Sunteți pe pagina 1din 6

Materials Science and Engineering A 510511 (2009) 5863

Contents lists available at ScienceDirect

Materials Science and Engineering A


journal homepage: www.elsevier.com/locate/msea

Long-term creep deformation property of modied 9Cr1Mo steel


K. Kimura , H. Kushima, K. Sawada
National Institute for Materials Science, 1-2-1, Sengen, Tsukuba, Ibaraki 305-0047, Japan

a r t i c l e

i n f o

Article history:
Received 9 January 2008
Received in revised form 10 March 2008
Accepted 30 April 2008
Keywords:
Creep deformation
Modied 9Cr1Mo steel
Creep data sheet
Transient creep
Accelerating creep

a b s t r a c t
The rst volume of Atlas of Creep Deformation Properties was published on modied 9Cr1Mo steels
in March 2007, as a part of the NIMS (National Institute for Materials Science) Creep Data Sheet series.
Creep deformation properties up to about 70,000 h have been investigated. No clear steady-state creep
stage has been observed, and creep deformation of the steel consists of transient and accelerating creep
stages. Good linear relationships between creep strain vs. time and creep rate vs. time were observed
within a transient stage in a loglog plot. It was appropriately expressed by a power law rather than an
exponential law, logarithmic law and Blackburns equation. With decrease in stress, the magnitude of
creep strain at the onset of accelerating creep stage decreased from about 2% in the short-term to less
than 1% in the long-term region. Life fraction of the time to specic strain of 1% creep strain and 1% total
strain, to time to rupture tended to increase with decrease in stress. The time to 1% total strain, that is
an important parameter for design of high temperature components, was observed to lie in the transient
creep stage in the short-term regime, however, it shifted to the accelerating creep stage in the long-term
regime. For evaluation of long-term creep strength properties, an experimental creep test data should be
extrapolated in consideration of the stress dependence of creep deformation properties.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Creep-strength-enhanced ferritic (CSEF) steels have been widely
used in a modern thermal power plant and those have contributed to improve energy efciency of the plant by means of
increasing steam temperature and pressure. Unexpected drop in
creep-rupture strength, however, was observed in the long-term
in comparison with the anticipated creep-rupture strength from
the short-term properties. In order to obtain a reliable estimation method of long-term creep strength, several new approaches
have been investigated on CSEF steels [15]. According to a reevaluation of long-term creep strength, the allowable tensile stress
value of several CSEF steels was reduced [610]. In order to better understand long-term creep strength properties of CSEF steels,
microstructural evolution and the degradation mechanism during creep exposure have been investigated. Precipitation and rapid
growth of Z-phase during creep exposure at elevated temperatures
has been found and many investigations have been conducted on
the Z-phase [1117]. This phase has been considered as one of the
potential causes of degradation, since growth of Z-phase consumes
many of the ne particles of MX carbonitride and reduces their
precipitation strengthening effect.

Corresponding author. Tel.: +81 29 859 2229; fax: +81 29 859 2201.
E-mail address: kimura.kazuhiro@nims.go.jp (K. Kimura).
0921-5093/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2008.04.095

It is useful to investigate creep deformation properties for the


evaluation of long-term creep strength property [18], since degradation due to microstructural change during creep exposure should
be reected on creep deformation. In March 2007, the rst volume of Atlas of Creep Deformation Properties on 9Cr1MoVNb
steel (ASME Grade 91), that is a typical CSEF steel was published
as a series of NIMS Creep Data Sheets [19]. In the present study,
creep deformation of 9Cr1MoVNb steels (ASME SA-213 T91)
is investigated and the stress dependence of creep deformation is
discussed.
2. Experimental procedure
Three melts, hereinafter referred to as heats, of modied
9Cr1Mo steel (ASME SA-213 T91) were used in this study. Chemical
composition and heat treatment condition of the steels are shown
in Tables 1 and 2, respectively. The steels have been produced by
different manufacturers and the contents of minor elements and
heat treatment condition indicate a little difference, however, no
obvious difference has been observed on grain size, hardness and
creep strength. Tensile test was conducted under a constant nominal strain rate of 5 105 s1 up to about 2% of total strain, and it
was increased to 1.25 103 s1 . Strain rate of the tensile test was
controlled by a differential transformer whose resolution was 1 m,
with an extensometer attached to the gauge portion of the specimen. Flow stress was evaluated under a constant nominal strain
rate of 5 105 s1 , as well as 0.2% offset yield stress, in addition to

K. Kimura et al. / Materials Science and Engineering A 510511 (2009) 5863

59

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

Si

Mn

Ni

Cr

MGA
MGB
MGC
SA-213, T91

0.10
0.09
0.09
0.080.12

0.38
0.34
0.29
0.200.50

0.40
0.45
0.35
0.300.60

0.015
0.015
0.009
0.020

0.001
0.001
0.002
0.010

0.12
0.20
0.28
0.40

8.53
8.51
8.70
8.009.50

Heat

Mo

Cu

Nb

Al

MGA
MGB
MGC
SA-213, T91

0.96
0.90
0.90
0.851.05

0.022
0.026
0.032

0.21
0.205
0.22
0.180.25

0.076
0.076
0.072
0.060.10

0.010
0.02
0.001
0.04

0.050
0.042
0.044
0.0300.070

Table 2
Heat treatment condition, grain size and hardness of the steels studied.
Heat

Normalizing

Tempering

Austenite grain
size number

Vickers
hardness

MGA
MGB
MGC

1045 C/10 min AC


1050 C/60 min AC
1050 C/10 min AC

780 C/60 min AC


760 C/60 min AC
765 C/30 min AC

9.2
8.8
9.2

221
221
230

tensile stress that is a ow stress under a constant nominal strain


rate of 1.25 103 s1 .
Creep data over a range of temperatures from 500 to 700 C of
the steels which was reported in the Atlas of Creep Deformation
Properties, No.D-1 of NIMS Creep Data Sheets [19], was used. Creep
deformation data at 550 and 600 C of the steels was analyzed and
instantaneous strain, i , creep strain, c , total strain, t , time to initiation of tertiary creep, tT , time to rupture, tR , and minimum creep
rate were evaluated. Denition of the parameters is described in
Fig. 1.
3. Results and discussion

Fig. 2. Stress vs. time to rupture curves over a range of temperatures from 500 to
700 C of 9Cr1MoVNb steel.

3.1. Creep strength


Stress vs. time to rupture curves over a range of temperatures
from 500 to 700 C of the steels are shown in Fig. 2. Although
slight difference in creep-rupture strength is observed at 700 C
and 30 MPa, creep-rupture strength of three steels are essentially
the same in the tested condition. At 500 C, rectilinear relationship
is observed, however, slope of the curves becomes steeper in the
long-term at 550 C and above. Minimum creep rate of the steels
at 550 and 600 C are plotted against stress and shown in Fig. 3.

Fig. 1. Creep curve and denition of parameters.

Fig. 3. Stress vs. minimum creep rate curves and ow stresses obtained by tensile
test at 550 and 600 C of 9Cr1MoVNb steel.

60

K. Kimura et al. / Materials Science and Engineering A 510511 (2009) 5863

Tensile strength, that is a ow stress under a nominal strain rate


of 1.25 103 s1 , and ow stress under a nominal strain rate of
5 105 s1 are also plotted in the same gure. Stress exponent, n
of the minimum creep rate is 16 at 550 C and 12 in the high-stress
regime at 600 C, however, it decreases to about 4 with decrease
in stress at 600 C. Large stress dependence of the minimum creep
rate at 550 C and in the high-stress regime at 600 C is equivalent
to strain-rate dependence of ow stress evaluated by tensile test,
similar to ASME Grade 122 and Grade 92 steels [20,21]. Precipitation of Z-phase was investigated on MGC heat and it was observed
after creep exposure for about 10,000 h at 600 C and about 5000 h
at 650 C [14,15]. However, the beginning of Z-phase formation did
not consist with inection of the stress-time to rupture curve in the
long-term [17]. The change in stress dependence of minimum creep
rate is considered to be caused by change in deformation mechanism from low stress and elastic regime to high-stress and plastic
regime [20,21].
Creep-rupture elongation and reduction of area of the steels are
shown in Fig. 4 for MGA, MGB and MGC heat. High ductility whose
rupture elongation is 20 to 40% and reduction of area is 80% and
above, is observed for all the steels in the short-term, however,
it decreases in the long-term, especially pertaining to a reduction of area at 650 C. MonkmanGrant relationship of the steels
is shown in Fig. 5. Parameters of A and B obtained by regression
analysis are also indicated in the gure. Good linear relationship
is observed for all the steels and those are equivalent to each
other.
3.2. Creep deformation properties
Initial strain values of the steels at 550 and 600 C are plotted against stress as shown in Fig. 6. Magnitude of initial strain
at 600 C is about 0.2% at 200 MPa, and it decreases to less than
0.1% at 100 MPa with decrease in stress. At 550 C, initial strain is
smaller than that at 600 C at the same stress condition, however,
it decreases with decrease in stress similar to that at 600 C.
Creep curves of MGC heat at 600 C over a range of stresses from
100 to 200 MPa are shown in Fig. 7 in a loglog plot. Magnitude
of creep strain at 0.1 h decreases with decrease in stress. Although
the slope of the creep curve slightly decreases with time within the
beginning of creep deformation less than 1 h, rectilinear relationship is observed in the transient creep stage. In order to express
linear creep curve within the transient creep stage in a loglog
plot, relation between creep strain and time in the transient creep
stage was analyzed with the following equations of (1) power law,
(2) exponential law, (3) logarithmic law [22] and (4) Blackburns
equation [23].
Power law

C = at b

(1)

Exponential law C = 0 + a{1 exp(bt)}

(2)

Logarithmic law C = 0 + a ln(1 + bt)

(3)

Blackburns eq.

C = 0 + a{1 exp(bt)} + c{1 exp(dt)}

(4)

where C is a creep strain and t is a time. 0 , a, b, c and d are constants obtained from analysis. Since C is a creep strain, it does not
include instantaneous strain. Consequently, 0 is not an instantaneous strain and it is used for the analysis except for power law, for
the purpose of improvement of tting accuracy. Results of regression analysis on a transient creep stage for 200 and 100 MPa at
600 C of MGC heat are described in Fig. 8 for 200 and 100 MPa. For
both stress conditions, a linear relationship between creep strain
and time in a loglog plot precisely corresponds to a power law. A
linear relationship in the transient creep stage could not be represented by the other three equations.

Fig. 4. Creep rupture ductility over a range of temperatures from 500 to 700 C of
9Cr1MoVNb steel.

Creep rate vs. time curves of MGC heat at 600 C over a range of
stresses from 100 to 200 MPa are shown in Fig. 9. Creep deformation of the steel consists of transient and tertiary creep stages and
no obvious steady state is observed. In the transient creep stage,
creep rate decreases linearly with increase in time in a double logarithmic plot. Creep rate in the transient creep stage at 200 and
100 MPa is calculated from the predicted creep curves shown in
Fig. 8, and those are described in Fig. 10 for 200 and 100 MPa. Linear relationship between creep rate and time within transient creep
stage in a loglog plot is accurately represented by a power law, as
is the creep curve. Creep deformation in the transient creep stage
of the steel could not be expressed by any of the other expressions. Similar results have also been observed on carbon steel, low
alloy CrMo steels and SUS 316 austenitic steel in this issue [24]. It
should be useful to analyze creep deformation properties within the
transient creep stage with a power law for several creep-resistant
materials.

K. Kimura et al. / Materials Science and Engineering A 510511 (2009) 5863

61

Fig. 5. MonkmanGrant plot steel over a range of temperatures from 500 to 700 C
of 9Cr1MoVNb steel.

Fig. 8. Comparison of observed and predicted creep curves in a transient creep stage
at (a) 600 C: 200 MPa and (b) 600 C: 100 MPa of MGC heat of 9Cr1MoVNb steel.

3.3. Stress dependence of creep deformation

Fig. 6. Stress dependence of initial strain at 550 and 600 C of 9Cr1MoVNb steel.

Fig. 7. Creep curves at 600 C of MGC heat of 9Cr1MoVNb steel.

Creep rate vs. creep strain curves of MGC heat at 600 C over a
range of stresses from 100 to 200 MPa are shown in Fig. 11. No obvious stress dependence has been observed on creep curve in Fig. 7
and creep rate vs. time curve in Fig. 9, however, stress dependence
of creep deformation is clearly recognized on creep rate vs. creep
strain curves. The magnitude of creep strain where a creep rate
shows minimum value decreases with decrease in stress, and stress
dependence of creep deformation is clearly detected in creep rate
vs. creep strain curve. At stress range of 140 MPa and above, creep
rate indicates minimum value at a creep strain of 23%, however, a
minimum creep rate is observed at a small creep strain of less than
1% at a lower stress condition of 120 MPa and below. In addition to

Fig. 9. Creep rate vs. time curves at 600 C of MGC heat of 9Cr1MoVNb steel.

62

K. Kimura et al. / Materials Science and Engineering A 510511 (2009) 5863

Fig. 12. Stress vs. times to 1% total strain, 1% creep strain, initiation of tertiary creep
and time to rupture at 600 C of MGC heat of 9Cr1MoVNb steel.

Fig. 10. Comparison of observed and predicted creep rate vs. time curves at (a)
600 C: 200 MPa and (b) 600 C: 100 MPa of MGC heat of 9Cr1MoVNb steel.

power-law creep deformation behaviour within the transient creep


stage, stress dependence of the onset creep strain of tertiary creep
stage should be considered for evaluation of long-term creep deformation properties, because design of high-temperature structural
components is controlled not only by creep-rupture strength, but
also creep-deformation properties. 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 [25], a temperature and time-dependent stress intensity limit,
St value is determined for each specic time by the lesser of the
following.

Stress vs. times to 1% total strain, 1% creep strain, initiation


of tertiary creep and time to rupture at 600 C of MGC heat of
9Cr1MoVNb steel is shown in Fig. 12. At 200 MPa, 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 decreasing with decrease in stress and it becomes negligibly small at 100 MPa.
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 9Cr1MoVNb steels are shown in Fig. 13. Life
fraction of the time to initiation of tertiary creep is within a range
of 0.60.7 at 550 C and a range of 0.50.7 at 600 C, and it is almost
constant independent of time to rupture. On the other hand, life
fraction of times to 1% total strain and 1% creep strain is smaller
than 10% of time to rupture in the short-term, and it increases with
increase in time to rupture. Life fraction of times to 1% total strain

(1) 100% of the average stress required to obtain a total strain of 1%,
(2) 80% of the minimum stress to cause initiation of tertiary creep,
(3) 67% of the minimum stress to cause rupture.

Fig. 11. Creep rate vs. creep strain curves at 600 C of MGC heat of 9Cr1MoVNb
steel.

Fig. 13. 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
9Cr1MoVNb steels.

K. Kimura et al. / Materials Science and Engineering A 510511 (2009) 5863

and 1% creep strain increases to about 50% of time to rupture in


the long-term at 600 C. It is almost the same as the life fraction of
initiation of tertiary creep. 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 denitely smaller than 100% of
the average stress required to obtain a total strain of 1% in the longterm. In the short-term, 1% creep strain is observed in a transient
creep stage and it is attained in a tertiary creep stage in the longterm as shown in Fig. 11. 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. Inection of stress vs. time to rupture curve
of the steel is considered to be caused by the stress dependence of
creep deformation property. The effect of stress on creep deformation property, especially on long-term creep deformation, should
be taken into account in consideration of microstructural change
during creep exposure.
4. Conclusions
Creep deformation property of 9Cr1MoVNb steels (ASME
SA-213 T91) was investigated and stress dependence of the creep
deformation property was discussed. Large stress dependence of
creep rupture life and minimum creep rate in the high-stress regime
decreased in the low stress regime. Large stress dependence of
the minimum creep rate in the high-stress regime was equivalent to strain-rate dependence of ow stress evaluated by tensile
test. It was considered to be caused by change in deformation
mechanism from low stress and elastic regime to high-stress and
plastic regime. Good linear relationship between creep strain vs.
time and creep rate vs. time in a double logarithmic plot were
observed. It was appropriately expressed by a power law rather
than exponential law, logarithmic law and Blackburns equation.
A magnitude of creep strain where a creep rate shows minimum
value decreased with decrease in stress. Life fraction of times
to 1% total strain and 1% creep strain increased with increase
in creep rupture life from several per cent of rupture life to
about 50% of that in the long-term at 600 C. 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%.
The effect of stress on creep deformation property, especially
on long-term creep deformation, should be taken into account
in consideration of microstructural change during creep exposure.

63

Acknowledgement
Part of this study was nancially 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 of Japan.
References
[1] K. Kimura, H. Kushima, F. Abe, in: R. Viswanathan (Ed.), Advances in Life Assessment and Optimization of Fossil Power Plants, EPRI, California, 2002.
[2] K. Kimura, H. Kushima, K. Sawada, in: A. Strang, et al. (Eds.), Engineering Issues
in Turbine Machinery, Power Plant and Renewables, MANEY, London, 2003, p.
443.
[3] K. Kimura, K. Sawada, K. Kubo, H. Kushima, in: Y.Y. Wang (Ed.), Experience with
Creep-strength Enhanced Ferritic Steels and New and Emerging Computational
Methods, PVP, vol.476, ASME, New York, 2004, p. 11.
[4] K. Maruyama, J.S. Lee, in: I.A. Shibli, et al. (Eds.), Creep & Fracture in High Temperature ComponentsDesign & Life Assessment Issues, DEStech Publications,
Inc., Pennsylvania, 2005, p. 372.
[5] B. Wilshire, P.J. Scharning, Scripta Mater. 56 (2007) 701.
[6] K. Kimura, Proceedings of the 2005 ASME Pressure Vessels and Piping Division
Conference, Denver, USA July 1721, 2005, PVP2005-71039.
[7] K. Kimura, Proceedings of the 2006 ASME Pressure Vessels and Piping Division
Conference, Vancouver, Canada, July 2327, 2006, PVP2006-ICPVT11-93294.
[8] ASME Boiler and Pressure Vessel Code Case 2179-6, August 4, 2006.
[9] ASME Boiler and Pressure Vessel Code Case 2180-4, August 4, 2006.
[10] ECCC data sheet, Steel ASTM Grade 92, 2005.
[11] A. Strang, V. Foldyna, A. Jakobova, Z. Kubon, V. Vodarek, J. Lenert, in: A. Strang,
et al. (Eds.), Advances in Turbine Materials, Design and Manufacturing, The
Institute of Materials, London, 1997, p. 603.
[12] A. Strang, V. Vodarek, in: A. Strang, et al. (Eds.), Microstructural Stability of
Creep Resistant Alloys for High Temperature Plant Applications, The Institute
of Materials, London, 1998, p. 117.
[13] P. Hofer, H. Cerjak, P. Warbichler, Mater. Sci. Technol. 16 (2000) 1221.
[14] K. Kimura, H. Kushima, F. Abe, K. Suzuki, S. Kumai, A. Satoh, in: A. Strang, et
al. (Eds.), Advanced Materials for 21st Century Turbines and Power Plant, The
Institute of Materials, London, 2000, p. 590.
[15] K. Suzuki, S. Kumai, H. Kushima, K. Kimura, F. Abe, Tetsu-to-Hagane 89 (2003)
691698 (in Japanese).
[16] H.K. Danielsen, J. Hald, Energy Mater. 1 (2006) 49.
[17] K. Sawada, H. Kushima, K. Kimura, M. Tabuchi, ISIJ Int. 47 (2007) 733.
[18] S. Holdsworth, Mater. High Temp. 21 (2004) 2532.
[19] Atlas of Creep Deformation Property, NIMS Creep Data Sheet, No.D-1, National
Institute for Materials Science, 2007.
[20] K. Kimura, K. Sawada, H. Kushima, Y. Toda, Proceedings of the Fifth International
Conference Advances in Materials Technology for Fossil Power Plants, Marco
Island, Florida, USA, October 35, 2007.
[21] K. Kimura, K. Sawada, H. Kushima, K. Kubo, Int. J. Mater. Res. 99 (2008) 395.
[22] F. Garofalo, Fundamentals of Creep and Creep Rupture in Metals, MacMillan,
New York, 1965.
[23] L.D. Blackburn, The Generation of Isochronous StressStrain Curves, ASME, New
York, 1972, p. 15.
[24] K. Sawada, M. Tabuchi, K. Kimura, this issue.
[25] ASME Boiler and Pressure Vessel Code, Section III, Division 1, Subsection NH,
2004.

S-ar putea să vă placă și