Vous êtes sur la page 1sur 7

Proceedings of the ASME 2012 Pressure Vessels & Piping Conference

PVP2012
July 15-19, 2012, Toronto, Ontario, CANADA

PVP2012-78575
EFFECTIVE MODELING OF FATIGUE CRACK GROWTH IN PIPELINES

Steven J. Polasik
Det Norske Veritas (USA), Inc.,
5777 Frantz Rd
Dublin, Ohio 43017
Tel: +1 614 761 6985
Email: Steven.Polasik@dnv.com

Carl E. Jaske
Det Norske Veritas (USA), Inc.,
5777 Frantz Rd
Dublin, Ohio 43017
Tel: +1 614 761 6916
Email: Carl.Jaske@dnv.com

ABSTRACT

INTRODUCTION

Pipeline operators must rely on fatigue crack growth models


to evaluate the effects of operating pressure acting on flaws
within the longitudinal seam to set re-assessment intervals. In
most cases, many of the critical parameters in these models are
unknown and must be assumed. As such, estimated remaining
lives can be overly conservative, potentially leading to
unrealistic and short reassessment intervals. This paper describes
the fatigue crack growth methodology utilized by Det Norske
Veritas (USA), Inc. (DNV), which is based on established
fracture mechanics principles. DNV uses the fracture mechanics
model in CorLAS to calculate stress intensity factors using the
elastic portion of the J-integral for either an elliptically or
rectangularly shaped surface crack profile. Various correction
factors are used to account for key variables, such as strain
hardening rate and bulging. The validity of the stress intensity
factor calculations utilized and the effect of modifying some key
parameters are discussed and demonstrated against available
data from the published literature.

One critical aspect to maintaining and managing the integrity


of any pipeline is assessing for the possibility of pressure cycle
fatigue-induced failure. Determining susceptibility to pressure
cycle fatigue involves at a minimum knowledge of the
pipeline manufacture, failure history and pressure cycle severity.
Once an operator determines the susceptibility, a suitable crack
identification methodology is chosen, such as a hydrostatic test
or an in-line inspection (ILI). Regardless of the methodology
selected, the operator must demonstrate that the pipeline
continues to operate in a safe condition. This is typically
accomplished by estimating fatigue crack growth and then reinspecting the pipeline within a timeframe chosen to minimize
detrimental fatigue crack growth, but also to minimize
inspection expenditures.
FATIGUE CRACK GROWTH ASSESSMENT
The general process DNV follows to estimate the remaining
fatigue life of crack-like defects is presented in Figure 1. The

Copyright 2012 by ASME

process illustrated in Figure 1 can be viewed as an extension of


the general approach for engineering critical assessment
developed previously [1]. The fracture mechanics model utilized
applies to a pipeline with an axially-oriented surface crack [2].
Internal pressure loading is considered such that only the
pressure-induced hoop stress is evaluated; discontinuities and
other stress concentrators are not assessed.

effective flaw with a semi-elliptical depth profile. In a manner


similar to that reviewed by Kiefner and Vieth [6], the effective
flaw is determined by evaluating all combinations of the depth
measurements that could possibly make up the flaw and using
the one predicted to result in the lowest failure pressure using
both the fracture mechanics model and flow stress. Figure 2
compares a detailed profile measured from a crack-like flaw to
the calculated effective flaw. Note that the effective flaw (the
one with the lowest predicted failure pressure) does not extend
for the entire length of the detailed profile nor does it have the
same maximum depth; yet, this effective flaw predicts the
lowest failure pressure out of all combinations of the detailed
measurements.

Characterize
Characterize Flaw
Flaw
Type
Type &
& Size
Size

Detailed
Detailed Flaw
Flaw
Profile?
Profile?

100%
Initial Depth Profile

Define
Define Equivalent
Equivalent
Semi-Elliptical
Semi-Elliptical Flaw
Flaw

90%

Define
Define Effective
Effective
Semi-Elliptical
Semi-Elliptical Flaws
Flaws

Effective Flaw

80%

(based
(based on
on length
length and
and depth)
depth)

Depth, %WT

70%

Compute
Compute Critical
Critical
Flaw
Flaw Size
Size

Paris
Paris Equation
Equation Fatigue
Fatigue
Model
Model
n
da/dN
=
C
K
da/dN = C Kn

60%
50%
40%
30%

Characterize
Characterize Operating
Operating
Pressure
Pressure Cycles
Cycles

20%
10%

Calculate
Calculate Estimated
Estimated
Remaining
Remaining Fatigue
Fatigue Life
Life

0%
0

FIG. 1 APPROACH FOR ESTIMATING REMAINING


FATIGUE LIFE

10

15

20
25
Axial Location, in

30

35

40

45

FIG. 2 RESULTS OF THE EFFECTIVE FLAW ANALYSIS


FROM A DETAILED FLAW DEPTH PROFILE

A fatigue crack growth assessment of a pipeline typically


begins by characterizing the loading spectrum, which is then
used to estimate remaining life [3]. The initial size of existing
defects can be based on ILI, excavation data or previous
hydrostatic tests. Fracture mechanics models are then used to
determine the final (critical) flaw size and remaining life is then
determined using a Paris Law approach [4] to estimate the time
required for the flaw to grow from the initial size to the critical
size.

When a detailed depth profile is not available, typically for


the case of an in-line inspection (only depth and length are
reported) or a hydrostatic pressure test (the depth and length of
remaining flaws must be inferred from the hydrostatic test
pressure), the profile is modeled as either a semi-ellipse or a
rectangle. In both cases, the crack-like flaw is evaluated using a
flow-strength based (f) criteria and a J fracture toughness-based
(Jc) criteria. As mentioned previously, the criterion giving the
lowest estimated failure pressure is used to predict failure. In the
absence of a detailed flaw profile, it is possible for the predicted
failure pressure to be more conservative than when a detailed
profile is used. For instance, it is easy to imagine the case where
an in-line inspection survey reported only the total length and
maximum depth of the flaw shown in Figure 2. For this
particular case the effective flaw used in the fracture mechanics
model would be significantly larger than the effective flaw from
the detailed profile analysis. The details of the model for
predicting failure are outside the scope of this paper; instead,
this paper will discuss the details of the model as it applies to
fatigue assessments.
Once the initial flaw is characterized, it is possible to
determine the (final) critical flaw size. As the fatigue assessment
model assumes that flaw growth is confined to depth (i.e., the
flaw length remains constant), the critical flaw size is
determined by calculating the depth at which a flaw with the

Rainflow Cycle Counting


Rainflow counting was developed to relate variable
amplitude fatigue loading to constant amplitude fatigue data. By
using simplified rainflow cycle counting, as described in ASTM
E1049 [5], on a representative pressure history, the number of
cycles that occurred for a specific pressure range is obtained.
The pressure range is then related to the stress in the pipe wall
through the use of Barlows formula.
Flaw Characterization
There are two methods to characterize the flaw shape [2].
The method that is used depends on the availability of a detailed
flaw profile. If a detailed flaw profile is available, the flaw-depth
profile is evaluated along its length and then approximated as an

Copyright 2012 by ASME

same effective length has a failure pressure equal to the


maximum (alternatively, at-site) operating pressure. This
approach is reasonable for typical flaws in pipelines because
they have large length to depth ratios. If the initial length-todepth ratio is small, then the change in this ratio would have to
be accounted for in the crack-growth calculation.

where Qf is a flaw shape factor [9,10,14], Fsf is a free surface


factor that depends on crack length and depth [14], a is the crack
depth, is the local stress that includes the Folias factor, E is the
modulus of elasticity, f3(n) is the Shih-Hutchinson [11,14]
solution that depends on strain-hardening exponent (n), and p is
the local plastic strain computed from using the stress-strain
curve. To the extent that the local stress is an effective value,
such as von Mises, the plastic strain is an effective value.

Flaw Growth
Perhaps the most common method used to estimate the
remaining fatigue life of crack-like flaws is the Paris Law
Method [4]. The Paris Law Method is a power law that describes
the incremental crack growth as a function of the stress intensity
factor range and has the general form:

da
= CK n
dN

Application to Fatigue
If the plastic zone that forms around the crack tip during
fatigue is sufficiently small, then linear elastic fracture
mechanics (LEFM) can be employed and the conditions at the
crack tip are uniquely defined by the stress intensity factor, K.
The crack growth rate is then characterized by Kmin and Kmax.
LEFM performs well in these cases.
The application of inelastic fracture mechanics to fatigue
crack growth is relevant only to those cases where significant
plastic deformation occurs. As summarized elsewhere [9],
fatigue crack growth under these conditions can be characterized
via the range in the J-integral, J, under the following
assumptions:
Temperature is constant,
Environmental factors are negligible,
Increment of crack growth is related to each cycle of
loading,
Three-dimensional crack is well approximated by a
two-dimensional crack model,
Damage or material cracking is assumed to take
place within a small process zone near the crack tip
(analogous to restricting inelastic deformation to a
process zone near the crack tip in LEFM),
Crack length must be much larger than the process
zone size.
J characterizes K, as the two are related [12]. More
specifically, this relationship is given by Equation 4:

(1)

where a is the depth of a surface flaw, N is the number of cycles,


da/dN is the incremental crack growth per cycle, C and n are
material constants, and K is the stress intensity factor range. As
discussed previously, a critical component of the fatigue
assessment includes a characterization of the pressure loading
spectrum since K is dependent on the magnitude of the
operating pressure and the number of cycles. The estimated
remaining life is simply the time required for the crack-like flaw
to grow from its initial depth to the final (critical) depth and is
determined by numerically integrating Equation 1.
FRACTURE MECHANICS MODEL
In order to estimate the value of K used in Equation 1,
DNV employs the fracture mechanics model briefly discussed
above in terms of characterizing the flaw. The fracture
mechanics model is the same as that developed and implemented
in CorLAS. CorLAS was developed to predict the failure
pressure of axial surface crack-like defects in cylindrical
components and has shown good agreement with both
experimental and actual failures. As discussed within a critical
review [7] of various crack assessment methodologies, the
fracture mechanics model within CorLAS produced the best
results, especially considering that actual flaw profiles can be
used to determine an effective equivalent flaw.
As it relates to fatigue, an effective stress intensity factor,
Keff, is calculated using the elastic portion of the J-integral for an
elliptical (or rectangular) flaw geometry with the effective flaw
dimensions. The total J-integral (Jt) for surface flaws is
described in detail in the CorLAS manual [8], and is
expressed as the sum of the elastic (Je) and plastic (Jp)
components:

Jt = Je + J p

J =

Q f Fsf a 2
E

(2)

+ Q f Fsf af 3 (n) p

(4)

where E = E for plane stress conditions and E = E / (1 - 2) for


plane strain, KI is the range in Mode I stress intensity factor
and is Poissons ratio. For typical operating conditions, pipe
dimensions and material properties, axial surface flaws in
pipelines can be approximated to be under plane stress
conditions.
As shown previously in Equation 2, the Jt for surface flaws
can be expressed as a sum of the elastic and plastic components.
Under high-cycle fatigue conditions, it is reasonable to assume
that the plastic component of Jt is negligible. Consequently, the
J in Equation 4 is comprised only of the Je component from
Equation 2, which is the first quantity in Equation 3.
Combining Equation 3 and Equation 4 yield:

Equation 2 can be further expanded:

Jt =

K I2
E

K = loc Q f Fsf a

(3)

(5)

Copyright 2012 by ASME

where loc is the local stress range. Recall that the standard
form of the Mode I stress intensity factor relates a correction
factor, Y, to the crack depth, a, and the far-field stress, , as in
Equation 6.

Initial Depth Profile


Effective Flaw Case 1
80%

(6)

M surf

50%
40%

20%
10%
0%
0

(7)

10

15

20
25
Axial Location, in

30

35

40

45

FIG. 3 EFFECTIVE FLAW ANALYSIS AND FINAL FATIGUE


PROFILE

As seen in Figure 3, the effective flaw estimated by the


CorLAS model for the entire initial flaw profile (Case 1) is
shorter in length and shallower in depth than the initial depth
profile. The final flaw measured in this particular example
extended by fatigue and occurred in a location similar to that of
the effective flaw, but the effective length is much longer. If, on
the other hand, the effective flaw is estimated only for the region
in which crack extension occurred (Case 2), the effective flaw is
approximately the same as the initial depth with a shorter total
length. As the stress intensity factor is more heavily influenced
by depth it is approximately 4% greater for Case 2 than for
Case 1.
In a 1982 study, Mayfield and Maxey [13] examined the
fatigue characteristics of the weld zone of electric resistance
weld (ERW) line pipe as well as the base metal. Pipe samples
were cyclically pressurized and crack growth from machined
notches was measured via the electric potential drop technique.
This study provides a method to compare the crack growth
measured to the growth estimated via the model described
herein.
Before examining the comparison of measured crack growth
in a pipe to the predictions, it is important to briefly discuss the
evolution of CorLAS and some of the components within the
model.
The original formulation of the fracture mechanics model in
CorLAS (Version 1) was found to provide good predictions of
actual SCC failure conditions, but gave somewhat conservative
predictions for very long cracks. In 2001, a PRC International
(PRCI) research project was initiated to develop improvements
to the CorLAS (Version 2) fracture mechanics model.
Version 2 included improved formulations for calculating values
of the J-integral for semi-elliptical surface flaws using published
results of finite element stress analysis [14]. As mentioned
previously, Rothwell and Coote [7] reviewed various crack
assessment methodologies, and determined that the fracture
mechanics model within CorLAS (i.e., Version 2) produced
the best results.
The transition from Version 1 to Version 2 involved changes
to multiple terms; plasticity was not included in Qf, but was

a
4 WT M
=
a
1
4 WT

60%

30%

where is the range in the far-field stress computed using


Barlows Formula and Msurf is the Folias factor for an elliptical
surface flaw (see Equation 8). Msurf is related to the Folias factor
used by Kiefner and Vieth [6], M, as shown in Equation 8,
where WT is the pipe wall thickness. The modification /4 is
needed to account for the area of an ellipse; for a rectangular
flaw, this factor is unnecessary.

Effective Flaw Case 2

70%

The correction factor Y is related to the quantities Qf and Fsf


as well as the transformation of to loc. The factor Qf is a
combination of an elliptical shape correction factor, Q, and the
surface-crack correction factor for an edge crack (1.122). This
factor is dependent on the ratio of the length to the depth. The
factor Fsf is computed from empirical relationships. Finally, loc
is defined in Equation 7.

loc = M surf

Final Depth Profile

90%

Depth, %WT

K I = Y a

100%

(8)

As the fracture mechanics model in CorLAS was


developed using a flat plate solution, the incorporation of the
Folias factor into loc applies a multiplication factor to account
for the additional stress intensity owing to the bulging effect that
occurs in cylinders, but is not present in the flat plate solution.
DISCUSSION
Anecdotally, DNV has seen good correlations between
calculated and actual fatigue lives from hydrostatic pressure
tests. The fatigue lives predicted using the model in these cases
appear reasonable, particularly for estimated fatigue lives based
on available detailed flaw profiles. As indicated, DNV has made
limited comparisons of actual fatigue crack profiles from
irregular flaws to the effective areas and depths determined by
CorLAS. In these comparisons, the effective lengths
calculated by CorLAS are shorter than actual lengths. On the
other hand, the calculated depths are larger.
Figure 3 is an extension of Figure 2 in which the effective
flaw (the estimated flaw that predicts the lowest failure pressure)
for two difference cases is compared to the final measured crack
profile and the initial measured flaw profile.

Copyright 2012 by ASME

incorporated elsewhere. As such, Qf (the flaw shape factor), was


modified to eliminate the effective plastic zone correction to
crack size, thereby reducing the elastic component, Je, when
compared between CorLAS Version 1 and Version 2. In
Version 1, Qf was based on the ratio of the far-field stress to the
yield stress (/y) being equal to 1.0. In Version 2, Qf is based
on the same ratio being equal to 0.0.
A comparison between Qf (Version 1) and Qf (Version 2) is
shown graphically in Figure 4. Qf is developed from the flaw
shape parameter for an elliptical crack [9]. Version 1 used the
factor for /y = 1.0 while Version 2 uses the factor for /y = 0.
The value is squared since J K2/E (see Equation 4). As is
expected, for all values of crack depth to crack length ratios
(a/2c), Qf (Version 1) is larger. A review of Equation 5 shows
that the lower Qf in Version 2 leads to lower stress intensity
factors and thus predict longer remaining fatigue lives than
would be predicted when using the Version 1 formulation.

0.250

0.200

Depth, in

0.150

0.050

Test 2 (NG-18 Report 130)


DNV Predicted Qf v2
DNV Predicted Qf v1

0.000
0

4000

6000

8000

10000

12000

14000

16000

FIG. 5 COMPARISON OF MEASURED FATIGUE GROWTH


AND PREDICTED GROWTH FOR NG-18 REPORT 130 TEST 2

Qf v1
Qf v2

1.60

0.250

Paris Law Coefficient: 2.1210-11 in/cycle ksi in


Paris Law Exponent: 4.0
Outside Diameter: 12.75 inch
Wall Thickness:
0.25 inch
Initial Depth:
0.050 inch (20% WT)
Initial Length:
2.0 inch
max:
72% SMYS

1.40
1.20
0.200

Qf

2000

Cycles

2.00
1.80

Paris Law Coefficient: 2.1210-11 in/cycle ksi in


Paris Law Exponent: 4.0
Outside Diameter: 12.75 inch
Wall Thickness:
0.25 inch
Initial Depth:
0.09123 inch
Initial Length:
2.0 inch
max:
50% SMYS

0.100

1.00
0.80
0.150

Depth, in

0.60
0.40

0.100

0.20
0.00
0.00

0.10

0.20

0.30

0.40

0.50

0.050

Test 3 (NG-18 Report 130)


DNV Predicted Qf v2
DNV Predicted Qf v1

a/2c

FIG. 4 COMPARISON OF FLAW SHAPE FACTORS


VERSION 1 AND VERSION 2

0.000
0

2000

4000

6000

8000

10000

12000

14000

Cycles

Figure 5 and Figure 6 compare the experimental results from


Mayfield and Maxey [13] for Test 2 and Test 3,
respectively, and the method described in this paper. In Test 2,
the maximum stress was 50% of the specified minimum yield
strength (SMYS) and it took approximately 7000 cycles before
the flaw started to show growth. The maximum stress in
Test 3 was 72% SMYS. The cycles in the figures have been
adjusted to reflect only the crack growth. Figure 5 and Figure 6
contains estimates for fatigue crack growth using the Version 1
and Version 2 formulations for Qf (all other components within
the Je calculation are the same). In both figures, the cycles have
been adjusted to only reflect crack growth. As seen in Figure 5
and Figure 6, the Version 2 formulation underestimates the
fatigue crack growth. Version 1 appears to match the actual
fatigue crack growth well for Test 3, but underestimates the
growth for Test 2. This indicates that the Qf component of the
CorLAS fracture mechanics model when applied to fatigue
using only the elastic portion of the J integral may need to
incorporate the plastic zone size correction factor.

FIG. 6 COMPARISON OF MEASURED FATIGUE GROWTH


AND PREDICTED GROWTH FOR NG-18 REPORT 130 TEST 3

Shen et al [15] simulated pipes and plates with an axial


rectangular crack with filleted corners under fatigue loading
using FEA and compared these results to those predicted using
BS 7910 and the Newman and Raju solutions for a plate with a
semi-elliptical crack.
In Figure 7, the stress intensity factors for semi-elliptical
flaws from both Version 1 and Version 2 of CorLAS as well
as estimates from API RP 579 [16] are overlaid with the results
compiled by Shen [15]. For shallow cracks, the results between
all of the methods appear to agree, with increasing scatter as the
crack depth increases and the largest scatter around 80% of the
wall thickness. For crack depths around 50% of the wall
thickness, all of the methods predict a lower stress intensity than
the BS 7910 methodology with both versions of CorLAS
showing the lowest value. At the largest crack depth, both
CorLAS Version 1 and API RP 579 [16] appear to predict
very similar stress intensities. A comparison of the FEA results
for both the Plate and Pipe scenarios in Figure 7 supports the
inclusion of the Folias factor in the CorLAS methodology.
5

Copyright 2012 by ASME

factor is less for a semi-elliptical crack in a pipe


than for a rectangular crack in a pipe, as is expected.

140

120

Plate (Newman & Raju)

cylinder (BS7910)

API RP 579

Plate (FEA)

Pipe (FEA)

CorLAS (Qf v2) Semi-ellipse

REFERENCES

CorLAS (Qf v1) Semi-ellipse

[1] Jaske, C. E., and Beavers, J. A. 2002, Development and


Evaluation of Improved Model for Engineering Critical
Assessment of Pipelines, Paper IPC02-27027,
Proceedings of ASME 2002 International Pipeline
Conference.
[2] Jaske, C. E., Polasik, S. J., and Maier, C. J., 2011,
Inelastic Fracture Mechanics Model for Assessment of
Crack-Like Flaws, Paper PVP2011-57099, Proceedings
of ASME 2011 Pressure Vessels and Piping Conference.
[3] Beavers, J. A, Maier, C. J., Jaske, C. E., and Bubenik, T.
A., 2006, Analysis of Crack-Like Flaws and Relative
Ranking of Pipeline Segments for Prioritizing Mitigation
Efforts, Paper IPC2006-10181, Proceedings of ASME
2006 International Pipeline Conference.
[4] Paris, P. C., Gomez, M. P., and Anderson, W. E., 1961, A
Rational Analytic Theory of Fatigue, The Trend in
Engineering, Vol. 13, No. 1, pp. 9-14.
[5] Anon., 2011, Standard Practices for Cycle Counting in
Fatigue Analysis, ASTM E1049-85(2011), American
Society for Testing and Materials, West Conshohocken,
PA.
[6] Kiefner, J. F., and Vieth, P. H., 1993, The Remaining
Strength of Corroded Pipe, Paper 29, Proceedings of the
Eighth Symposium on Line Pipe Research, Catalog No.
L51680, American Gas Association, Inc., Arlington, VA.
[7] Rothwell, A. B., and Coote, R. I., 2009, A Critical Review
of Assessment Methods for Axial Planar Surface Flaws in
Pipe, Proceedings of Pipeline Technology 2009,
Oostende, Belgium, October 12-14.
[8] Jaske, C. E., 1996, CorLAS 1.0 User Manual Computer
Program for Corrosion-Life Assessment of Piping and
Pressure Vessels, Version 1.0, CC Technologies Systems,
Inc., Dublin, OH.
[9] Jaske, C. E., 1984, Damage Accumulation by Crack
Growth Under Combined Creep and Fatigue, Ph.D.
Dissertation, The Ohio State University, Columbus, OH.
[10] Jaske, C. E., 1986, Estimation of the C* Integral for
Creep-Crack-Growth Test Specimens, The Mechanisms
of Fracture, ASM International, Materials Park, OH, pp.
577-586.
[11] Shih, C. F., and Hutchinson, J. W., 1975, Fully Plastic
Solutions and Large Scale Yielding Estimates for Plane
Stress Crack Problems, Report No. DEAP S-14, Harvard
University, Cambridge, MA.
[12] Anderson, T. L., 2005, Fracture Mechanics: Fundamentals
and Applications, Taylor & Francis, Boca Raton, FL.
[13] Mayfield, M. E., and Maxey, W. A., 1982, ERW Weld
Zone Characteristics, Final Report to AGA/PRCI, Catalog
No. L51427, Pipeline Research Council International, Falls
Church, VA.
[14] Jaske, C. E., and Beavers, J. A., 2001, Integrity and
Remaining Life of Pipe with Stress Corrosion Cracking,
Final Report PR 186-9709, Catalog No. L51928, Pipeline
Research Council International, Falls Church, VA.

100
G318 (X-46)
c = 25 mm (0.984 in)
L = 50 mm (1.969 in)
Ri = 157.2 mm (6.189 in)
OD = 324 mm (12.75 in)
WT = 4.8 mm (0.189 in)
h=0.8SMYS = 1091 psig

K, ksiin

80

60

40

20

0
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

a, in

FIG. 7 COMPARISON OF STRESS INTENSITY FACTORS


FOR AXIAL CRACKS IN PLATES AND PIPES

CONCLUSIONS
The discussion above indicates the following:
The CorLAS fracture mechanics model that forms
the basis of the K calculations was developed to
model failure of axial crack-like defects and
accurately predicts failure,
Comparisons to measured fatigue crack growth for
two machined flaws in pipe demonstrate that the
current formulation of Je, and therefore K
(Version 2) under-estimates fatigue crack growth,
CorLAS Version 2 was modified to remove
plasticity from the components that make up Je, with
plasticity being accounted for in Jp,
The current formulation of Je does not incorporate a
plastic zone correction factor to the crack depth in
the Qf component,
When Qf does incorporate the plastic zone
correction factor (i.e., CorLAS Version 1), the
estimated fatigue crack growth error for these two
cases is reduced,
Version 1 and Version 2 are expected to provide an
upper and lower bound for the stress intensity factor
as these formulations either ignore the plastic zone
size (/y = 0.0) or fully take it into account
(/y = 1.0),
Both versions of Qf predict lower stress intensity
factors than those estimated by both BS 7910 and
API RP 579 [16] for the cases evaluated,
The incorporation of the Folias factor to account for
the additional stress owing to the bulging effect that
occurs in cylinders (it is not incorporated in the flat
plate solution) appears to be justified by FEA
analysis,
Comparisons of semi-elliptical flaws to rectangular
cracks via the FEA analysis are less conclusive, but
do demonstrate that the predicted stress intensity

Copyright 2012 by ASME

[15] Shen, G., Adeeb, S. M., Coote, R. I., Horsley, D. J., Tyson,
W. R., Gianetto, J. A., and Bouchard, R., 2006, Fatigue
Crack Driving Force for Axial Surface Cracks in Pipes,
Paper IPC2006-10177, Proceedings of ASME 2006
International Pipeline Conference.

[16] Anon., 2007, Fitness-For-Service, Standard API 5791/ASME FFS-1, American Petroleum Institute,
Washington, DC.

Copyright 2012 by ASME

Vous aimerez peut-être aussi