Sie sind auf Seite 1von 18

Reprinted from JAI, Vol. 8, No.

4
doi:10.1520/JAI103712
Available online at www.astm.org/JAI

V. Kalyanasundaram,1 A. Saxena,2 S. Narasimhachary,1


and B. Dogan3

ASTM Round-Robin on Creep-Fatigue


and Creep Behavior of P91 Steel
ABSTRACT: The American Society for Testing and Materials (ASTM),
through its Committee E08 on Fatigue and Fracture subcommittee E08.05 on
Creep-Fatigue Crack Formation, has recently developed a new standard for
creep-fatigue testing (ASTM E2714-09). This paper describes the plans and
preliminary results from a round-robin being presently conducted in support
and verication of this new standard. The choice of the test material (ASTM
Grade P91), the design of the round-robin test matrix, and a machining plan
for the specimens are described. The results of microstructural analysis,
tensile testing, and creep deformation and rupture testing are also presented
along with some preliminary results from creep-fatigue testing. A new analytical model for representing the creep deformation characteristics of this material is also presented and evaluated using the creep data generated as part
of the round-robin program. The results of the round-robin creep-fatigue
testing will be used to make appropriate modications to the test standard.
KEYWORDS: creep-fatigue, crack formation, statistical data analysis, creep
deformation and rupture

Introduction
ASTM E2714-09 [1] covers the determination of mechanical properties pertaining to creep-fatigue crack formation in nominally homogeneous materials by
the use of test specimens subjected to uniaxial forces under isothermal

Manuscript received December 22, 2010; accepted for publication March 3, 2011;
published online April 2011.
1
Dept. of Mechanical Engineering, Univ. of Arkansas, Fayetteville, AR 72701.
2
Dept. of Mechanical Engineering, Univ. of Arkansas, Fayetteville, AR 72701
(Corresponding author), e-mail: asaxena@uark.edu
3
EPRI, Charlotte, NC 28262.
Cite as: Kalyanasundaram, V., Saxena, A., Narasimhachary, S. and Dogan, B., ASTM
Round-Robin on Creep-Fatigue and Creep Behavior of P91 Steel, J. ASTM Intl., Vol. 8,
No. 4. doi:10.1520/JAI103712.
C 2011 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Copyright V
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
23
Downloaded/printed by

(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

24 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

conditions. It concerns creep-fatigue testing at strain rates and/or cycles involving sufciently long hold times to induce creep deformation (and oxidation)
during cyclic deformation where cycles to crack formation are affected by
creep.
This test method is applicable to the determination of deformation and
crack formation or nucleation properties as a consequence of either constantamplitude strain-controlled tests or constant-amplitude force-controlled tests
with hold times. It is primarily concerned with the testing of round bar specimens subjected to uniaxial loading in either force or strain control, wherein the
latter is recommended. As a result of this round-robin, future improvements
planned for this standard include, but are not limited to, a more denitive precision and bias statement.
A total of 16 participants from laboratories all over the world are conducting tests under a coordinated set of test conditions. The primary objective of the
round-robin is to conduct creep-fatigue tests to characterize the number of
cycles for crack formation while using the procedures specied in the standard
to assess variability in the results. The list of participants, their afliation, and
the planned set of tests for each of them is provided in Table 1.

Test Material
The candidate materials considered for the round-robin were from a wide variety of materials that are used in high temperature applications and included gas
turbine materials for aircraft and land-based engines, fossil power plants, and
nuclear reactor materials. Other considerations included material availability in
sufcient quantities, offers of sponsorship for material and machining costs,
and the interest level in testing the material among the participating volunteers.
The test material selected is ASTM Grade P91 steel (donated by the Electric
Power Research Institute (EPRI), Charlotte) that has a creep rupture strength
of 94 MPa at 600 C for a life of 105 h. The nominal chemical composition of P91
steel in wt % is given in Table 2 [2]. There are two other round-robins with similar objectives that are under way. The Idaho National Laboratory is conducting
a separate round-robin on IN617 (a nickel-based superalloy) and the Japan Society for Promotion of Science is conducting one on ASTM Grade P92 material.
While these are parallel independent studies, an effort is planned to integrate
the ndings from all the studies at a later date.

P91 SteelBrief Overview


Modied tempered martensitic heat resistant steels of type 9Cr-1Mo-Nb-0.2V
(P91/T91 as per ASTM A335 [3]/ASTM A213 [4], respectively) developed in the
United States in the early 1980s, are used in main steam pipes, superheaters,
headers, boilers, and turbines in supercritical and ultra-supercritical fossil
power generation plants [5]. As a replacement for low alloy ferritic and austenitic stainless steels, the high creep strength of P91 permits its use in relatively
thinner-wall components. This class of steel also offers the advantages of

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

Steel Grade P91

Si

0.31

0.11

0.45

Mn

2
2
2

2
2
2
2
2

60.75 %

2
2
2
2
2
2
2

2
4

60.5 %/10
min Hold

0.011

P
0.009

60.75 %/10
min Hold

0.19

Ni

8.22

Cr

0.94

Mo

0.005

As

0.21

0.07

Nb

0.006

Al

0.16

Cu

0.039

2
4

0.001

Bal.

Fe

60.5 %/30
min Hold

Sb, Sn

2
2
2
2
2
2
2
2
Conduct creep deformation and rupture testing and metallography
and TEM of the test material

2
2
2

60.5 %

60.25 %

P91 Crack Formation

TABLE 2Nominal chemical composition of P91 steel (wt %) 2.

MPA Stuttgart:Andreas Klenk/Karl Meile


British Energy: Mike Spindler
Politecnico di Milano: S. Beretta
BiSS, India: R. Sunder

Idaho National Laboratory: Laura Carroll


EMPA Switzerland: Stuart Holdsworth
Tech U. Darmstadt: Alfred Scholz
CRIEPI, Japan: Y. Takahashi
GE, Schenectady: David Knorr
BAM, Berlin: Hellmuth Klingelhoeffer
NASA, Glenn: Brad Lerch
ANSTO Australia: Warwick Payton
Georgia Tech: Rick Neu
Tohoku Univ.: A. T. Yokobori
NRC, Canada: Jonathan Tsang
Univ. of Arkansas: Ashok Saxena

StrainAmplitude %!

Participants
;

TABLE 1List of participants and test matrix for the ongoing ASTM round-robin (Numbers indicate the total number of tests planned under
those conditions).

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 25

FIG. 1(a) P91 pipe section used for the ASTM round-robin testing; (b) cross-sectional view of section 2 of the pipe section; (c)
machining layout for subsection 2-1 (Material courtesy of EPRI).

26 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 27

low thermal expansion, high thermal conductivity, high corrosion cracking


resistance, a low oxidation rate, and good inspectability and weldability [6].
These advantages allow this material to be a potential candidate to replace currently used P22 (2-1/4 CrMo) steel and an alternative for future applications in
power plants operating at service conditions greater than 600 C and a steam
pressure of 300 bars.

Test Specimens and Machining Plan


Test specimens were machined from a pipe section 482 mm in outer diameter,  1 m in length and 47.5 mm in wall thickness [see Fig. 1(a)]. The pipe
was renormalized to ensure consistency with the original microstructure,
then cut in to three sections along its length and labeled as pipe sections 1,
2, and 3, respectively. For this round-robin, the cut section 2 was again cut
in to six equal subsections along the cross-section and labeled as shown in
Fig. 1(b). Specimen blanks were machined from these subsections to accommodate various specimen sizes and geometries as provided by the participants. Figure 1(c) shows the machining layout for subsection 2-1 as an
example. Specimen sizes and grip congurations were selected by the participants, but they were required to be within the allowable limits of the standard. Figures 2(a) and 2(b) show two typical specimen geometries and sizes
requested by participants as part of the round-robin for performing creepfatigue testing and for the supporting creep deformation and rupture testing,
respectively.

Microstructural Characterization, Tensile and Creep Behavior


Figure 3(a) shows the microstructure of the renormalized P91 material wherein
the martensite phase can be observed in the lath microstructure arranged in
packets (3040 lm in size) within prior austenite grains that were on average
100 lm in diameter. Optical micrographs at high magnications revealed extensive carbide precipitates rich in chromium, namely, M23 C6 , where M Cr (sizes
up to 100 nm), that can be located at grain and packet boundaries and ne carbonitride precipitation, namely, MX, where M Nb or V and X C and/or N,
inside the packets similar to the schematic shown in Fig. 3(b). It has been found
that this material has excellent creep resistance as its martensitic microstructure includes signicant dislocation content in the lath structure introduced by
heat treatment [7].
The University of Arkansas was responsible for conducting tensile (performed on-site at Bangalore Integrated System Solutions (BiSS), India) and
creep rupture tests and the subsequent metallographic evaluation of the failed
test coupons. Tensile tests were performed, at room temperature and at the
high temperature 625 C chosen for creep-fatigue testing, on dog-bone specimens with a gage length and diameter of 13 and 5 mm, respectively, at an initial
strain rate of 0:002 s1 . The tensile properties obtained from these tests are

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

FIG. 2(a) Creep-fatigue and (b) creep specimen layout used in the ASTM round-robin (all dimensions are in inches).

28 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

FIG. 3(a) Optical micrograph of P91 steel microstructure with the inset showing inclusions and (b) shows a schematic of the
microstructure 9.

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 29

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

30 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

TABLE 3Tensile test results of P91 steel at room and high temperature (625 C).
Test Temperature
 C= F
24/75
625/1157

0.2 % Yield Strength


(MPa/ksi)

Ultimate Tensile Strength


(MPa/ksi)

Elongation
(%)

532.6/77.2
325.1/47.1

708.4/102.7
343.7/49.8

26
33

listed in Table 3 and were found to be comparable to those published in the literature for P91 steels.
Creep tests were performed on smooth round specimens with a gage length
and diameter of 25 and 5 mm, respectively, which were designed as per the
ASTM E139 standard [8] [see Fig. 2(b)]. All tests were carried out at 625 C (898 K).
Creep tests were carried out under uniaxial static (constant stress and temperature) loading conditions in a lab-controlled atmosphere (2062 C and 50 % relative
humidity). The external static load was applied using dead weights and a calibrated
LVDT with a repeatability of 0:1 lm was employed to measure elongation during
the tests. The test temperature was monitored continuously during the entire span
of the tests using two K-type thermocouples wound mechanically at the top and
bottom ends of the specimen gage length. The temperature difference between the
top and bottom thermocouples during any test was continuously monitored and
found to be within the allowable limits of 62 C of the test temperature.
The LarsonMiller parameter PLM is used as a predictive parameter to
evaluate the stress level required for a given rupture time [10] and is denoted
analytically as
PLM Tk  CLM logtr =1000

(1)

where:
Tk test temperature (K),
tr creep  rupture time (h), and
CLM constant (30 for P91 steel).
The PLM plotted against the applied engineering stress (external static load
divided by the initial cross-sectional area of the specimen) r in MPa is used to
obtain the value of r for a given tr . For this round-robin, the test matrix was
designed such that tr ranges from 5000 to 150 h and the corresponding stress
levels range from 102 to 152 MPa (see Fig. 4). The data from the tests conducted
as part of this program are plotted with the data from the literature [11] in Fig.
4 and were found to lie perfectly within the trends of the literature data, thus
once again conrming that the test material behavior is within the limits of typical P91 steels.
The creep deformation mechanism of P91 steels can be dominated depending on stress and temperature, by cross-slip and dislocation climb > 70 MPa,
and by grain boundary diffusion < 70 MPa [12]. It has been found by
others that for martensitic steels, the metallurgical changes are of vital
importance as they strongly affect creep resistance properties leading to loss in
creep rupture strength [13]. From a microstructural perspective, it has been
Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 31

FIG. 4Larson Miller parameter plot from Ref 11, overlaid with data from the current
round-robin tests at 625 C.

observed that the M23 C6 precipitates in P91 link up by rapid directional coarsening during the primary creep regime. The resistance to creep deformation by
hindering of dislocation motion is increased by the evolution of this microstructure during the nal stages of primary creep and during the entire secondary
creep regime [14].
Creep strain versus time curves exhibited very short primary ( 10 % of tr )
creep followed by secondary or steady-state and substantial tertiary creep stages
under the tested conditions (see Fig. 5). Because of dynamic creep recovery
due to subgrain growth and dislocation migration/annihilation in grain
sub-boundaries, the primary creep strain rate starts to decrease until it reaches
a plateau at the onset of secondary or steady-state creep. The secondary creep
regime represents a dynamic equilibrium between work hardening and creep
recovery processes, wherein a balance between generation of new dislocations
and annihilation of existing dislocations is achieved [15]. The hindrance to dislocation motion due to the evolving precipitates mentioned above must also be
part of the overall dynamic equilibrium during secondary creep stage. The minimum creep rate _e ss as observed in the secondary creep regime is linearly t
using Nortons power law _ ss Arn (where A and n are material constants) and
the power law exponent n and constant A are found to be  8 and
9:5 1021 , respectively (see Table 4).
The onset of tertiary creep is characterized by accelerating creep rates as a
result of a combination of the following reasons [16]:
(1) increasing stress experienced in localized region(s) of the material due
to necking phenomenon
(2) increasing creep strain accumulation that leads to cavity formation and
growth and consequent damage that develops in the form of creep cavities over time

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

FIG. 5Creep rupture (experimental versus modeled) curves for the completed ASTM round-robin tests of P91 steel at 625 C.

32 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 33

TABLE 4Steady-state creep rate as a function of stress for P91 test material at 625 C.
Applied Stress
(MPa)
101.5
130.0
136.8
138.3
142.7
151.5

Steady-State Creep
Rate h1
0.000 35
0.001 66
0.003 81
0.005 23
0.007 79
0.007 48

Power Law Constants


n 8:24 A 9:53 1021

(3) microstructural changes that are not associated with the accumulated
damage
The creep rupture ductility for P91 steel is high with the nal longitudinal
elongation varying from 1619 %. It can also be observed from Fig. 5 that the
creep strains associated with the primary and the secondary creep regimes are
relatively smaller than those associated with the tertiary creep regime, especially at the higher stress levels. Localized necking occurs at times past  95 %
of the creep rupture time (i.e., tr ) when macro-cracking also appears and contributes further to rupture elongation. Similar such observations have been
noted in other studies for this class of steels [9,12,13].
Optical microscopic analysis of ruptured creep specimens showed that the
cracking mode during creep rupture is a predominantly transgranular (ductile)
fracture with an array of microvoids growing through grains along crystallographic planes towards the nal rupture location [see Fig. 6(a)]. It is possible
that such a failure mode is due to plastic deformation at high stresses typically
used in accelerated laboratory testing. It is further noted that these arrays were
found most commonly near material inclusions [precipitates and/or secondary
phases, see Fig. 6(b)].

FIG. 6Transgranular (ductile) fracture mode as observed in P91 steel (a) before and
(b) after etching with Nital (3 % nitric acid in methanol) solution (Test condition:
151.5 MPa, 625 C).

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

34 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

Creep Deformation Model


In this section, the creep deformation and rupture data are represented using
mathematical equations also known as extrapolation models. Such models are
essential for data analysis and structural analysis. Since the creep data presented on P91 clearly exhibit primary, secondary, and tertiary creep behaviors,
the focus of this exercise was on choosing a model that is capable of representing the creep behavior in all three regimes and one that also accounts for
changes in deformation kinetics as a function of stress and temperature. Since
all tests performed were at one temperature, the temperature capability of the
chosen model cannot be veried from the current data set. A compilation of the
existing models as applicable to any high temperature material and P91 steel in
particular is provided in Ref [17]. After carefully evaluating all of these models,
the authors found that the logarithmic creep strain prediction (LCSP) model to
be most suitable for representing the data on P91steel [18]. This model has the
least number of tting constants and its mathematical form has the natural
shape of a creep curve at constant stress. However, in its current form, this
model suffers from two shortcomings:
(1) The model does not reduce to the correct form when the boundary
conditions at the start and at the end of any creep test are applied (i.e.,
as one asymptotically approaches time t 0 and t tr ).
(2) Further, the model does not take full advantage of parameters that can
be measured from the actual tests to reduce the number of tting
constants.
Hence, a modied LCSP model is proposed here by addressing the above
shortcomings of the original model (see Eq 2). As per the modied model, at
any time t in a creep test beyond 1 h, engineering creep strain et is given as
 
1=p

logtr b
C
 1 fj
logt fs
logt b

(2)

where:
fs

p 
p

p

b
log1 C  logru C ; fj logru C
logtr

pr; T p0 p1 logr p2 =T 273 and b, p0 ; p1 ; p2 , and C are tting


parameters,
T test temperature  C,
r applied external stress (MPa),
t engineering creep strain at time t,
tr time to failure by creep rupture (h),
t time to given engineering creep strain (h), and
e1 ; eru et at time t 1 h and uniform strain at rupture, respectively.
As noted earlier, typically from 95 % of tr and beyond, macro-cracking starts
to appear because of localized necking and results in an unstable specimen

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 35

FIG. 7Pictorial representation to empirically measure the uniform strain at rupture,


eru . Vertical lines numbered 1, 2, and 3 indicate locations where diameter is measured
to compute dunif . Untested specimen in the top is kept as a reference for the creep ruptured specimen below.

response thereafter. We also know that the reduction in area varies signicantly
based on the necking characteristics. Hence, the engineering creep strain at
rupture (t at time t tr ) is taken to be the measured uniform strain at rupture,
ru (i.e., strain generated in the specimens gage length before the onset of
necking in a localized region). This strain can be empirically obtained by evaluating the reduction in area Ared in the specimen gage length in regions away
from the neck. Rather than using the conventional ductility equation for Ared ,
an analytical form as shown in Eq 3 is employed in this work. As shown in
Fig. 7, the diameter of the specimen at locations 1, 2, and 3 is measured and
averaged to obtain dunif and this value is used to compute Aunif :
Ared

Aorig  Aunif
Aorig

(3)

where:
Aorig original cross-sectional area in the specimen gage length and
Aunif uniform cross-sectional area obtained by using dunif (the average of
diameters in regions away from the neck).
Minimizing human measurement errors and the effect of complex strain
proles across the specimen gage length were the primary reasons for using the
average value of dunif rather than using a single diameter value in the
unnecked region. The choice of position of locations to compute dunif does not

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

36 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

TABLE 5Modied LCSP model tting constants for P91 steel at 625 C.
Parameter

Value
0.27
1.65
9.15
2553.74
6.8

b
p0
p1
p2
C

seem to play a signicant role, as long as they are chosen to be approximately


equidistant (to capture the diameter variation across the specimen gage length
of the creep ruptured specimen) away from the neck to the edge of the gage
length markers. Also, it was observed in this work that a choice of three such
locations is optimal and a larger number of locations does not improve the
consistency in the average value of dunif .
To perform non-linear regression (NLR) of the available creep rupture data,
the SOLVERV software routine, available as an additional plug-in in Microsoft
Excel, is employed in this work. A few of the tested creep rupture data were
randomly chosen and few commercial NLR softwares including SOLVERV were
used to non-linearly t this data by minimizing the squares of errors (difference
between the actual and predicted strain values). It was found that SOLVERV
did equally as well, if not better, in all of these cases. Thus, the performance of
this package as compared with the commercially available NLR softwares was
carefully analyzed and evaluated before its eventual selection. This routine
searches for the set of values of the tting factor constants or parameters that
minimize the sum of squares of residuals, i.e., of differences between each
observation and the corresponding prediction. From the available test data for
P91 steel, the modied LCSP model provided the tting constants as shown in
Table 5. As can be seen, the model has relatively fewer tting constants as compared to the other models and is consequently more robust [18]. The tted
strain versus time histories are compared to actual measurements in Fig. 5. The
model successfully describes the data at various stress levels. Figure 8 demonstrates the tting efciency of the modied LCSP model for P91 steel by comparing predicted and measured creep rates. Once again, the prediction seems to
correlate well with the experimental data.
R

Creep-Fatigue Tests
The creep-fatigue round-robin test matrix has been designed to address precision in the results obtained from tests when utilizing the procedure outlined in
the new standard E2714-09. More specically, the objective is to quantify intraand inter-laboratory variability in the creep-fatigue data. It is anticipated that
the standard will be revised as warranted by the results of the round-robin. For
example, questions such as what percentage drop in the force provides the most
consistent measure of the number of fatigue cycles for crack formation are

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

FIG. 8Predicted creep strain rate by modied LCSP model for P91 steel as compared with that of experimental values for two different test conditions at 625 C.

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 37

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

38 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

FIG. 9Pilot tests results from creep-fatigue testing of P91 steel, where Na is the number of cycles to failure. Data courtesy of Dr. Stuart Holdsworth of EMPA, Switzerland
and Dr. Yukio Takahashi of CRIEPI, Japan.

expected to be addressed. As noted earlier, 16 laboratories worldwide are


involved in conducting the creep-fatigue testing as per the test matrix given in
Table 1. The creep-fatigue tests are being conducted at 625 C at strain amplitudes of 60.25, 60.5, and 60.75 %. The hold times for the tests are 0, 10, and
30 min at the peak tensile strain. The strain rate chosen for loading and unloading in a triangular wave-form will be 0.025 % per second. The participants are
asked to report the number of cycles at various level drops in force starting with
a 2 % drop. To determine the creep-fatigue round-robin test conditions, pilot
tests were conducted by Dr. Stuart Holdsworth of Swiss Federal Laboratories
for Materials Science and Technology (EMPA), Switzerland and Dr. Yukio
Takahashi of Central Research Institute of Electric Power Industry (CRIEPI),
Japan and the results of these pilot tests are shown in Fig. 9.
These results demonstrate that the selected test conditions reported in
Table 1 meet the criteria for round-robin testing. The test temperatures
and hold times chosen are sufciently high for creep-fatigue interactions to be
present during the testing.
Summary
In order to further rene the precision and bias statements for the recently
developed ASTM standard E2714-09 for creep-fatigue testing, a round-robin
program has been initiated on P91 steel with the participation of 16 laboratories

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

KALYANASUNDARAM ET AL., doi:10.1520/JAI103712 39

worldwide. The specimen machining, material characterization, and pilot test


studies to nalize test parameters for this round-robin have been completed. As
part of this round-robin, this paper also reports the results of microstructural
analysis, tensile and creep deformation, and rupture tests performed thus far on
P91 steel. Preliminary microscopic analysis of ruptured creep specimens shows
that material undergoes predominantly transgranular (ductile) fracture with an
array of microvoids found most commonly near precipitates and/or secondary
phases. A modied LCSP model has been successfully developed to accurately
represent the creep deformation characteristics of this material with the
available data. Some preliminary results from the creep-fatigue tests are also
presented.

Acknowledgments
The writers would like to acknowledge the technical assistance offered by Jeff
Mincy, University of Arkansas and Jeff Metz, Struers in metallographic
specimen preparation and related work. The data offered by NIRM, EMPA, and
CRIEPI (used in Fig. 9) are also highly appreciated.
References
ASTM E2714-09, Standard Test Method for Creep-Fatigue Testing, Annual Book
of ASTM Standards, Vol. 03.01, ASTM International, West Conshohocken, PA.
[2]
Parker, J. (P91 material offered by Kent K. Coleman, EPRI) private communication, 2009.
[3]
ASTM A33510b, Standard Specication for Seamless Ferritic Alloy-Steel Pipe for
High-Temperature Service, Annual Book of ASTM Standards, Vol. 01.01, ASTM
International, West Conshohocken, PA.
[4]
ASTM A21310a, Standard Specication for Seamless Ferritic and Austenitic
Alloy-Steel Boiler, Superheater, and Heat-Exchanger Tubes, Annual Book of ASTM
Standards, Vol. 01.01, ASTM International, West Conshohocken, PA.
[5]
Sikka, V. K., Ward, C. T., and Thomas, K. C., Modied 9Cr-1Mo SteelAn
Improved Alloy for Steam Generator Application, Ferritic Steels for HighTemperature Applications, A. K. Khare, Ed., Am. Soc. Met, Metals Park, OH, 1983,
pp. 6584.
[6]
Hald, J., Metallurgy and Creep Properties of New 912 % Cr Steel, Steel Res., Vol.
67, 1996, pp. 369374.
[7]
Hasegawa, Y., Ohgami, M., and Okamura, Y., Advanced Heat Resistant Steel for
Power Generation, R. Viswanathan and J. Nutting, Eds., The University Press, Cambridge, United Kingdom, 1999, p. 655.
[8]
ASTM E13906, Standard Test Methods for Conducting Creep, Creep-Rupture,
and Stress-Rupture Tests of Metallic Materials, Annual Book of ASTM Standards,
Vol. 03.01, ASTM International, West Conshohocken, PA.
[9]
Fournier, B., Sauzay, M., Barcelo, F., Rauch, E., Renault, A., Cozzika, T., Dupuy, L.,
and Pineau, A., Creep-Fatigue Interactions in a 9 Pct Cr-1 Pct Mo Martensitic Steel:
Part II. Microstructural Evolutions, Metall. Mater. Trans. A, Vol. 40, 2009, pp. 330341.
[10] Larson, F. R. and Miller, J., A TimeTemperature Relationship for Rupture and
Creep Stresses, Trans. ASME, Vol. 74, 1952, pp. 765775.
[1]

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

40 JAI  STP 1539 ON CREEP-FATIGUE INTERACTIONS

[11]

[12]

[13]

[14]

[15]
[16]

[17]
[18]

Gold, M., Tanzosh, J., Swindeman, R. W., Maziasz, P. J., and Santella, M. L., Safe
Use Limits for Advanced Ferritic Steels in Ultra-Supercritical Power Boilers,
CRADA Final Report No. ORNL00-0598, U.S. Department of Energy, Washington,
D.C., 2003, pp. 110.
Gaffard, V., Besson, J., and Gourgues-Lorenzon, A. F., Creep Failure Model of a
Tempered Martensitic Stainless Steel Integrating Multiple Deformation and
Damage Mechanisms, Int. J. Fract., Vol. 133, 2005, pp. 139166.
Sklenicka, V., Kucharova, K., Svoboda, M., Kloc, L., Bursik, J., and Kroupa, A.,
Long Term Creep Behavior of 912 % Cr Power Plant Steels, Mater. Charact., Vol.
51(1), 2003, pp. 3548.
Spigarelli, S., Cerri, E., Bianchi, P., and Evangelista, E., Interpretation of Creep
Behavior of a 9Cr-Mo-Nb-V-N (T91) Steel Using Threshold Stress Concept, Mater.
Sci. Technol., Vol. 15, 1999, pp. 14331440.
Orlova, A. and Cadek, J., Dislocation Structure in High Temperature Creep of Metals and Solid Solution, Mater. Sci. Eng., Vol. 77, 1986, pp. 118.
Annigeri, R., 1997, Life Prediction Methodology for Thermal-Mechanical Fatigue
and Elevated Temperature Creep Design, Ph.D. dissertation, Pennsylvania State
University, PA.
Creep-Resistant Steels, F. Abe, T.-U. Kern, and R. Viswanathan, Eds., Woodhead
Publishing Ltd., Abington Hall, Cambridge, United Kingdom, 2008, p. 700.
Holmstrom, S., 2010, Engineering Tools for Robust Creep Modeling,  Ph.D. dissertation, The Aalto University School of Science and Technology, Espoo, Finland.

Copyright by ASTM Int'l (all rights reserved); Tue Mar 3 15:51:13 EST 2015
Downloaded/printed by
(INSTITUTO POLITECNICO NACIONAL) pursuant to License Agreement. No further reproductions authorized.

Das könnte Ihnen auch gefallen