Sunteți pe pagina 1din 12

Long-term Creep Strength of 9-12Cr Creep Strength Enhanced Ferritic

Steels

K. Kimura, K. Sawada and H. Kushima

National Institute for Materials Science, 1-2-1 Sengen, Tsukuba, Ibaraki 305-0047,
Japan

ABSTRACT

Creep deformation property of 9-12Cr ferritic creep resistant steels was investigated. With
decrease in stress, a magnitude of creep strain at the onset of accelerating creep stage
decreased from about 2% in the high-stress regime to less than 1% in the low-stress regime. A
time to 1% total strain was observed in the transient creep stage in the high-stress regime,
however, it shifted to the accelerating creep stage in the low-stress regime. Life fraction of the
times to 1% creep strain and 1% total strain tended to increase in the low-stress regime.
Difference in stress dependence of the minimum creep rate was observed in the high- and
low-stress regimes with a boundary condition of 50% of 0.2% offset yield stress. Stress
dependence of the minimum creep rate in the high-stress regime was equivalent to a strain
rate dependence of the flow stress evaluated by tensile test, and a magnitude of stress
exponent, n, in the high-stress regime decreased with increase in temperature from 20 at
550°C to 10 at 700°C. On the other hand, n value in the low-stress regime was about 5, and
creep deformation in the low stress regime was considered to be controlled by dislocation
climb. Creep rupture life was accurately predicted by a region splitting analysis method by
considering a change in stress dependence of creep deformation property.

1. Introduction

In order to harmonize increase in demand of electrical energy and global environment,


improvement of energy efficiency of thermal power plant is one of the key issues. Creep
strength enhanced ferritic (CSEF) steels have been widely used for high temperature
structural components of Ultra Supercritical (USC) thermal power plant and energy efficiency
of the plant has been improved by increasing steam temperature and pressure. However,
unexpected drop of the long-term creep rupture strength from the predicted value extrapolated
from the short-term creep rupture strength was observed on the CSEF steels. High
temperature equipments are designed according to an allowable stress which is determined by
long-term creep strength properties of creep deformation and creep rupture. Consequently,
accurate evaluation of long-term creep strength is considered to be a most important issue to
ensure safety and reliability of those components.
In order to establish a reliable and accurate evaluation method for long-term creep strength of
CSEF steels, several new approaches have been investigated [1-5]. A region splitting analysis
method proposed by Kimura et al. [1-3] evaluates creep strength independently for high- and
low-stress regimes divided by 50% of 0.2% offset yield stress under a nominal strain rate of
5×10-5 s-1. Multi region analysis of creep rupture data in consideration of change in activation
energy for creep rupture life was proposed by Maruyama et al. [4]. A rationalization and
extrapolation method of creep fracture data based on relationships which involve the
activation energy for matrix diffusion and the ultimate tensile stress values at the creep
temperature was proposed by Wilshire et al. [5]. Influence of stress on creep rupture life is
considered in a region splitting analysis method [1-3] and Wilshire’s proposal [5]. Creep
deformation behaviour is affected by microstructural change during creep exposure and a
stress influences on microstructural change. In this paper, creep deformation property of
several CSEF steels is investigated and stress dependence of the creep deformation property is
discussed.

2. Experimental Procedure

Three type creep strength enhanced ferritic (CSEF) steels of ASME Grades T91, P92 and
P122 were used. Chemical composition and heat treatment condition of the steels are shown
in Tables 1 and 2. Creep test was conducted over a range of temperatures from 500 to 700°C.
The minimum creep rate of the unbroken creep test was also used for the investigation.
Tensile test was conducted under a constant nominal strain rate of 5×10-5 s-1 up to about 2%
of total strain, and that was increased to 1.25×10-3 s-1 and a tensile strength was evaluated.
Flow stress was evaluated under a constant strain rate of 5×10-5 s-1, as well as 0% and 0.2%
offset yield stresses, in addition to tensile strength that is regarded as a flow stress under a
constant nominal strain rate of 1.25×10-3 s-1.

Table 1 Chemical composition (mass%) of the steels studied.


Steel C Si Mn Ni Cr Mo Cu W V Nb Al B N
T91 0.09 0.29 0.35 0.28 8.79 0.90 0.032 - 0.22 0.072 0.001 - 0.044
P92 0.11 0.10 0.41 0.17 9.26 0.42 - 1.67 0.16 0.057 0.01 0.002 0.0462
P122 0.12 0.30 0.60 0.32 10.65 0.34 0.85 1.89 0.19 0.05 0.007 0.0029 0.054

Table 2 Heat treatment condition of the steels studied.


Steel Normalizing Tempering
T91 1050°C × 10min / Air cooling 765°C × 30min / Air cooling
P92 1070°C × 120min / Air cooling 780°C × 120min / Air cooling
P122 1050°C × 60min / Air cooling 770°C × 360min / Air cooling

3. Results and Discussion


3.1 Stress dependence of creep deformation

Stress vs. time to rupture curves over a range of temperatures from 500 to 700°C of T91 steel
are shown in Figure 1. A rectilinear relationship is observed at 500°C, however, slope of the
curves becomes steeper in the long-term at 550°C and above. This change in slope of the
relation makes difficult to predict long-term creep strength accurately. Degradation due to
microstructural change during long-term creep exposure reduces creep strength of the steel,
and it may cause the curve to be steeper.

Figure 1 Stress versus time to rupture curves of the T91 steel.

Creep rate vs. time curves of T91 steel at 550 and 600°C are shown in Figure 2 (a) and (b),
respectively. Creep deformation of the steel consists of transient and tertiary creep stages and
no obvious steady state creep stage is observed at both temperatures. Good linear relationship
between creep rate and time is recognized in the transient creep stage at both temperatures in
a double logarithmic plot. With decrease in stress, minimum creep rate decreases and creep
rupture life increases, and feature of the creep rate vs. time curve is essentially the same
regardless of stress.

Creep rate vs. creep strain curves of T91 steel at 550 and 600°C are shown in Figure 3 (a) and
(b). No obvious influence of stress on creep deformation is observed in the creep rate vs. time
curve (Fig.2), however, stress dependence of creep deformation property is clearly observed
on creep rate vs. creep strain curve. In the high stress regime, creep rate indicates minimum
value at a creep strain of 0.02 to 0.03, however, a minimum creep rate is observed at a smaller
creep strain of less than 0.01 at the low stress regimes. Creep strain where a creep rate shows
minimum value decreases with decrease in stress. Stress dependence of the onset creep strain
of tertiary creep stage should be taken into account for an evaluation of long-term creep
strength property, because design of high temperature structural components is controlled by
not only creep rupture strength, but also creep deformation property.
Figure 2 Creep rate versus time curves of the T91 steel at (a) 550°C and (b) 600°C.

Figure 3 Creep rate versus creep strain curves of the T91 steel at (a) 550°C and (b) 600°C.
According to the rules for construction of nuclear power plant components regulated in
ASME Boiler and Pressure Vessel Code, Section III, Division 1, Subsection NH [6], a
temperature and time-dependent stress intensity limit, St value is determined for each specific
time by the lesser of the followings.

• 100% of the average stress required to obtain a total strain of 1%


• 80% of the minimum stress to cause initiation of tertiary creep
• 67% of the minimum stress to cause creep rupture

Stress vs. times to 1% total strain, 1% creep strain, initiation of tertiary creep and time to
rupture at 550 and 600°C of T91 steel are shown in Figure 4 (a) and (b). At the highest stress
condition, time to initiation of tertiary creep is about 10 times longer than times to 1% total
strain and 1% creep strain, however, difference between those parameters decreases with
decrease in stress.

Figure 4 Stress versus times to 1% total strain, 1% creep strain, initiation of tertiary creep and
time to rupture of T91 steel at (a) 550°C and (b) 600°C.
Changes in life fraction of times to 1% total strain, 1% creep strain and initiation of tertiary
creep with increase in time to rupture at 550 and 600°C of T91 steel are shown in Figure 5.
Life fraction of time to initiation of tertiary creep is within a range of 0.5 to 0.7, independent
of time to rupture. Life fraction of times to 1% total strain and 1% creep strain is smaller than
10% of time to rupture in the short-term, however, it increases with increase in time to rupture
and approaches to almost the same level as that of the initiation of tertiary. It indicates that an
initiation of tertiary creep is more important parameter for the stress intensity limit, St, than
time to a total strain of 1%, since 80% of the minimum stress to cause initiation of tertiary
creep is definitely smaller than 100% of the average stress required to obtain a total strain of
1% in the long-term. Increase in life fraction of the times to 1% total strain and 1% creep
strain in the long-term should be derived from the onset of tertiary creep stage within a
smaller strain than that in the short-term, as shown in Fig.3.

Figure 5 Changes in life fraction of times to 1% total strain, 1% creep strain and initiation of
tertiary creep with increase in time to rupture of T91 steel at (a) 550°C and (b) 600°C.
3.2 Stress dependence of minimum creep rate

The minimum creep rate of P92 and P122 steels are plotted against stress and shown in
Figures 6 and 7. Tensile strength and flow stress under a constant strain rate of 5×10-5 s-1 are
also plotted in the same figure. Numerical values in the figure indicate the stress exponent
value, n of the power law. The stress dependence of the minimum creep rate is clearly divided
into two regimes. The magnitude of the stress exponent, n, in the high stress regime is larger
than that in the low stress regime, and it decreases with increase in temperature from about 20
at 550°C to about 10 at 700°C. On the other hand, the n value in the low stress regime is
smaller and in the range of 4 to 7. Moreover, tensile strength and flow stress at a strain rate of
5×10-5 s-1 are located on the linearly extrapolated lines from the stress dependence of
minimum creep rate in the high-stress regime. The stress dependence of minimum creep rate
in the high stress regime is equivalent to the strain rate dependence of flow stress.

Figure 6 Stress versus minimum creep rate curves of P92 steel. Tensile strength and flow
stress are plotted on the corresponding strain rate. Numerical values indicate the stress
exponent, n of the power law.
Figure 7 Stress versus minimum creep rate curves of P122 steel. Tensile strength and flow
stress are plotted on the corresponding strain rate. Numerical values indicate the stress
exponent, n of the power law.

According to a change in stress dependence of the minimum creep rate, a boundary stress
between high- and low-stress regimes is evaluated. 0% offset yield stress and 50% of 0.2%
offset yield stress are plotted against a boundary stress and shown in Figure 8 (a) and (b),
respectively. A magnitude of 0% offset yield stress is plotted with its error of measurement of
about ±10MPa. There is a good correspondence between 0% offset yield stress and the
boundary stress between high- and low-stress regimes, as shown in Fig. 8(a). It indicates that
the low-stress regime is equivalent to an elastic range below the proportional limit, and the
high-stress regime corresponds to a plastic range. Consequently, the large stress dependence
of the minimum creep rate in the high-stress regime should be caused by a contribution of
plastic deformation. It should also be a cause of large creep strain in the transient creep stage
in the high-stress regime as shown in Fig. 3 for T91 steel. On the other hand, creep
deformation in the low-stress regime is considered to be governed by diffusion controlled
phenomena. In addition to 0% offset yield stress, good correspondence between 50% of 0.2%
offset yield stress and the boundary stress between high- and low-stress regimes is observed
for both P92 and P122 steels, as shown in Fig. 8(b). It indicates that 50% of 0.2% offset yield
stress is regarded to be a boundary stress between high- and low-stress regimes where the
magnitude of stress exponent changes, and it is empirically corresponds to 0% offset yield
stress that is a proportional limit stress of the steels.
Figure 8 Relations between (a) 0% offset yield stress, (b) 50% of 0.2% offset yield stress and
a boundary stress between high stress and low stress regimes where the magnitude of stress
exponent changes.

3.3 Region splitting analysis of creep rupture life

The stress dependence of the minimum creep rate changes at the stress corresponding to a
proportional limit, that is empirically equivalent to 50% of 0.2% offset yield stress. According
to the above observation, the creep rupture data of P92 and P122 were divided into two
groups of high stress and low stress regimes with a boundary condition of 50% of 0.2% offset
yield stress, and the creep rupture life of both regimes were individually analyzed by a
Larson-Miller parameter with a quadratic function of logarithm stress. This is a creep rupture
life assessment method proposed by Kimura et al. as a Region Splitting Analysis [1-3]. The
evaluated creep rupture life curves by means of the above method are described with the data
plot in Figure 9 (a) and (b) for P92 and P122 steels, respectively. The creep rupture life data
of both steels are precisely described by the predicted creep rupture life curves.

Comparison of the predicted and observed creep rupture lives is shown in Figure 10 (a) and
(b) for P92 and P122 steels, respectively. Creep rupture life in the high- and low-stress
regimes is shown by open and solid symbols, respectively. Good correspondence between
predicted and observed creep rupture lives is observed from short-term to long-term. Accurate
life prediction by means of a region splitting analysis in consideration of 50% of 0.2% offset
yield stress is confirmed on the steels. In order to evaluate creep rupture strength of the CSEF
steels, it should be analyzed on high- and low-stress regimes independently with a boundary
of 50% of 0.2% offset yield stress. Evaluation of long-term creep strength of the CSEF steels
should be conducted on the creep test data in the low stress regime, since it is essentially
different from the creep strength in the high stress regime, and the high temperature
components are operated under stress condition below the proportional limit stress.
Figure 9 Predicted creep rupture life by means of a region splitting method of (a) P92 and (b)
P122 steels.

Figure 10 Comparison of observed and predicted creep rupture lives of (a) P92 and (b) P122
steels.
4. Conclusions

Creep deformation property and its stress dependence were investigated on the several creep
strength enhanced ferritic steels of T91, P92 and P122, and a reliable life prediction method
was discussed.

1. Creep strain where a creep rate shows minimum value of T91 steel decreased with decrease
in stress. It was indicated that an initiation of tertiary creep should be more important
parameter for the stress intensity limit, St, than time to a total strain of 1%, since 80% of the
minimum stress to cause initiation of tertiary creep was definitely smaller than 100% of the
average stress required to obtain a total strain of 1% in the long-term.

2. The stress dependence of the minimum creep rate of P92 and P122 steels in the high-stress
regime was equivalent to the strain rate dependence of the flow stress. The value of the stress
exponent, n in the high-stress regime was larger than that in the low-stress regime. The
boundary stress where the magnitude of stress exponent changes corresponded to a
proportional limit stress which was empirically equivalent to 50% of 0.2% offset yield stress.

3. The large stress dependence of the minimum creep rate in the high stress regime was
considered to be caused by a contribution of considerable plastic deformation due to stresses
higher than the proportional limit. Creep deformation in the low stress regime was considered
to be governed by diffusion controlled phenomena and dislocation climb was a candidate rate
controlling mechanism.

4. It has been concluded that the long-term creep strength of creep strength enhanced ferritic
steels should be evaluated on the creep test data in the low stress regime below 50% of 0.2%
offset yield stress, since it is essentially different from the creep strength property in the high
stress regime.

Acknowledgement

A part of this study was financially supported by the Budget for Nuclear Research of the
Ministry of Education, Culture, Sports, Science and Technology, based on the screening and
counseling by the Atomic Energy Commission.

References

[1] K. Kimura, H. Kushima and F. Abe: „Degradation and Assessment of Long-term Creep
Strength of High Cr Ferritic Creep Resistant Steels”, EPRI Int. Conf. on Advances in Life
Assessment and Optimization of Fossil Power Plants, Orlando, USA, (2002).
[2] K. Kimura, H. Kushima and K. Sawada, „Long-term Creep Strength Prediction of High Cr
Ferritic Creep Resistant Steels Based on Degradation Mechanisms”, Engineering Issues in
Turbine Machinery, Power Plant and Renewables, 6th Int. Charles Parsons Turbine Conf.,
Dublin, Ireland, 443 – 456 (2003).
[3] K. Kimura, K. Sawada, K. Kubo and H. Kushima, „Influence of Stress on Degradation
and Life Prediction of High Strength Ferritic Steels”, PVP-Vol.476, ASME/JSME Pressure
Vessel and Piping Conf., San Diego, USA, 11-18 (2004).
[4] K. Maruyama and J.S. Lee, „Causes of Overestimation of Creep Rupture Strength in
11Cr-2W-0.3Mo-CuVNb Steel”, Creep & Fracture in High Temperature Components -
Design & Life Assessment Issues, ECCC Creep Conf., London, UK, 372-379 (2005).
[5] B. Wilshire and P.J. Scharning, „Long-term creep life prediction for a high chromium
steel”, Scripta Mat., Vol.56, 701-704 (2007).
[6] ASME Boiler and Pressure Vessel Code, Section III, Division 1, Subsection NH, (2007).

S-ar putea să vă placă și