You are on page 1of 8

Available online at www.sciencedirect.

com

Procedia Engineering 55 (2013) 591 – 598

6th International Conference on Cree


ep, Fatigue and Creep-Fatigue Interaction [CF-6
6]

Advances in Creep Damagge/Life Assessment Technology fo


or
Creep Strength Enhanced Ferritic Steels
mitsu Masuyama∗
Fujim

Graduate School of Engineering, Kyushu Institute of Technology, 1-1, Sensui-cho, Tobata, Kitakyushu 804-8550, Japan

Abstract

The creep damage/life assessment technology for creeep strength enhanced ferritic (CSEF) steels such as Gr.91, 92, 122, etc.
has been strongly demanded by power plant operatorr and equipment manufacturers however creep damage mechan nisms of
these steels are not well clarified. Since the creeep degradation mechanism of CSEF steels is very compliccated in
comparison with the conventional CrMo steels, tthe study on CSEF steels must be grounded on the findiings on
microstructural degradation and creep softening in martensitic structure composed of very fine martensite lath,, block,
packet and prior austenite grains, and precipitationn and dislocation structures. This paper presents new findiings on
microstructural creep degradation, the hardness creepp life model and the simplified tertiary creep modeling of the Omega
method for creep life assessment of CSEF steels aas the advances in creep damage/life assessment technology for its
practical applications in future.

©
© 2013 TheAuthors.
2013 The Authors. Published
Published by Elsevier Ltd. Ltd. Seleection and/or peer-review under responsibility of the Indira Gand
by Elsevier dhi
Centre forand
Selection Atomic Research
peer-review under responsibility of the Indira Gandhi Centre for Atomic Research.

Keywords: Creep damage; life assessment; creep strengthh enhanced ferritic (CSEF) steels; hardness creep life model; tertiarry creep
modeling; Omega method

1. Introduction

The creep strength enhanced ferritic (CSEFF) steels such as Gr. 91, 92, 122, etc. developed in 1980s and
1990s have extensively used in modern fossil-ffired power plants with high steam parameter for last deecades.
However a number of creep failure experiencees with these steels used in the superheater/reheater tubes, hot
reheat pipes and headers have been reported. IIt is essential to evaluate the material conditions and rem
maining
life of the materials for long-term use in poweer plant components, but it becomes apparent that thesee steels
exhibits different degradation behavior from cconventional CrMo steels due to the fine lath structure of the
tempered martensite and its complicated creep ddegradation mechanism in base metal and welds. Accord dingly,
it is difficult to apply conventional techniquess for the diagnosis of remaining life, and new techniqu ues are
required for the appropriate life management oof equipment using these steels and for the achievement of life


E-mail address: masuyma@tobata.isc.kyutech.ac.jp

1877-7058 © 2013 The Authors. Published by Elsevier Ltd.


Selection and peer-review under responsibility of the Indira Gandhi Centre for Atomic Research.
doi:10.1016/j.proeng.2013.03.300
592 Fujimitsu Masuyama / Procedia Engineering 55 (2013) 591 – 598

extension[1]. In the present paper new findings on microstructural creep degradation, hardness creep life model
and simplified tertiary creep modeling developed for creep damage/life assessment of CSEF steels are
introduced as the advances in this field.

2. Microstructural creep degradation in CSEF steels

Since the CSEF steels have tempered martensite structure, as-received hardness is around 220HV. This
hardness declines by only 5% when subjected to thermal aging for 30,000h at 650 , and declines by 20-30%
due to creep after 10,000 to 30,000 h exposure associated with microstructural degradation. However it is very
difficult to determine the creep degradation and material conditions through the optical microstructure in case
of CSEF steels, while the conventional CrMo steels with ferrite and pearlite structures or coarse grained
martensitic structures for base metal or heat affected zone respectively can be determined their degradation
rating and creep damage percentage. Fig.1 schematically illustrates the creep damage progress for CrMo steels
and CSEF steels as well as the damage detecting method applied to the first half of the creep life. The micro
cracks and macro cracks in the latter half stage III and stage IV of the creep life are detectable by non-
destructive testing and replication technique. In case of CrMo steels microstructural changes (stage I) under
the optical microscope followed by creep cavitation process (stage II) can be classified against the creep
damage rating, but both stages in CSEF steels are not distinguishable, and creep damage rating is impossible by
the microstructural observation. Recently microstructural changes are simulated employing the free energy
concept as a thermodynamic model for steels. According to the investigation result [2] on Gr.91 and Gr.92 the
system energies are reduced with creep time toward the saturated values which are higher than 0 of the
equilibrium state, and it is seen that there are no marked difference between grip portion and gauge portion for
Gr.91, and larger drop in the energy for Gr.91 than Gr.92 due to the relatively greater structural degradation.
These results suggest the structural simulation would be helpful to assess the material conditions and the creep
resistance of the materials.

Fig. 1. Creep damage progress and damage detecting methods.

M23C6 and Laves phase can be observed as precipitates at the tempered martensite grain boundaries using an
optical microscope, and TEM allows observation of MX as well. MX is extremely fine, and does not show any
major change in particle size resulting from thermal aging or creep. In contrast M23C6 and Laves phase show
coarsening due to thermal aging and creep. There are four kinds of grain boundaries in the structure, i.e., prior
austenite, packet, block and lath, while prior to the creep exposure M23C6 is seen at all grain boundaries and
Fujimitsu Masuyama / Procedia Engineering 55 (2013) 591 – 598 593

MX is observed at the lath boundaries and within the lath. After creep exposure new Laves phase precipitation
is seen at the grain boundaries. It is known that the sizes of these precipitates change due to heating and creep.
According to particle size distribution measurement results [3] for MX, M23C6 and Laves phase on the grip and
gauge potions of 10Cr-1Mo-1W-V-Nb steel, average values on the grip portion are approximately 30nm for
MX, 80nm for M23C6 and 350nm for Laves phase, while on the gauge portion these are approximately 50nm,
130nm and 400nm, respectively. Looking at the precipitate size distribution measurement results3) in the case
of thermal aging at 600 , Laves phase precipitation is not seen prior to the thermal aging, with distribution
occurring due to the heating. With respect to the Laves phase size distribution, the peak location moves rapidly
toward the maximum in conjunction with heating time. Specifically, the average size that is 70nm at around
1,000h changes to 400nm at around 33,000h. In contrast, the average sizes of MX and M23C6 approximately
50nm and 80nm respectively, do not exhibit any major changes due to heating. From the foregoing, Laves
phase undergoes considerable coarsening even without the stress, while it can be seen that growth without
stress for MX and M23C6 is quite limited, and that even the effect of stress on MX is quite small. In this
context, M23C6 appears particularly sensitive to the action of stress. According to the precipitate size
distribution measurement result [4] before and after creep (interrupted at 0.9 of t/tR) performed on Gr.91, there
is no change in the average particle size for all precipitates, and only the precipitates in excess of 100nm are
seen to experience coarsening. Precipitates in excess of 100nm correspond to M23C6 and Laves phase, and it is
considered that coarsening of Laves phase is due to heating during creep, while that of M23C6 is due to the
action of stress during creep exposure.
The CSEF steels are composed of very fine martensite lath structures with approx. 1ȝm width at normalized
and tempered conditions, and the morphology of lath structures correlates with the crystallography. The
morphology and crystallography of martensite lath structures have been extensively studied by use of
metallographic instruments. Particularly EBSD (Electron Backscattering Diffraction) technique recently
becomes a strong tool to reveal the crystallography changes of martensite lath structures during creep as well as
the conventional techniques of TEM and SEM, because the crystallographic orientation of martensitic
structures changes with recovery and creep deformation. Measuring the misorientation angles is a way to know
the information about grain boundary changes due to creep degradation. Resent research work [5] has clarified
the changes in lath boundary length per unit area in terms of creep strain under the several creep conditions
using interrupted creep specimens so as that the length of lath boundary increases a bit corresponding to the
generation of new boundaries within the lath shown at the beginning of creep, then linearly decreases with
creep strain due to the recovery of lath structure with subgrain growth.
A TEM photograph [6] in Fig. 2 [7] indicates the structure of Gr.91 with 25% tensile strained at 820
which is just below the Ac1 temperature. Normalized and tempered as-received material has extremely high
dislocation density, with an approx. 1ȝm width lath structure, but deformation at high temperatures results in
dramatic lowering of the dislocation density, and substantial growth in lath size can be seen without any
evidence of phase transformation. Also, while it is thought that precipitates originally exist at the lath
boundaries or other grain boundaries, the photograph clearly shows that these are dispersed within the grains,
and the alignment of precipitates suggests that they had been at the original grain boundaries. It has already
been known that the grain size of this lath (inclusive of subgrains) is related to creep resistance, and as
indicated in Fig. 3 [6], creep resistance declines in proportion to the cube of the lath width. The lath width is
also strain-dependent, and as shown in Fig. 2 [7], this increases proportionately to strain until the strain reaches
approx. 0.1. It appears that necking occurs in cases where the strain is above approx. 0.1, and, although the lath
width is shown as a constant value in order to separate the location of structural observation from the necking
area, it is considered that there is actually a wide range in which the proportional relationship holds. It is clear
that such lath boundary movement or grain growth is induced by the action of stress and the resulting strain.
Based on consideration [7] of this process, precipitates are strongly bound, with the dislocation networks
(serving to increase internal stress) moving as a result of the action of stress and reaching the lath boundaries.
When these networks are swept by the movement of the lath boundaries, it is thought that they are absorbed,
thus increasing the interface energy. Considering the growth of the lath structure, given some form of change
to the precipitate or interaction between the precipitate and the dislocation to which it is bound, the dislocation
could shift away and reach to the lath boundary, thus being the direct cause of the boundary movement. The
594 Fujimitsu Masuyama / Procedia Engineering 55 (2013) 591 – 598

lath width would be inversely proportional to the load stress (lath width being greater with lower stress), and as
a result the lath structure or subgrains that had grown would have a major effect on hardness. Especially in the
case of ruptured material, it is predicted that hardness will decline with lower load stress. With respect to the
lath width distribution as well, while a normal distribution applies to cases of high stress above approximately
100MPa, the distribution in the case of low stress is characterized by two maxima, and mixing of coarse and
fine grains is reported [8]. Hardness, on the other hand, is proportional to the square root of the dislocation
density remaining after creep deformation, such that a relationship can be established between hardness and
lath width. Accordingly, because lath growth behavior differs depending on stress, it is predicted that
differences will also occur in terms of changes in hardness. As a result of investigation [9] into the influence of
lath width and dislocation density on the hardness of Gr.91, it has been clarified that increased lath width and
reduced dislocation density cause hardness to decline. As lath width and dislocation density are directly related
to creep resistance measurement of hardness can be understood to enable estimation of material creep
resistance.

Fig. 2. Relationship between lath width and minimum creep rate for Fig. 3 Relationship between lath width and strain for Gr.91 steel.
Gr.91 steel.

3. Novel creep damage/life assessment techniques

3.1. Hardness model for creep life assessment for CSEF steels

Consideration thus far clearly indicates that hardness drop due to creep, and that the amount of decline is
strongly related to life consumption or creep degradation, and the hardness of tempered martensitic structure is
influenced by the structural conditions of grain boundaries, subgrain size, precipitates, dislocation density and
creep strain [10]. The CSEF steel in the present context experiences reduced hardness due to thermal aging,
but the amount of decline is small, and is actually extremely large under the action of stress. That is, reduced
hardness in the crept materials is based on structural changes dependent upon stress, and there is no doubt that
the assessment of life from hardness measurement values has a firm basis in materials science. The hardness
drop measured for Gr.91 thermally aged and creep interrupted for base metal and welds (minimum hardness in
the heat affected zone), taking the Larson-Miller parameter as a variable, can be demonstrated as shown in Fig.
4 [11]. It is found from the figure as that the hardness data from the higher stress tests at 98MPa and above
scatter at relatively lower values of the Larson-Miller parameter, while the data from the lower stress tests
scatter at higher parameter values, thus being divided into two parts (including the welds data). The reason
Fujimitsu Masuyama / Procedia Engineering 55 (2013) 591 – 598 595

why the hardness versus Larson-Miller parameter plot can be divided into two parts is considered to be strain-
induced softening for the higher stress test and stress-induced softening for the lower stress test, without any
dependence of temperature and time. The hardness measured on the grip potion well corresponds to that of the
aged specimens. The hardness (H) measured for creep specimen linearly reduced with the creep life fraction,
and there is a linear relationship between hardness and creep life fraction of 0.2 to 0.9 for base metal and welds.
When the initial hardness (H0) for base metal and welds are defined as the hardness of the grip portion of the
base metal and that of the base metal of the welded joint respectively, the hardness ratio is shown as being
linear against the creep life fraction as indicated in Fig. 5 [12]. The hardness ratio (H/H0) can be expressed as
below,

H/Ho = 0.98-0.15t/tR (1)

where t/tR is creep life fraction.


The hardness drop (ǻH) due to creep and thermal aging can be demonstrated as a function of Larson-Miller
parameter as shown in Fig. 4. Three lines represent the hardness drop due to creep or aging for high stress
creep test, low stress creep test, and thermal aging test respectively. The logarithmic hardness drop is
expressed by following equation.

lnǻH = lnǻH0 + Ks×(LMP) (2)

where ǻH: hardness drop, ǻH0: initial hardness drop, Ks: coefficients for high stress (98MPa and 115MPa),
low stress (71MPa) and aging, LMP: T(20 + log t), T is absolute temperature in K. t is time in hour. lnǻH0 is
assumed to be 0. Therefore the equation (2) can be expressed in following form.

lnǻH = Ks(LMP) = Ks×T(20 + log t) (3)

Fig .4 Relation between hardness drop and parameter for Gr.91. Fig. 5 Relation between hardness ratio and Larson-Miller creep
life fraction for Gr.91.

As creep life fraction t/tR is given by the equation (4), initial hardness H0 is not necessary to be provided, but
only the temperature and operating hours as well as the hardness value instantaneously measured are inputs to
calculate the creep life fraction consumed.

t/tR = 1/0.15(0.98 – H/H0) = 1/0.15{0.98 – H/{H + exp(Ks×T(20 + log t))} (4)


596 Fujimitsu Masuyama / Procedia Engineering 55 (2013) 591 – 598

Using this hardness model creep lives of creep specimens were predicted to compare with the observed
creep life fractions, and found that predicted lives are longer than the observed one in the approximate range of
factor of 2 for high stress softening and factor of 3 for low stress softening [11].

3.2. Simplified tertiary creep modeling method

In order to assess the creep damage or creep life a tertiary creep modeling so called the Omega method [12]
is adopted by API 579, and the logarithm of the strain rate is well expressed by the following equation over a
wide range of tertiary creep, irrespective of stress and temperature.

ln ȑ = ln ȑ0 + Ωε or ȑ = ȑ0 exp (Ωε) (5)

where ȑ is the creep rate, ȑ0 is the initial creep rate, and Ω is the strain rate acceleration factor (dlnȑ/dε).
When the duration of primary creep is short enough, the creep life and creep strain can be expressed by Ω
and ȑ0, and the remaining creep life is expressed by following equation.

tR – t i = 1 / Ω ȑ I (6)

Therefore to obtain the remaining life the values of Ω and the instantaneous true creep strain rate ȑ i are
needed at the given time t i, while these values are sometimes difficult to be determined. In order to simplify
the remaining life prediction by Omega method it was considered that the instantaneous true creep strain rate ȑ i
could be estimated from the average creep rate, ȑavg, which should be the ratio of the creep deformation to the
operating duration, and the parametric expression of the value of Ω could be developed. In this sense ȑ i is
expressed by αȑavg (α is the coefficient), and Ω is expressed by the parametric function of stress and
temperature as mentioned below [13]. To express the parametric expression of the value of Ω Larson-Miller
parameter is employed and the time to rupture in the equation is converted to KMG ȑ (Monkman-Grant
relationship), where KMG is the Monkman-Grant constant, and ȑ is the minimum creep rate. Then since the
minimum creep rate ȑ is expressed by Norton’s law as ȑ = Aσ n, KMG ȑ becomes KMGAσ n. Accordingly the
temperature-stress parameter, Φ (T, σ ) is expressed by following equation.

Φ (T, σ )= T [ C + log { KMG / (A σ n ) } ] (7)

where C is the parameter constant.


The magnitude of the Ω value increases with decreasing both stress and temperature at the temperature of
500 and above. The Ω values can be expressed in terms of temperature-stress parameter Φ (T, σ ), and the
linear relationship between both the Ω value and Φ (T, σ ) estimates the Ω value to predict the remaining creep
life. Accordingly the deformation record of the equipment at the given operating hours, and the data bases of
the material used such as Monkman-Grant constant and Norton’s stress exponent as well as the operating stress
and temperature simply provides the remaining creep life.

3.3. Practical creep life assessment systems for CSEF steels

The remaining creep life assessment, in general, essentially consists of the theoretical calculation, non-
destructive examination and destructive test of serviced materials, and levels 1, 2 and 3 approaches are
proposed [14]. In the case of the conventional CrMo steels with ferrite and pearlite structures or coarse grained
martensitic structures in the heat affected zone replication technique to observe the optical microstructure to
determine the degradation rating is very useful, while the hardness is not a good indicator of the degradation of
annealed CrMo steels. On the other hand it is very difficult to determine the creep degradation rating and
material conditions through the optical microstructure in the case of CSEF steels. Several kinds of non-
destructive damage detecting techniques such as replication method, electrochemical method, potential drop
method and hardness method are so far proposed and investigated, however those techniques have not yet been
Fujimitsu Masuyama / Procedia Engineering 55 (2013) 591 – 598 597

well established. The creep degradation mechanism in the martensitic steels is composed of grain boundaries,
lath width, precipitates, dislocation density and creep strain, and these microstructural parameters strongly
influence to the softening of martensitic structures. Therefore the hardness measurement method is a promising
technique for material conditions and creep life assessment. To perform the systematic creep life assessment of
CSEF steels, at first analytical calculation using tertiary creep modeling, the Omega method simplified as
previously mentioned is recommended to be made using data base and measured dimensional deformation.
The electric potential drop ratio value measured in the damaged (creep-softened) potion to that at the portion
without stress or referenced potion is well correlated to the hardness drop which is corresponded to the creep
life fraction as shown in Fig. 6 [13]. When the hardness is measured on the concerned portion of the
component, the measured result is directly addressed to the creep life prediction using the hardness model
proposed with needless of initial hardness value. Then finally the combined evaluation of the assessment
results from the above mentioned simple methods will provide the accurate creep life prediction for CSEF
steels in the systems shown in Fig. 7.

Fig. 6. Relationship between potential drop and hardness drop Fig. 7. Creep life assessment system for CSEF steels.
due to creep.

4. Summary

The creep damage/life assessment for CSEF steels has been studied worldwide from the various aspects.
Based on the findings on microstructural degradation and creep softening in martensitic structure composed of
very fine martensite lath, block, packet and prior austenite grains, and precipitation and dislocation structures,
hardness creep life model and tertiary creep modeling method were investigated and a systematic creep life
assessment system was proposed. The creep degradation in CSEF steels is strongly related to the changes in
microstructural factors influence to the softening of martensitic steels. Consequently the systematic use of
hardness modeling combining with the potential drop method and tertiary creep modeling was thought to be a
promising technique for the material conditions assessment and the creep life assessment of CSEF steels.

References

[1] F.Masuyama: Advanced technology in creep life prediction and damage evaluation for creep strength enhanced ferritic steels, Proc.
HIDA (High-temperature Defect Assessment) -5 Conference, Guilford, UK, June 23-25, (2010)
598 Fujimitsu Masuyama / Procedia Engineering 55 (2013) 591 – 598

[2] Y.Saito, T.Kita, Y.Murata, M.Morinaga, T.Koyama, Y.Hasegawa, M.Igarashi: Proc. 47th Symposium on Strength of Materials at
High Temperatures, (2009), 81-85.
[3] H.Cerjak, P.Hofer, B.Schaffernak: The influence of microstructural aspects on the service behavior of advanced power plant steels,
ISIJ International, 39(1999)874-888.
[4] F.Masuyama, N.Nishimura, A.Sasada: Precipitation behavior during creep of modified 9Cr-1Mo steel, CAMP-ISIJ, 11(1998)1245
[5] M.Mitsuhara, S.Morioka, S.Hata, K.Ikeda, H.Nakashima: Analysis of creep degradation behavior in lath martensite based on
crystallographic orientation changes, Report of the 123rd Committee on Heat-resisting Metals and Alloys, Japan Society for the
Promotion of Science, 50(2009)37-43.
[6] F.Masuyama, N.Nishimura: Effect of structures and residual elements on creep behavior of modified 9Cr-1Mo steel, Strength of
Materials (eds. Oikawa H., et al.), (1994)657-660.
[7] T.Endo, F.Masuyama, K.S.Park: Change in hardness and substructure during creep of mod.9Cr-1Mo steel, Tetsu-to-Hagane,
88(2002)526-533.
[8] K.Suzuki, S.Kumai, H.Kushima, K.Kimura, F.Abe: Heterogeneous recovery and precipitation of Z phase during long term creep
deformation of modified 9Cr-1Mo steel, Tetsu-to-Hagane, 86(2000)550-557.
[9] K.Sawada, K.Maruyama, Y.Hasegawa, T.Muraki: Creep life assessment of high chromium ferritic steels by recovery of martensitic
lath structure, Key Engineering Materials, 171-174(2000)109-114.
[10] T.Endo, F.Masuyama, K.S.Park: Changes in Vickers hardness and structure during creep of mod. 9Cr-1Mo steel, Materials
Transaction, 44(2003)239-246.
[11] F.Masuyama: Hardness model for creep life assessment of high strength martensitic steels, Materials Science and Engineering A,
510-511(2009)154-157.
[12] M.Prager: The omega method –an engineering approach to life assessment, Journal of Pressure Vessel Technology, Transaction of
ASME, 122(2000)273-280.
[13] F.Masuyama, T.Tokunaga, N.Shimohata, T.Yamamoto, M.Hirano: Comprehensive approach to creep life assessment of martensitic
heat resistant steels, Proc. ECCC Creep Conf., (2009)31-42.
[14] F.Masuyama: Life assessment and extension of welded structures for high temperature components, Welding in the World,
32(1993)51-64.

You might also like