Sunteți pe pagina 1din 18

Department of Mechanical, Materials & Manufacturing Engineering

Material Models and Modes of Failure - MM4MMM


Convenor: Dr W Sun (Coates Building B68, w.sun@nottingham.ac.uk)

VISCO-PLASTICITY MODEL AND THERMAL MECHANICAL FATIGUE


1.

UNIFIED VISCO-PLASTICITY MODEL

1.1

The Basic Model

For situations where plasticity (including effects such as cyclic hardening/softening


(see Figure 1), etc.) and creep are occurring simultaneously, and where reverse
loading may also be occurring, unified constitutive models can be used. One of such
models is the Chaboche, unified, visco-plasticity model [1] which is capable of
representing such material behaviour.

(a)

ii

ii

iii

iii

(b)

Figure 1 Schematic representation of the possible cyclic behaviours: (a) A. initial non-linear hardening; B. secondary
linear: i. Hardening; ii. Saturation; iii. Softening; and (b) A. initial non-linear softening; B. secondary linear:
i. hardening; ii. Saturation; iii. softening.

Cyclic material behaviour can take many forms, as Figure 1 shows. The region A in
Figure 1(a) shows that materials subjected to cyclic loading can initially exhibit nonlinearly hardening (a non-linear increase in load carrying capacity,
, with cycle
number N). Alternatively, region A in Figure 1(b) shows that materials can soften in a
non-linear manner (a non-linear decrease in load carrying capacity,
, with cycle
number N). In both cases, a secondary cyclic behaviour is usually observed in the
form of either linear hardening (linear increase in load carrying capacity), saturation
(stabilisation of load carrying capacity) or linear softening (linear decrease in load
carrying capacity), as shown in regions B in Figures 1(a) and (b) by i, , ii and iii,
respectively. It is the cyclic behaviour up to this point which is described here. Beyond
this point is the failure of the material (shown in each case in Figure 1 as a non-linear
decrease in load carrying capacity). This decrease in load carrying capacity which
occurs in the later stages involves the initiation and growth of cracks.
The uniaxial form of the Chaboche, unified, visco-plasticity model is as follows:53

f
Z

sgn

(1)

1 x0

x 0,
sgnx 0
1 x 0

x x 0
x
0 x 0

where:

(2)
(3)

f Rk

and

(4)

The elastic domain is defined by f 0 and the inelastic domain by f 0 (i.e. equation
(4) is the yield function for the model).
The evolution of the R and

values are controlled by the following equations: bQ Rp


R
i C i ai p i p

(7)

v Zp

p d p

R k sgn E p
v

(5)
(6)
(8)
(9)
(10)

As can be seen from equations (5) to (8), the model variables, i.e., hardening
(isotropic, drag-stress, R , and kinematic, back-stress, ) and viscous-stress, v , are
dependent on the value of inelastic-strain, p , calculated from equation (1). is the
stress, E, k , b , Q , a 1 , C 1 , a 2 , C 2 , Z and n are temperature-dependent material
constants, p is the accumulated inelastic-strain and is the total-strain.
The model takes account of both kinematic hardening and isotropic hardening. Figures
2 shows (viewed on the -plane) the basic physical interpretation of both types of
hardening and the effect they have on the yield surface, when viewed in threedimensional principal stress space.

54

(a)

(b)
Figure 2 Schematic -plane representation of (a) isotropic hardening; and
(b) kinematic hardening.

Equation (8) defines the viscous-stress, which is of a similar form to the widely used
Norton creep law. Equation (9) shows that the variable, p , which is used in the
calculation of many of the other variables is the accumulation of all the tensile and
compressive inelastic-strain, p , accumulated within the material, during every cycle.
More recently, a modification [2] has been made to the model in order that the
secondary, linear, behaviour is incorporated into the model predictions. This has been
achieved by the inclusion of an additional term into the equation for isotropic
hardening, i.e. equation (5) is modified to the following:
R bQ R p Hp
(11)
The additional Hp term ensures that once the non-linear hardening has saturated (i.e.
the bQ R p term as well as the kinematic hardening/softening contribution are no
longer changing), hardening/softening can still occur in a linear fashion as shown in
regions B in Figures 1. The sign of H determines whether the cyclic linear behaviour is
hardening (positive) or softening (negative). Alternatively, if a zero value of H is used,
the secondary behaviour is that of linear saturation as shown by ii in regions B in
Figure 1.
The material model requires the identification of eleven material constants. In order to
do this, strain-controlled, fully-reversed ( R 1 ), saw-tooth waveforms can be used,
as shown in Figure 2(a). Figures 2(b) and (c) show a typical schematic representation
of a stress-strain loop resulting from this applied waveform, assuming that the strain
amplitude reached is such that the elastic limit of the material is exceeded and
significant plasticity occurs.

55

(b)

(a)

(c)

Figure 3 Schematic representations of (a) a fully-reversed ( R 1 ), saw-tooth, strain-controlled waveform; (b) first
stress-strain loop resulting from this saw-tooth waveform and (c) stabilised stress-strain loop resulting from this sawtooth waveform.

t
(a)

(b)

(c)

Figure 4 Schematic representation of (a) an adapted saw-tooth waveform to include a dwell period at max ; (b) a
stress-strain loop including a maximum strain dwell period, and (c) creep relaxation behaviour during the maximum
strain dwell period.

For components in which stress relaxation may occur during a loading cycle, data from
strain-controlled, fully-reversed ( R 1 ) tests with constant strain dwells is also often
used; a typical dwell-type is shown in Figure 4(a). Schematic diagrams of the typical
behaviour observed in cycles with constant strain dwells are shown in Figures 4(b) and
(c).
1.2

Determination of Material Constants for the Visco-plasticity Model

The accurate determination of the material properties in the visco-plasticity models


require a complex procedure [3]. For example, in the case of the Chaboche unified
viscoplasticity model, Eqns. (1) to (11), firstly, a step by step procedure was used to
estimate the initial values of the parameters using the experimental results. These
initial values are then used to obtain an optimised set of material constants in a
simultaneous parameter optimisation routine based on two objective functions (stressstrain loops and the hardening/softening curve). In total, the material model requires
the identification of 10 material constants at a given temperature. The procedure for
the determination of the initial material constants from experimental data as well as
the subsequent optimisation procedure for the determination of the final material
constants have been described previously [4]. Temperature dependency will be
obtained by fitting the material properties derived from a number of constant
temperatures using iso-thermal test data [3]. Detailed procedure for determining the
material properties for visco-plasticity models is considered as beyond the scope of
this module, which can be seen in [3,4].
56

1.3

Multi-axial Formulations

In an elasto-viscoplasticity material behaviour, the total strain


elastic strain tensor
and the irreversible strain tensor .

, is the sum of the

The elastic strain is calculated using the linear elastic law

The plastic flow rule is a multiple of the plastic normal direction n and the plastic
multiplier

is the effective stress

The normal is presented as

Noted that the normal n and kinematic hardening X are both deviators. The isotropic
hardening law is
The kinematic hardening law is

From the above equation, it can be seen that the only unknown is the plastic multiplier
dp. And the key problem to solve the set of equations is to calculate dp.
In the viscoplastic condition, the yield function equates to the viscous stress, which
can be expressed by the Nortons law.

The nortons law

57

2.

THERMO-MECHANICAL FATIGUE

2.1

Introduction

Thermo mechanical fatigue (TMF) is a complex subject because of the interaction of


several failure modes including fatigue, oxidation and creep with a wide variety of
complex thermal and mechanical loads. Many models exist. In addition, the material
properties needed for these calculations are difficult and very expensive to obtain
experimentally.
TMF is caused by combined thermal and mechanical loading where both the stresses
and temperatures vary with time. This type of loading can be more damaging by more
than an order of magnitude compared with isothermal fatigue at the maximum
operating temperature. Material properties, mechanical strain range, strain rate,
temperature, and the phasing between temperature and mechanical strain all play a
role in the type of damage formed in the material. These types of loadings are most
frequently found in start-up and shut-down cycles of high temperature components
and equipment. Typically design lives are a few thousand cycles and involve significant
plastic strains.
One of the major differences between isothermal and thermal mechanical fatigue is
constraint. When heated, structures develop thermal gradients as they expand.
Expansion near stress concentrators is often constrained by the surrounding cooler
material. In this case thermal strain is converted into mechanical strain which causes
fatigue damage in the structure. Total constraint exists when all of the thermal strain
is converted into mechanical strain. Over constraint can occur in a stress concentration
where the mechanical strain is greater than the thermal strain. One measure of the
degree of constraint is the ratio of the thermal and mechanical strain rates. TMF
loading is often described to be in-phase (IP) or out-of-phase (OP). A schematic
illustration of the two loading situations is given in Figure 5.

(a)

(b)

Figure 5 Schematic representation of (a) in-phase (IP); (b) out-of-phase (OP).

58

A schematic illustration of the stress-strain responses under these two loadings is


given in Figure 6 [5]. In IP loading the maximum temperature and strain occur at the
same time. In OP loading, the material experiences compression at highest
temperature and tension at lower temperatures. IP loading is more likely to cause
creep damage during tensile stresses at high temperatures. OP loading is more likely
to cause oxidation damage because an oxide film can form in compression at the
higher temperature and then rupture during the subsequent low temperature tensile
portion of the loading cycle where the oxide film is more brittle.

Figure 6 Stress-strain loops for in-phase (IP) and out-of-phase (OP) loadings.

There are many active mechanisms in the TMF process. For discussion it is convenient
to consider damage from three primary sources: fatigue, oxidation and creep [6].
Damage from each process is summed to obtain an estimate of the total fatigue life,
Nf .

1
1
1
1
fatigue creep oxidation
Nf
Nf
Nf
Nf

(12)

Frequently one of the damage mechanisms is dominant. From a modelling prospective,


this suggests that the individual damage models and their associated material
properties must be accurate only for those conditions where the life is dominated by
that failure mechanism.
2.2

Thermal-Mechanical Fatigue Analysis of Power Plant Steels and Pipes (A


Case Study)

2.2.1 Introduction
Increasing temperatures and pressures for increased efficiency and reduced CO 2
emissions has become an ongoing trend for power generation plants. New advanced
materials that allow for significant increases in operating temperature are essential for
this need. Due to the intermittent nature of renewable energy generation,
conventional power generation plants are now subjected to a higher frequency of
59

thermo-mechanical cycling, demanded by flexible operation. This indicates that more


attention should be given to the problem of thermo-mechanical fatigue in component
life assessment and in the design of new plants.
The need for the power generation industry to improve the thermal efficiency of power
plant has led to the development of 9-12%Cr martensitic steels. The research on P91
and P92 steels started in the late 1970s and the early 1990s, respectively. Most of
research on these materials has focussed on their high temperature creep strengths.
Recently, the introduction of more cyclic operation of power plant has introduced the
possibility of fatigue problems. Bore cracking, e.g. Figure 7, due to the effects of
varying steam warming has been reported. The temperature cycling causes thermal
gradients between the inside and outside of components and this can cause cyclic
stress levels to be of concern.

Figure 7 Bore crack of a pipe weld caused by primarily downshock thermal fatigue.

The behaviour of power plant components can be simulated using the finite element
(FE) method. For example, the stresses in and failure behaviour of pipe weldments
under creep condition have been extensively studied. The accuracy of the modelling
work depends on the capability of the material constitutive equations for the
conditions which include cyclic mechanical loading and temperatures. Viscoplasticity
models are preferred due to their ability to represent both cyclic plasticity and viscous
creep behaviour. This type of model is widely used in aeroengine application with
nickel base alloys. The material constants can be determined from isothermal tests
and the resulting model can be used in anisothermal situations. However,
viscoplasticity models are relatively unused for representing the behaviour of power
plant materials. Experimental tests under cyclic loading conditions have been
performed [7] in order to improve the understanding of the behaviour of 9-12%Cr
steels under thermal fatigue or thermal-mechanical fatigue conditions.
2.2.2 Experimental investigation
The P91 and P92 materials
P91 and P92 are typical examples of the type of creep resistant steels, used in power
plant superheater, reheater tubes, headers and often in high temperatures steam
pipework. In general, P91 steel consists of 9% chromium and 1% molybdenum while
P92 steel consists of 9% chromium, 1.75% tungsten and 0.5% molybdenum. Both of
60

these steels are of the ferritic/martensitic types, which are commonly grouped as 912%Cr steels. Ferritic/martensitic type steel has a lower coefficient of thermal
expansion, compared to an austenitic type steel, which is an advantage when dealing
with thermal fatigue problems. The P92 steel has better creep strength by
approximately 10-20% over the P91 steel; thus, P92 pipe wall thicknesses can be
reduced, resulting in improved behaviour under thermal fatigue situations. The
chemical compositions, in accordance with ASTM standard, of P91 and P92 steels are
given in Table 1.
Table 1 Chemical compositions of the P91 and P92 steels (wt%)

P91
P92
P91
P92

Cr

Mo

Si

8.60

1.02

0.12

0.34

8.62
Al
0.00
7
0.01
9

0.33
V

0.10
Nb
0.07
0
0.07
6

0.45
N
0.06
0
0.04
7

0.24
0.21

S
<0.00
2
0.002
W
0.03
1.86

P
0.017
0.015
B
0.003
4

Low cycle iso-thermal and TMF tests


An Instron 8862 TMF test machine, which utilizes radio-frequency induction heating,
was used to perform all experiments. The coil design enables the temperature
gradient along the gauge section to be controlled to within 10oC of the target
temperature. The maximum achievable load for the machine is 35kN. The machine is
controlled by a servo electric screw driven actuator. Strains were measured using
high temperature extensometers with a gauge length of 12.5mm.
The test specimens were machined from P91 and P92 steam pipe sections. Figure 8
shows the dimensions of the cylindrical specimens used for the uniaxial tests. The
plain section of the specimens is 15mm in length and 6.5mm in diameter; these were
finished by fine machining and polishing to an average roughness value of 0.8m.

Figure 8 Uniaxial LCF/TMF specimen.

61

0.6

0.6

0.4

0.4

0.2
0.0
-0.2
-0.4

10

15

20

Strain, (%)

Strain, (%)

Fully-reversed isothermal and aniso-thermal tests were conducted for both materials,
under strain control conditions, with a strain rate of 0.1%/s, for a range of
temperatures and strain amplitudes. The cyclic period is 20s for the continuous strain
cycling tests, see Figures 9. The tests were carried out to failure, defined as the
number of cycles at which a 30 percent drop in the maximum stress occurs. Tests with
dwell periods at the tensile peak strains for times of 2 and 5 minutes were also applied
to P91 and P92 specimens, respectively, in order to get stress relaxation data for the
materials.
The strain amplitudes and strain rates were maintained constant
throughout the tests. All of the tests performed under an-isothermal conditions were
carried out with either in-phase (IP) or out-of-phase (OP) conditions [7].

= 0.5%
-rate = 0.1%/s

-0.6
(a)

0.2
0.0
-0.2

30

60

90

120

= 0.5%
Hold = 2 minutes

-0.4
-0.6

Time (seconds)

(b)

Time (seconds)

Figure 9 Strain histories in a cycle: (a) Saw-tooth type waveforms;


(b) Saw-tooth type waveforms with a strain holding dwell at max.

2.2.3 Material behaviour model and properties


For the P91 and P92 power plant steels, the Chaboche, unified, viscoplasticity model is
used to describe the visco-plasticity behaviour and in the finite element modelling [79]. The material constants for the Chaboche model for the P91 steel, determined using
isothermal test data, are presented in Table 2.
Table 2 The material constants for viscoplasticity model of P91 steel.
Temp.
(oC)
400
500
600

E
(MPa)
197537.
0
181321.
6
139395.
2

k
Q
(MPa) (MPa)

a1
(MPa)

96

-55.0

0.45

150.0

90

-60.0

0.6

98.5

85

-75.4

1.0

52.0

C1
2350.
0
2191.
6
2055.
0

a2
(MPa)
120.0
104.7
67.3

C2
405.
0
460.
7
463.
0

Z
(MPa.s1/n)

2000

2.25

1875

2.55

1750

2.7

2.2.4 Low cyclic fatigue behaviour of P91 and P92 steels


62

Cyclic stress-strain behaviour under iso-thermal and TMF conditions


Strain controlled, iso-thermal and aniso-thermal low cycle fatigue tests have been
performed on the P91 and P92 steels with temperatures in the range 400 to 675 oC,
with strain ranges,
0.2 to 0.6% [7]. Full cyclic stress-strain responses
have been used to derive the material constants in the visco-plasticity models using an
optimization procedure. Figures 10 show examples of the stress-strain loops for the
tests of the P91 and P92 steels, at 600oC, with holding dwells, during which significant
stress relaxation can be seen. Figure 11 shows an example of the comparison of the
experimental and the simulated stress-time curves using the unified model for the P91
steel at 600oC, from which high accuracy of the predictive capability from the viscoplasticity model can be seen. Temperature dependencies have been derived using the
iso-thermal material properties in the tested temperature range. This has been used to
simulate the TMF tests in order to validate the predictive capability of the viscoplasticity models under aniso-thermal conditions. An example of this is given in Figure
12.

63

400.0

300.0

300.0

200.0

200.0

100.0

100.0

0.0
-0.006 -0.004 -0.002 0.000 0.002 0.004 0.006
-100.0

MPa

MPa

400.0

0.0
-0.006 -0.004 -0.002 0.000 0.002 0.004 0.006
-100.0

-200.0

-200.0

-300.0

-300.0

-400.0
Strain

-400.0
Strain

Figure 10(a) Stress-strain loop (straincontrolled = 0.5%, dwell = 2 mins, N = 1)


for P91 at 600C.

Figure 10(b) Stress-strain loop (straincontrolled, = 0.5%, dwell = 5 mins, N


= 1) for P92 at 600C.

300

500 C

400.0

200
200.0

0
6400

6500

6600

6700

6800

-100
-200

Exp
Simulated

(MPa)

(MPa)

100

0.0
-0.004

-0.002

0.000
-200.0

675 C

0.002

0.004

Exp
Simulated

-400.0

-300
S train

Figure 11 Experimental and simulated stresstime curves from strain controlled test for P91
o
at 600 C ( = 0.5%, dwell = 2 mins).

S train

Figure 12 Stress-strain loop obtained from strain


controlled TMF test for P92 (out-of-phase, 500675C, N = 1).

Cyclic isotropic softening


During strain range controlled, isothermal, low cycle fatigue tests, for both P91 and
P92 steels, a distinct cyclic softening behaviour has been observed if specimens are
taken to failure [7]. The stress range, /2 (defined as half the difference between the
maximum and minimum stresses in a cycle, found at the end points of the tensile and
compressive loadings, respectively), evolved as load cycles progressed. In general,
materials initially exhibit a rapid non-linear cyclic softening (stage I), which is followed
by a relatively long period of linear softening (stage II). Ultimately, damage will
accumulate and micro-cracks will be formed, in the stage III softening, which
eventually leads to failure, as illustrated, schematically, in Figure 13(a). A typical
example, for the P91 steel test at 600oC ( = 0.5%) is shown in Figure 13(b).
64

320

max

No dwell
2 mins dwell

290

/ 2 (MPa)

Stage I

260

a c
d

230
200

Stage II

Stage III
N
Nf

170
0

200

400

600

800

S train

Figure 13(a) Schematic illustration of cyclic softening


of an isothermal low cycle fatigue test.

Figure 13(b) Cyclic isotropic softening of the P91


steel at 600oC.

Material behaviour models capable of describing stage I and stage II have been
developed [e.g. 7]. The development of material models which can accurately describe
the full life cycle, including stage III is currently ongoing. The need for such models
and the associated numerical procedures are extremely important, if accurate
prediction of the component life under low cycle thermal mechanical fatigue and multiaxial loading conditions is to be achieved.
Microstructural degradation
Detailed microstructural investigations of test pieces taken from the isothermal tests
have been performed using transmission electron microscopy (TEM) and scanning
electron microscopy (SEM). These investigations were performed at various life
fractions in order to establish the microstructure evolution that takes place throughout
the life of a test piece.
Microstructural evolution of the P91 steel, under cyclic loading, seems to occur on a
sub-grain scale [7]. Figure 14 shows the results for the as-received material as well
as for the interrupted test pieces at cycle numbers 200, 400 and 656 (corresponding
to points (a), (b), (c) and (d), in Fig. 7(b), respectively). Based on the bright field TEM
images, the sub-grain sizes, at these various stages, are 0.383, 0.507, 0.551 and
0.604 m, respectively. It can be seen that the sub-grains are coarsened with cycle
number, particularly within the initial 200 cycles (also, the stress amplitude decreases
non-linearly before achieving a stabilisation stage at around the 200th cycle). It is
difficult to clearly identify the sub-grain evolution, at different life fractions, using SEM.
However, the SEM images show that a small number of cracks start to develop as the
softening curve begins to accelerate as the test piece begins to fail, see Figure 14(e).
Up to this point, it has been found that the value of the cyclic Youngs modulus in each
cycle is similar to the initial Youngs modulus; the modulus value is an indirect
measurement of damage. Transgranular cracks were observed at many locations on
the test piece which ran to failure (at 656 cycles). Between the 400 th and the 656th
cycle, the Youngs modulus values and the stress amplitudes both decreased. The
65

accelerated cyclic softening behaviour, in the final stage of cyclic loading, is associated
with the propagation of cracks within the material.

Figure 14 Bright field TEM images for (a) the as-received material, and the sub-grain evolution
which occurs in 0.5% strain-controlled test at 600C for cycles (b) 200, (c) 400 and (d) 656.
Also, (e) a SEM image of a crack which initiated on the surface of a specimen at 400cycle.

2.2.5 Thermal-mechanical response of pipes


Finite element simulation
A coupled, temperature-displacement, axisymmetric pipe model has been used to
predict the thermal-mechanical responses of pipes, as shown in Figure 15.

Figure 15 Axisymmetric finite element pipe model.

Representative thermal and mechanical loading


A more realistic material model for thermal-mechanical fatigue behaviour is the
Chaboche unified viscoplasticity model which was discussed earlier. In this subsection, FE results obtained for a realistic main steam pipe (ri = 0.126m, ro = 0.162m)
with representative loading, are presented. User defined code (UMAT with ABAQUS) or
the Z-mat software (licensed by Transvalor/ENSMP) can be used to implement the
multiaxial form of the Chaboche constitutive equations into the ABAQUS FE code [8].
The material constants are based on those for P91 steel found in Saad et al. [7]. A
representative power plant steam temperature and pressure loading cycle, defined in
Figure 16, was applied to the inner surface of the pipe. Figure 17 shows the
temperature history generated at the inner and outer pipe wall surfaces. The axial
stress histories, presented in Figure 18, follow the trend of steam temperature, with
the inner pipe surface mostly in compression and the outer pipe surface mostly in
tension, until the end of the cycle. This is because in this case, the inner surface
66

always remains significantly hotter than the outer surface, except right at the end of
the cycle. It should be noted that the hoop stress follows a similar trend to that of the
axial stress and so is not presented here.

Figure 16 Representative steam temperature and


pressure loading cycle for power plant pipes.

Figure 17 Inner and outer pipe wall surface temperature histories.

Figure 18 Inner and outer pipe surface axial stress histories.

67

Cyclic stress-strain responses of a plain pipe


Figure 19 shows the axial strain history obtained from the modelling of the pipe over
the course of the first four cycles, with loading of the form given in Figure 16. As
expected, the axial strain history follows the trend of the steam temperature. The
quite complex axial stress-axial strain loops, obtained for the inner and outer pipe
surfaces, during the four cycles are plotted in Figure 20.

Figure 19 Pipe slice axial strain history.

Figure 20 Inner and outer pipe surface axial stress-strain loops.

An energy-based life prediction method


Such stress-strain loops shown in Figure 20 may be used to predict the fatigue life of
the pipe using an energy-based method. Since the hysteresis loop area represents the
plastic strain energy dissipated, or plastic work done during a given cycle, it can be
used as the basis for quantifying the fatigue damage that has occurred during each
cycle. For example, for a stabilised cycle, the plastic work done (hysteresis loop area),
WPd , can be correlated with the number of cycles to failure, N, via:
68

WPd = AN

(13)

where A and are material constants. Alternatively, the accumulated plastic work
done may be used. In multiaxial cases, the plastic strain energy dissipated per cycle
can be determined from the sum of the hysteresis loop areas associated with all nine
tensor components (Figure 21).

Figure 21 Dissipated plastic strain energy lifing approach.

2.2.6 Discussion
Existing and new power plants are faced with new challenges to deliver energy more
securely and more efficiently, due to the rise of energy demand, required reduction in
specific emissions and the increasing deployment of renewable energy. Power plants
have to operate more efficiently and yet more flexibly, to compensate for the irregular
supply from renewable sources. Hence, advanced materials are required and these
need robust characterisation for extremely severe long-term operating conditions (e.g.
creep, fatigue, thermo-mechanical fatigue, oxidation and their interactions, at high
temperature). Due to the change of operating conditions of power generation plants,
the traditional design and life prediction methods or procedures, based on static load
situations, may not be sufficient. High temperature structural integrity assessment,
therefore, should take into account the severe, thermal-mechanical loading conditions,
which exist. In particular, the TMF behaviour of components with material and
geometry discontinuities, such as occurs in welds, branched pipes and pipe bends
must be taken into account.
This work provides useful input to the future application of thermal-mechanical
analyses of power plant components for the purpose of component life assessment.
For example, the high level of axial stress and strain concentrations, experienced at
the inner surface of a pipe, due to cyclic thermal and mechanical loads, gives an
indication of the potential crack initiation sites in pipes. The behaviour may become
more significant for the case of a pipe with a circumferential weld; the heterogeneous
nature of welds makes structures containing welds more vulnerable to thermalmechanical fatigue failure.
Further work is in progress to improve the accuracy and capability of visco-plasticity
modelling for the prediction of thermal-mechanical fatigue of power plant materials
and components. In particular, improved optimization procedures will be developed
69

and used in order to produce more accurate and more efficient methods for
determining the full set of the temperature dependent material properties. In addition,
the prediction of cyclic softening behaviour will be improved to allow the models to be
extended for the prediction to include crack growth stage. The current model limits the
stress prediction to only about two thirds of the number of cycles to failure. Also,
further research will be carried out to more accurately predict the stress relaxation
behaviour and to predict all stages of the cyclic softening behaviour, through a more
detailed understanding of cyclic softening mechanisms related to micro-crack/damage
formation and growth. New life prediction methodologies, such as those based on
plastic strain energy approaches, need to be established and validated.
REFERENCES
[1]

Chaboche, J. L. and Rousselier, G. On the Plastic and Viscoplastic Constitutive equations - Part 1: Rules
Developed with Internal Variable Concept. J. Pressure Vessel Technology 105, 153-158, 1983.

[2]

Lemaitre J, Chaboche JL. Mechanics of Solid Materials. Cambridge: Cambridge University Press, 2000.

[3]

Tong, J., Zhan, Z. L. and Vermeulen, B. Modelling of cyclic plasticity and viscoplasticity of a nickel-based alloy
using Chaboche constitutive equations. Int. J. Fatigue 26 (8), 829837, 2004.

[4]

Gong Yunpeng, Hyde C. J., Sun W. and Hyde T. H. Determination of material properties in the Chaboche
unified viscoplasticty model. J. Materials: Design & Applications 224 (1), 19-29, 2010.

[5]

Jaske, C.E. Thermal Fatigue of Materials and Components, ASTM STP 612, 1976, pp. 170-198.

[6]

Sehitoglu, H. Advances in Fatigue lifetime Predictive Techniques, ASTM STP 1122, 1992, pp. 47-76.

[7]

Hyde C. J., Sun W., Hyde T. H., Saad A. A. Thermo-mechanical fatigue testing and simulation using a
viscoplasticty model. J. Computational Materials Science 56, 29-33, April 2012.

[8]

Tanner D. W. J., Sun W. and Hyde T. H. FE analysis of a notched bar under thermo-mechanical fatigue using a
unified viscoplasticity model. Procedia Engineering 10, 1081-1086, 2011 (Elsevier).

[9]

Sun W., Tanner D. W. J., Hyde T. H. and Saad A. A. Thermal-mechanical fatigue behaviour of 9-12% Cr power
plant steels and pipes. 2nd SuperGen Int. Conf. 8-9 September 2012, Hangzhou China.

70

S-ar putea să vă placă și