You are on page 1of 9

ARTICLE IN PRESS

International Journal of Pressure Vessels and Piping 85 (2008) 80–88


www.elsevier.com/locate/ijpvp

Factors influencing creep model equation selection$


S.R. Holdswortha,, M. Askinsb, A. Bakerc, E. Gariboldid, S. Holmströme, A. Klenkf,
M. Ringelf, G. Mercklingg, R. Sandstromh, M. Schwienheeri, S. Spigarellij,
(on behalf of Working Group 1 of the European Creep Collaborative Committee)
a
EMPA, Dübendorf, Switzerland
b
RWE, Swindon, UK
c
British Energy Generation Ltd., Barnwood, UK
d
Politecnico di Milano, Italy
e
VTT, Espoo, Finland
f
MPA, Stuttgart, Germany
g
ISB, Milan, Italy
h
KIMAB, Stockholm, Sweden
i
IfW, Darmstadt, Germany
j
Università Politecnica delle Marche, Ancona, Italy

Abstract

During the course of the EU-funded Advanced-Creep Thematic Network, ECCC-WG1 reviewed the applicability and effectiveness of
a range of model equations to represent the accumulation of creep strain in various engineering alloys. In addition to considering the
experience of network members, the ability of several models to describe the deformation characteristics of large single and multi-cast
collations of e(t,T,s) creep curves have been evaluated in an intensive assessment inter-comparison activity involving three steels, 214CrMo
(P22), 9CrMoVNb (Steel-91) and 18Cr13NiMo (Type-316).
The choice of the most appropriate creep model equation for a given application depends not only on the high-temperature
deformation characteristics of the material under consideration, but also on the characteristics of the dataset, the number of casts for
which creep curves are available and on the strain regime for which an analytical representation is required. The paper focuses on the
factors which can influence creep model selection and model-fitting approach for multi-source, multi-cast datasets.
r 2007 Elsevier Ltd. All rights reserved.

1. Introduction  to co-ordinate the generation, exchange/collation and


assessment of creep data in Europe;
The European Creep Collaborative Committee (ECCC)  to interact with, and supply information to European
was formed in 1992 to provide European industry (in standards organisations and their technical committees;
particular alloy producers, power plant manufacturers and  to mutually exchange technical information relaying to
electricity utilities) with a means of influencing the creep current and future activities on material developments;
rupture strength values required for the large number of and
European product and design standards in preparation  to develop common rules for creep data generation,
during the early 1990s. Initially, the principal aims of exchange/collation, and assessment.
ECCC were
The Technical Working Group of ECCC (WG1)
$
develops methods and guidelines for the generation,
This article appeared in its original form in Creep & Fracture in High collation/exchange, and assessment of creep data [1b–e].
Temperature Components: Design & Life Assessment Issues, 2005.
Lancaster, PA: DEStech Publications, Inc. With regard to assessment, the initial focus of this group
Corresponding author. Tel.: +41 44 823 4732; fax: +41 79 290 8242. was the consideration of large, multi-source, multi-cast,
E-mail address: stuart.holdsworth@empa.ch (S.R. Holdsworth). multi-temperature creep-rupture datasets to predict long-time

0308-0161/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijpvp.2007.06.009
ARTICLE IN PRESS
S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88 81

Nomenclature tf ðT; sÞ time to a specific creep strain as a function of


temperature and stress
ECCC European Creep Collaborative Committee T temperature
n stress exponent Z parameter quantifying effectiveness of master
p time exponent creep equation to predict times to specific
Q activation energy for creep strains (see Eq. (2))
R universal gas constant e, ee, ei strain, elastic strain, instantaneous plastic
Rpe/t/T, Ru/t/T creep strength and rupture strength for a strain
given time and temperature ef, ep, eper creep strain, plastic strain, permanent strain
sA-RLT standard deviation of residual log times _; _f;min ; _ave strain rate, minimum creep strain rate,
t time average strain rate
tu, tu,max observed time to rupture, maximum observed s stress
time to rupture se(t,T) stress to give a specific strain as a function of
tpe/s/T, tp=s=T observed and predicted times to given time and temperature
plastic strain o; o _ damage, rate of damage accumulation

H1 (continuous)
1E+2
H2 (continuous)
600°C 165 164 140 120 H3 (continuous)
150
170 140 115100 H4 (continuous)
H5 (interrupted)
180
1E+1 H6 (interrupted)
STRAIN, %

90
80
1E+0
80
72
56
50

1E-1

1E-2
1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
TIME, h

Fig. 1. Continuous and interrupted measurement creep curves available for 6 casts of Steel-91 at 600 1C (curves with data points are the result of
interrupted measurement tests; labels are test stresses in MPa).

(X100 kh) strength values for standards. More recently, Many different model equations are used to represent creep
attention has moved to the assessment of creep strain data strain accumulation characteristics ranging from simple
[2,3]. The requirement for a representative description of a phenomenological to complex constitutive (e.g. [2–26]). Their
material’s ef(t,T,s) creep strain behaviour is no longer just effectiveness to model primary, secondary, and/or tertiary
for scientific interest and metallurgical understanding.1 The creep deformation for specific applications can vary with
creep deformation behaviour of engineering components is material characteristics and source data distribution. For
now routinely evaluated using PC-based finite-element example, not all model equations and fitting procedures are
analysis tools. Design engineers require the parameters for suitable for the prediction of mean long time creep strength
model equations to describe the long-time creep behaviour behaviour. Assessment inter-comparisons have been per-
of a specified alloy (not simply the characteristics of a formed on large multi-source, multi-cast, multi-temperature
single cast), typically in the primary and secondary datasets to establish the most effective approaches for the
deformation regimes. In contrast, remaining life assessment description of mean-alloy (rather than single cast) creep
engineers are more likely to require the best model deformation behaviour in evaluations based on working
description for a single cast of material, in the secondary datasets established for Steel-91 and Type-316.
and tertiary creep regimes. Prior to the start of the large multi-cast dataset evaluation
activity, a first creep data assessment inter-comparison
1
A list of nomenclature is given in Nomenclature. The terminology used examined a single-cast 214CrMo (P22) dataset [2,3]. The T,s
is as recommended in the ECCC Recommendations volume 2 [1b]. conditions for the creep curves in this dataset were uniformly
ARTICLE IN PRESS
82 S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88

75 100 90 80 78
1E+2 125 123 63 125
98
700°C 190 165
98 50
90 62
49

62
1E+1

STRAIN, %
40
1E+0

1E-1 123b
123e
123m
123o

1E-2
1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
TIME, h

Fig. 2. Interrupted measurement creep curves available for 4 casts of Type-316 at 700 1C (low-strain sections of curves without data points determined by
continuous measurement; labels are test stresses in MPa).

distributed and the study, involving the application of 16 Table 1


model equations by 9 analysts, resulted in the development Classical representations of primary, secondary and tertiary creep
of a parameter (Z) to provide a measure of model-fitting Model equation Source reference Used in
effectiveness in specific creep strain regimes. As a generality, Eq.a
specific model equations are better suited to representing
creep strain accumulation characteristics for a given material Primary creep
Power law: f ¼ Atp ! f ¼ Asn tp Graham [17] (A.5)
in either the primary/secondary regime or the secondary/
Logarithmic: ef ¼ B log(1+bt) Phillips [18]
tertiary regime, although some models can be suitable for Exponential: ef ¼ C(1exp(ct)) McVetty [19] (A.1)–(A.4),
both. The Z-parameter provides a means of quantifying (A.9)
model-fitting effectiveness and is defined as Sinh law: f ¼ D sinhðct1=3 Þ Conway [20]

Z ¼ 102:5:sARLT , (1) Secondary creep


Power law: _f;min ¼ A0 sn Norton [21] (A.1)–(A.4)
where sA-RLT is the standard deviation of residual log time Sinh law: _f;min ¼ _0 sinhðs=s0 Þ Nadai [22]
for specific creep strains. Ideally for single-cast assessments Tertiary creep
Z-2, whereas ZX4 is unacceptable [2,3]. The applicability Exponential: f ¼ MðexpðmtÞ  1Þ McHenry [23] (A.1), (A.4)
of Z-parameter to large multi-source, multi-cast, multi- asn csk Rabotnov- (A.2)–(A.5)
_f ¼ _ ¼
; where o
temperature datasets has also been evaluated and the ð1  oÞq ð1  oÞr Kachanov [24,25]
outcome is considered below.
Omega: _f ¼ _0 expðOÞ Sandström- (A.9)
Kondyr [26]
2. Working datasets
a
See Appendix A.
ECCC-WG1 has employed 3 working datasets for their
consideration of creep strain assessment. Details of the An added assessment complication was that some of the test
single-cast 214CrMo working dataset have been given else- records were the results of continuously monitored unin-
where [2,3]. Details of the multi-cast Steel-91 and Type-316 terrupted creep tests and some were the result of interrupted
datasets are summarised in the following sections. tests.2 While this situation can be typical for a large multi-
source, multi-cast, multi-temperature dataset collated to
2.1. Steel-91 provide representative long-time alloy creep strength values
for standards purposes, it added to the analytical challenge
Steel-91 is a 9CrMoVNb martensitic stainless steel. The through the provision of some creep curves comprising
Steel-91 dataset used in the second WG1 creep strain
assessment inter-comparison comprised a total of 90
ef(t,T,s) and ep(t,T,s) creep test records representing the (footnote continued)
constant load (or stress) to a uniaxial testpiece held at constant
behaviour of 6 casts at 11 temperatures in the range temperature [3]. In continuous-measurement tests, the creep strain, ef, is
450–650 1C.2 The maximum test duration exceeded 50 kh. monitored without interruption by means of an extensometer attached to
the gauge length of the testpiece. In interrupted tests, the total plastic
2
Creep strain ef(t) or ep(t) curves are determined from the results of strain, ep, is measured optically at room temperature during planned
continuous measurement or interrupted tests involving the application of a interruptions (i.e. ep ¼ ei+ef, where ei is instantaneous plastic strain [1b]).
ARTICLE IN PRESS
S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88 83

Table 2
Summary of Z values associated with model-fits to Steel-91 dataset

Model equation Fitting approach Z values for times to specified strains in %

0.2 0.5 1.0 2.0 5.0

A.1 Modified theta model [4] 1 74 19 17 8 12


A.2 BJF model [5] 1 113 104 50 32 11
A.4 Li–Akuluv model [7,8] 1 143 42 18 6 4
A.5 Modified Graham–Walles model 3 7 7 6 7 5
A.7 Bolton model [10] 3 38 7 7 6 5
A.8 MHG model [11–13] 2 6 7 6 6 5
A.9 Modified omega model [14,15] 1 138 19 11 8 5
A.10 Modified Sandström model [16] 2 11 7 6 6 6

many individual ef(t) co-ordinates and some with only a secutively for individual deformation regimes) to
small number of ep(t) co-ordinates (e.g. Fig. 1). establish the model parameters for specific conditions
of T and s.
2.2. Type-316 (ii) Determination of the temperature and stress depen-
dence of the selected model parameters to define the
Type-316 steel is a 18Cr13NiMo austenitic stainless steel. material mean master equation for all ef(t,T,s).
The Type-316 dataset used in the third WG1 creep strain
assessment inter-comparison comprised a total of 98 ep(t,T,s) Approach 2:
interrupted creep test records representing the behaviour of 4
casts at 6 temperatures in the range 500–700 1C (e.g. Fig. 2). (i) Determination of specific tf ðt; sÞ co-ordinates from
A wide variation in strength between the four casts was individual experimental creep curves either directly
observed. The maximum test duration exceeded 130 kh. In (unconstrained by a formal model description) or as a
lower temperature, higher stress tests in particular, the result of model fitting (with a model different to that
magnitude of instantaneous plastic strain was a significant used for final fitting).
proportion of the total plastic strain. (ii) Parametric model fitting of the specific co-ordinates to
establish parameters to define mean master equation
3. Creep strain data assessment for material in the form of e(,,s) or tf ðT; sÞ:

3.1. Model equation options Approach 3:

A wide range of creep model equations are in use today (i) Derivation of mean s ðt; TÞ relationships from specific
to represent the high-temperature time-dependent defor- observed tf ðT; sÞ co-ordinates from individual experi-
mation behaviour of engineering materials (e.g. [2–16], mental creep curves.
Appendix A). Many of these comprise components (ii) Model fitting with derived s ðt; TÞ to establish para-
originating from a small number of classical representa- meters for selected model to define mean master
tions of primary, secondary, and/or tertiary creep deforma- equation for material.
tion (e.g. [17–26], Table 1).
The multi-cast creep data assessment inter-comparisons
involved the application of 10 models by 10 different 3.3. Steel-91
analysts. These models are listed in Appendix A.3
Approach 1 was used to fit models A.1, A.2, A.4, and
3.2. Model fitting A.9 to the Steel-91 dataset (Table 2). In addition,
Approach 2 was used to fit models A.8 and A.10, with
Different approaches may be adopted to model-fit Approach 3 being used to fit models A.5 and A.7. The Z
ef(t,T,s) creep strain behaviour. values achieved with the 8 assessments are summarised in
Table 2. The lowest values achieved for the Steel-91 multi-
source, multi-cast dataset are significantly higher than
Approach 1:
those achieved during the assessment of the 214CrMo single-
(i) Model-fitting individual experimental creep curves with cast dataset [2,3]. The inherent cast-to-cast variability in
the selected model equation (simultaneously or con- the Steel-91 data led to best Z values typically in the range
6–7 for times to creep in the primary and secondary
3
The model equations used in the single-cast creep data assessment regimes. This compares with an achievable target of 2 for
inter-comparison are given elsewhere [2,3]. times in the same regimes for a single-cast dataset of creep
ARTICLE IN PRESS
84 S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88

curves generated for a uniform distribution of tempera- establish the parameters for model A.10 (Table 3). This
tures and stresses. The best Z values for the multi-source proved to be a difficult dataset to assess and the activity is
multi-cast dataset were achieved using model-fitting still incomplete. Where available, the Z values associated
approaches 2 or 3 (italicised in Table 2). with the assessments of the multi-cast Type-316 dataset are
Examples of the A.5, A.7, and A.8 model fits determined generally larger than those reported for Steel-91 in Table 2.
by these approaches are compared with experimental creep These are undoubtedly a consequence of the significant
measurements in Fig. 3. Z values of 6–7 may have to be variation in strength between the four casts and uncertain-
regarded as acceptable for model equations representing ties associated with the determination of ei, in particular, at
alloy mean behaviour determined from the results of large higher stresses and lower application temperatures in the
multi-source, multi-cast datasets. creep range for the austenitic stainless steel.
Predicted 10 and 100 kh strength values for (a) 0.2% The lowest Z values obtained in the multi-cast Type-316
and (b) 1.0% creep strain determined using the A.1, creep strain data assessment inter-comparison were
A.5, A.7, and A.8 model equations are given in Fig. 4. achieved by employing an Approach 2 model fitting
There is good agreement for the creep strength values procedure (Table 3).
determined using model-fitting procedures involving Ap-
proach 2 or 3.
4. Creep model selection
3.4. Type-316
The selection of creep model and model-fitting approach
Approach 1 was used to fit models A.2, A.3, and A.6 to can depend on several factors including material char-
the Type-316 dataset, while Approach 2 was used to acteristics, data distribution, and practical application.

100 100
H3 (continuous) H2 (continuous)
H5 (interrupted) H4 (continuous)
H6 (interrupted) Bolton model fit
10 Bolton model fit 10 MHG model fit
MHG model fit Modified Graham-Walles
Mod Graham Walles
STRAIN, %
STRAIN, %

1 1

0.1 0.1

Steel-91, 550°C, 170MPa Steel-91, 600°C, 150MPa


0.01 0.01
0.1 1 10 100 1000 10000 100000 0.01 0.1 1 10 100 1000 10000
TIME, h TIME, h

100 10
H1 (continuous) H4 (continuous)
H5 (interrupted) H6 (interrupted)
H6 (interrupted) Bolton model fit
10 Bolton model fit
MHG model fit
1 Modified Graham-Walles
MHG model fit
STRAIN, %
STRAIN, %

Modified Graham-Walles

0.1
0.1

Steel-91, 600°C, 120MPa Steel-91, 600°C, 80MPa


0.01 0.01
0.1 1.0 10.0 100.0 1000.0 10000.0 100000.0 0.1 1 10 100 1000 10000 100000 1000000
TIME, h TIME, h

Fig. 3. Comparison of creep curve fits to Steel-91 experimental data for (a) 550 1C/170 MPa tests (casts H3, H5 and H6), (b) 600 1C/150 MPa tests (casts
H2 and H4), (c) 600 1C/120 MPa tests (casts H1, H5 and H6), and (d) 600 1C/80 MPa tests (casts H4 and H6).
ARTICLE IN PRESS
S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88 85

400 400
Modified Theta (A1,1) Modified Theta (A1,1)
PREDICTED 0.2% CREEP STRENGTH, MPa

PREDICTED1.0% CREEP STRENGTH, MPa


350 Modified Graham-Walles (A5,3) 350 Modified Graham-Walles (A5,3)
Bolton model fit (A7,3)
Bolton model fit (A7,3)
MHG model fit (A8,2)
MHG model fit (A8,2)
300 300

250 250 10kh

10kh
200 200

150 150 100kh

100 100kh 100

50 50

0 0
450 500 550 600 650 450 500 550 600 650
TEMPERATURE, °C TEMPERATURE, °C

Fig. 4. Comparison of predicted 10 and 100 kh strength values, were determined for all casts of Steel-91, (a) R0.2/t(T) and (b) R1.0/t(T) (open points are
predicted strength values for 10 kh; solid points are predicted strength values for 100 kh; the number following the model equation designation in the
legend label parentheses refers to the employed model-fitting approach as defined in Section 3.2).

Table 3
Summary of Z values associated with model-fits to Type-316 dataset

Model equation Fitting approach Z values for times to specified strains in %

0.2 0.5 1.0 2.0 5.0

A.2 BJF model [5] 1 – 26 66 25 13


A.3 Modified Garofalo model [6] 1 tbd tbd tbd tbd tbd
A.6 Baker–Cane model [9] 1 tbd tbd tbd tbd tbd
A.10 Modified Sandström model [16] 2 14 10 8 9 6

tbd—Still ‘to be determined’ by all-cast assessment.

4.1. Material characteristics interrupted measurement tests it is measured indirectly in a


companion hot tensile test.
The effectiveness of a creep equation to represent a
material’s characteristic e(t) curve shape can depend on 4.2. Data distribution
features such as the relative proportions of primary,
secondary, and tertiary creep as fractions of the strain Model selection and the choice of model-fitting approach
and time at rupture (e.g. Fig. 5a), and the way in which can be influenced by the distribution of the data to be
they vary over the T,s regime of interest (Fig. 5b). The assessed. A creep dataset typically consists of a number of
Z-parameter (Eq. (1)) provides a useful indication of model e(t,T,s) curves (creep test records), each comprising a
effectiveness [2,3]. For well-distributed, single-cast e(t,T,s) number of e(t) observations. Both the e(t,T,s) curve and
creep curve datasets, Z values of -2 are achievable. For e(t) datapoint distribution characteristics are influential.
multi-source, multi-cast datasets such as those for Steel-91 The WG1 single-cast 214CrMo working dataset provided a
and Type-316 investigated in the WG1 assessment inter- good example of a set of creep curves, which were uniformly
comparison, Z values of better than 6–7 are not possible. distributed in terms of temperature and stress (curves for 6
The magnitude of instantaneous plastic strain relative to stresses at each of 5 temperatures), with each creep curve
total plastic strain can be significant, in particular, for comprising the same number of e(t) observations (400)
austenitic steels at low application temperatures in the [2,3]. Nevertheless, there are still potential difficulties with
creep range. In contrast, at high application temperatures, the analysis of such a dataset when either primary or tertiary
ei is usually small and often disregarded. In continuous creep dominates and effective model fitting the entire creep
measurement tests, ei is measured directly whereas for curve can be compromised by an imbalance in the number
ARTICLE IN PRESS
86 S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88

1.0 1.0
450C, 520MPa
450C, 430MPa
650C, 120MPa

NORMALISED PLASTIC STRAIN


NORMALISED PLASTIC STRAIN

0.8 0.8 650C, 80MPa


Type-316
650°C

0.6 0.6

0.4 0.4
2¼CrMo
540°C
0.2 Steel-91 0.2
600°C

0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
NORMALISED TIME NORMALISED TIME

Fig. 5. Examples of creep curve shape variations for (a) 214CrMo, Steel-91 and Type-316 at typical application temperatures and (b) Steel-91 over a wide
T,s regime.

of observations in each regime. In such circumstances, it is mean creep behaviour for well-specified alloys from large
necessary to redress the balance of e(t) data points or to multi-source, multi-cast strain-time datasets.
model fit each regime independently [2,3]. As a generality, specific model equations are better
In contrast, the multi-cast Steel-91 dataset comprised 90 suited to representing creep strain accumulation character-
creep curves for 6 casts of the steel. The temperatures and istics of a given material in either the primary/secondary
stresses for which the creep curves were generated were not regime or the secondary/tertiary regime, although some
uniformly distributed and the number of e(t) observations models can be suitable for both.
per curve ranged between 2 and 250, the creep curves Certain model-fitting approaches are more applicable to
having been generated by interrupted and continuous large datasets, in particular, those involving a mixture of
measurement testing. For such datasets, Approach 2 or 3 continuous and interrupted measurement creep test records
appears to be the most appropriate procedures for model with a wide spectrum of ef(t,T,s) data points per unit time.
fitting (Tables 2 and 3), both approaches effectively Such approaches effectively average the multi-cast ef(t,T,s)
averaging the multi-cast ef(t,T,s) data in a balanced way data in a balanced way prior to model fitting.
prior to final model fitting. Whereas values for the model-fitting effectiveness para-
meter of Z-2 are achievable for single-cast creep strain
4.3. Practical application datasets which are well distributed in terms of (T,s), Z
values of 6–7 may have to be regarded as acceptable for
Model selection can also depend on the purpose for model equations representing alloy mean behaviour
which the materials creep strain description is required. determined from the results of large multi-source, multi-
Typically, the priority of the scientist is for a creep model cast datasets.
equation to have a sound physical basis in the primary,
secondary, and tertiary regimes. As a generality, it is more
Appendix A. Creep strain equations adopted in assessment
important for design and assessment engineers for the
inter-comparisons
model equation to be simple to implement and effective in
its description of creep deformation at long times (e.g. A.7
A.1. Modified theta model [4]
[10]). For design engineers, effective modelling is more
important in the relatively low-strain primary/secondary
creep regimes whereas for remaining life assessment f ¼ y1 ð1  expðy2 tÞÞ þ y5 t þ y3 ðexpðy4 tÞ  1Þ,
engineers, the priority is more likely to be an accurate where logðyi Þ ¼ ai þ bi T þ ci s þ d i Ts
knowledge of secondary/tertiary behaviour.
and y5 ¼ Asm expðQ=RTÞ. ðA:1Þ
5. Summary
A.2. BJF model [5]
Results are reviewed of an ECCC work programme to
investigate procedures for the practical representation of f ¼ n1 ð1  expðtÞb þ n2 tÞ,
ARTICLE IN PRESS
S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88 87

Z
the P91 case F(e,s) is of the form: a0 þ a1 log s2 þ
where t ¼ ðseff =A1 Þn expðQ=RT Þ dt
a2 log 3 :
s  s m
seff ¼ and o _ ¼B . ðA:2Þ
1o 1o A.9. Modified omega model [14,15]

 
1 1
A.3. Modified Garofalo [6] f ¼  ð lnðtu  tÞ þ lnðtu ÞÞ
O 2C tr
þ C tr ð1  expðmtr tÞÞ. ðA:9Þ
per ¼ i þ f1max ½1  expðDðt=t12 Þu Þ
þ _ p;min t þ c23 ðt=t23 Þf . ðA:3Þ
A.10. Modified Sandström [16]

Pnpol
A.4. Li-Akuluv model [7,8] ðF ð;sÞþf 2 ðTÞ=f 1 ðTÞð1=f 1 ðTÞÞ ajþ1 log sj Þ
t ¼ 10 j¼0 , (A.10)
  where F ð; sÞ ¼ b1 log 4 þ b2 log 2 þ b3 T log s þ b4 T þ
_f;min _i  _f;min
f ¼ ln 1 þ ð1  expðktÞÞ þ _s t b5 log s and any time temperature parameter satisfying
k _ f;min TTPðt; TÞ ¼ f 1 ðTÞ log t þ f 2 ðTÞ can be used.
þ T ðexpðt=tt Þ  1Þ. ðA:4Þ
References

[1] ECCC Recommendations. Creep data validation and assessment


A.5. Modified Graham–Walles procedures. In: Holdsworth SR, et al., editors. ECCC. publ., (a) vol.
1: Overview, (b) vol. 2: Terms and terminology, (c) vol. 3: Data
  acceptability criteria, data generation, (d) vol. 4: Data exchange and
sð1 þ Þ n1 m1 collation, (e) vol. 5: Data assessment, (f) vol. 6: characterisation of
_ f ¼ eðQ1=TÞ 10A1  þ eðQ2=TÞ microstructure and physical damage for remaining life assessment, (g)
1o
 n2 vol. 7: Data assessment—creep crack initiation, (h) vol. 8: Data
A2 sð1 þ Þ assessment—multi-axial, (i) vol. 9: Component assessment.
 10 [2] Holdsworth SR, Merckling G. ECCC developments in the assessment
1o
of creep-rupture data. In: Proceedings of sixth international Charles
where Parsons Conference on engineering issues in turbine machinery,
power plant and renewables, Trinity College, Dublin, 16–18
o ¼ eðQD=TÞ 10AD ðsð1 þ ÞÞnD mD . (A.5) September, 2003.
[3] Holdsworth SR. Developments in the assessment of creep strain and
ductility data. Mater High Temp 2004;21(1):125–32.
[4] Evans RW, Wilshire B. Creep of metals and alloys. Institute of
A.6. Baker–Cane [9] Metals; 1985.
[5] Jones DIG, Bagley DL. A renewal theory of high temperature creep
and inelasticity. In: Proceedings of I. mech. E conference on creep
f ¼ Atm þ p þ fs þ s ðl  fÞ and fracture: design and life assessment at high temperature, MEP,
  15–17 April, 1996. p. 81–90.
t=tu  f 1f=lf [6] Granacher J, Möhlig H, Schwienheer M, Berger C. Creep equation
 l ðA:6Þ for high temperature material. In: Proceedings of seventh interna-
1f
tional conference on creep and fatigue at elevated temperatures
where l ¼ u =s ; s ¼ _m tu and f ¼ tp =tu . (Creep 7), 3–8 June, NRIM, Tsukuba, 2001. p. 609–16.
[7] Li JCM. A dislocation mechanism for transient creep. Acta Metall
1963;11:1269.
A.7. Bolton characteristic strain model [10] [8] Akulov NS. On dislocation kinetics. Acta Metall 1964(12):1195.
[9] Baker AJ, O’Donnell MP. R5 high temperature structural integrity
ðRu=t=T =R=t=T Þ  1 assessment of a cracked dissimilar metal weld vessel test. In:
f ¼  . (A.7) Proceedings of second international conference on integrity of high
ðRu=t=T =sÞ  1 temperature welds, 10–12 November 2003, London.
[10] Bolton JL. Design considerations for high temperature bolting. In:
Strang A, editor. Proceedings of conference on performance of
bolting materials in high temperature plant applications, York,
A.8. MHG model [11–13] 16–17/6/94. 1994; p. 1–14.
[11] Grounes M. A reaction rate treatment of the extrapolation methods
in creep testing. J Basic Eng Ser D Trans ASME 1969.
t ¼ expðTF ð; sÞ þ CÞ, (A.8) [12] Manson SS, Haferd AM. A linear time-temperature relation for
extrapolation of creep and stress rupture data. NASA TN 2890 1953.
where the F(e,s) function is freely selected from multilinear [13] Holmström S, Auerkari P. Prediction of creep strain and creep
combinations of s and e with an optimised value of C. For strength of ferritic steels for power plant applications. In: Proceedings
ARTICLE IN PRESS
88 S.R. Holdsworth et al. / International Journal of Pressure Vessels and Piping 85 (2008) 80–88

of Baltica conference on life management and maintenance for power [20] Conway JB, Mullikin MJ. An evaluation of various first stage creep
plants, VTT Symposium 234, 8–10 June, Espoo. equations. In: Proceedings of AIME conference, Detroit, Michegan,
[14] Prager M. Development of the MPC Omega method for life assessment 1962.
in the creep range. ASME J Press Vessel Technol 1995(May):95–103. [21] Norton FN. The creep of steel at high temperature. New York:
[15] Merckling G. Metodi di calcolo a confronto per la previsione McGraw-Hill; 1929.
dellulteriore esercibilità in regime di scorrimento viscoso. In: [22] Nadai A. The influence of time upon creep. The hyperbolic sine creep
Proceedings of conference on fitness for service, Giornata di Studio law. Stephen Timoshenko anniversary volume. New York: Macmil-
CESI-CONCERT, Milan, 28 November, 2002. lan; 1938.
[16] Sandström R. Extrapolation of creep strain data for pure copper. [23] McHenry D. A new aspect of creep in concrete and its application to
J Test Eval 1999;27:31–5. design. Proc ASTM 1943;43:1069.
[17] Graham A, Walles KFA. Relations between long and short time [24] Kachanov LM. Izv Akad Navk SSR 1958;8:26.
properties of commercial alloys. JISI 1955;193:105. [25] Rabotnov YN. Creep problems in structural members. Amsterdam:
[18] Phillips F. The slow stretch in india rubber, glass and metal wire North-Holland; 1969.
when subjected to a constant pull. Philos Mag 1905;9:513. [26] Sandström R, Kondyr A. A model for tertiary creep in Mo- and
[19] McVetty P G. Factors affecting the choice of working stresses for CrMo-steels. In: Proceedings of third international conference on
high temperature service. Trans ASME 1933;55:99. mechanical behaviour of materials, Cambridge, 1979.

You might also like