Sunteți pe pagina 1din 8

Assessment of Fatigue Damage from Variable Amplitude Loads in Risers

Fengjie Yin, Mark Cerkovnik


2H Offshore Inc., Houston, TX, USA

Al Conle
University of Windsor, Windsor, ON, Canada

ABSTRACT
The in-service fatigue loading of risers may include
contributions from first and second order vessel motions, wave
loadings, vortex induced vibration (VIV), vortex induced motion
(VIM), pressure variation and slugging. All of these loadings are
variable amplitude random loadings where the sequence of cycles can
change the fatigue outcome. The loading sequence effects can be
addressed in analysis through cycle counting methods and through use
of the most appropriate methods of damage accumulation. High loads
can be responsible for either fatigue crack growth retardation or
acceleration depending up loading sequence. The Rainflow Counting
method takes into account the whole loading algorithm and assumes the
structure memory sustains all through the loading. The Simple Range
Count method counts cycles by every single reversals and does not take
into account the load sequence or material memory. Fatigue damage
accumulation can be accomplished using a linear model like PalmgrenMiner or a nonlinear model. Fatigue life can be estimated using crack
growth methods, S-N or strain life methods. For each method some
tools are available that can account for sequence effects. The fatigue
analysis of the riser system is illustrated step by step. This study
particularly looks at how different cycle counting methods affect the
fatigue damage estimate obtained with the S-N approach, strain-life
approach and fatigue crack growth approach. In the study Rainflow and
Simple Range cycle counting methods are applied to variable amplitude
loads typical of deepwater riser systems.

KEY WORDS: variable amplitude loading, cycle counting, Rainflow


counting, Simple Range counting, load sequence, material memory,
Palmgren-Miner.

INTRODUCTION
Risers experience variable amplitude fatigue loading caused by first
and second order vessel motions, wave loadings, vortex induced
vibration (VIV), vortex inducted motion (VIM), pressure variation and
slugging. Developing a realistic representation of service load is a key
step to successful fatigue analysis and design of deepwater riser
systems.

Fatigue analysis of risers usually includes the following steps: 1.


Conduct a computer simulation of the global riser system or acquire
real-time monitoring data describing the response of the riser system; 2.
Generate stress time traces all along the riser; 3. Conduct cycle
counting with the stress time histories to decompose the variable
amplitude loading history into a spectrum of simple constant amplitude
loading cycles (stress histograms); 4. Calculate fatigue damage and
predict fatigue life contributed by each bin in the stress histograms
using S-N, fatigue crack growth, or strain-life methods; 5. Assess the
fatigue damage or the life of the structure by the Palmgren-Miner linear
damage rule (Palmgren, 1924, Miner, 1945) or nonlinear damage rule.
If the crack growth approach is adopted, then fatigue crack growth life
is calculated instead of step 4 and step 5.
STRESS TIME HISTORIES
There are usually two approaches to obtain a stress time histories for
risers: 1. Computer simulation with finite element analysis considering
load contributions from first and second order vessel motions, wave
loadings, VIV, VIM, pressure variation and slugging; 2. Real-time
monitoring with pressure and temperature sensors, accelerometers, and
strain gages.
If time domain simulation is conducted to obtain the stress time
histories, long-term sea states or short-term fatigue analysis is usually
proceeded with either vessel motion time traces or RAO (response
amplitude operator). The simulation is normally conducted with three
hour simulation. For each simulation, time traces of the effective
tension and bending moments are extracted at selected locations along
the length of the riser. The stress at any selected location around the
riser circumference as shown in Figure 1 is calculated using the
following equation:
(1)
Where,
F
A

is the axial force


is the pipe wall area;

My
Mz
I
R

is the moment about the local y-axis;


is the moment about the local z-axis;
is the moment of inertia of the cross section;
is the radius
is the angle of rotation anti-clockwise from the local y-axis

every part of every overall range once.


Cycle counting methods were invented in the early years of aircraft
loads monitoring device development in order to summarize long
duration accelerometer or strain signals into a tabular summary which,
in the days prior to computer usage, allowed the signal to both be
characterized in a graphic form and to reduce the number of damage
calculations. After much discussion and testing of the best methods to
use for cycle counting (Downing and Socie, 1982) it became apparent
from following the axial specimen stress-strain deformation, behavior
that the closure of a stress-strain hysteresis loop seemed to be the
appropriate definition of a fatigue damaging event. Matsuishi and
Endo (1968) described the Pagoda Raindrop method of cycle
counting; a name that was eventually altered to Rainflow Cycle
Counting in studies by others. The literature (BS5400, 1980; Downing
and Socie, 1982; Conle, Oxland, and Topper, 1988) has described
several variants of Rainflow counting but all the studies aim at
simulating the closure of stress-strain hysteresis loops; as if the signal
were to be played back as a control for testing an un-notched axially
loaded sample.
In this study theories and sample calculations with two cycle counting
methods are discussed: Rainflow counting and Simple Range counting
methods. The algorithm for standardized Rainflow counting method is
detailed in ASTM E-1049 (2005). The procedure is summarized as
follows: 1. Count flow along roof; 2. Starting from the highest peak
go down to the next reversal. The rain flow runs down and continues
unless the flow reaches the end of the time trace or flow is interrupted
by the previous flow, or flow reaches a starting point of greater
magnitude. 3. Completed flow paths along roofs for a given case are
counted as one half cycle. Starting and ending points counted must be
touching a roof; 4. Half cycles from each case are then added to
combine into a stress histogram.

Figure 1 Circumferential Locations on Riser for Stress Calculation


When the stress time histories are obtained from real-time monitoring,
filtering, screening, and editing of the data obtained is essential. The
reasons are: 1. The measurements from the sensors may have spurious
spike readings which are not experienced by the riser. These false
readings need to be removed from the load time traces; 2. The
measurements may include an event that is not expected to repeat in the
future (for example the high pressure spike from hydro-testing); 3. The
riser may have routine events scheduled / expected to occur in the
future, so they are not captured from the measurements but need to be
added in the time traces.
CYCLE COUNTING
The objective of all kinds of cycle counting methods is to compare the
effect of variable amplitude load histories to fatigue data and fatigue
curves obtained with constant amplitude load cycles. Different cycle
counting methods can change results significantly. There are a few
commonly agreed rules upon good cycle counting methods: a good
cycle counting method must count a cycle with the range from the
highest peak to the lowest valley and seek to count other cycles in a
manner that maximizes the ranges that are counted, that is to say,
intermediate fluctuations are less important than the overall differences
between highest and lowest points; a good counting method shall count

Figure 2 Rainflow Counting Algorithm


One sample time trace is shown in Figure 2 and the result from the
Rainflow counting is given in Table 1. Note that the result is kept
unitless because it could come from counting load, stress, or strain time

histories.
Table 1 Rainflow Counting Result
Cycle
A-D-I
B-C-B
E-H-E
F-G-F

Range
39
9.5
28
5

Simple Range counting is counting every difference between two


successive reversals. The Simple Range counting approach counts both
positive and negative ranges. Each is counted as one-half cycle. The
result from the Simple Range counting with the same time trace as
shown in Figure 2 is given in Table 2.
Table 2 Simple Range Counting Results
Event
A-B
B-C
C-D
D-E
E-F
F-G
G-H
H-I

No. of Cycles
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5

Figure 3 Illustration of Material Memory Event


(Reprinted with Perminssion Reference)
http://fde.uwaterloo.ca/Fde/Crackgrowth/pdpropPaper2013.pdf
A similar path from a test with large cyclic strains is shown in Figure 4.
The presence of material memory events and the resulting closed
hysteresis loops are evident at these large amplitudes as well. Although
there is no study known to the authors of material memory events
observed in the plastic zone of a propagating crack, one could infer that
such events are also important to the correct simulation of a
propagating crack.

The closure of a stress strain hysteresis loop, as shown in Figure 3 is


termed a material memory event (i.e.: a stress strain path that is
interrupted by a signal reversal will be resumed when the signal returns
to that reversal value within certain conditions). In the Figure 3
deformation and loading start at the zero strain and zero stress origin
"O" and proceeds in tension along the monotonic curve to point A,
where the deformation direction reverses towards compression and
reverses again, towards tension at point B. From there the stress-strain
locus moves in the tensile direction to point C where another reversal
occurs and the locus moves to point D. From D the next half-cycle of
strain moves all the way to F. Along this half-cycle however two
memory events occur. The first is at E where the material remembers
the old reversal point C and closes the small elastic hysteresis loop DC. The second event is at the old point A where the hysteresis loop B-A
is closed and the path of the stress-strain locus returns to the original
monotonic path.
A simple range count of the deformation path shown in Fig. 1 would
yield the following ranges: O-A, A-B, B-C, C-D, and D-F. Note that
the stress-strain hysteresis loop half cycle that goes from B to A has
now been split into the piece B-C and another piece D-F. The segment
B-C would yield less predicted damage than the segment B-A when
one counts damage on a half cycle basis. It is also easy to imagine the
effect of several small unloads during the path from B to C. The large
overall half cycle would be divided into smaller, less damaging
segments. Theoretically if there were substantial noise on the signal a
simple range count would degenerate into many small ranges.

Figure 4 Stress-Strain Path with Large Overall Strain Amplitude


Showing Material Memory Events
Simple Range counting ignores the material memory. Rainflow cycle
counting avoids both the segmentation problem and the small noise
effect problem that Simple Range counting method encounters. The
smaller loops that interrupt a large half-cycle are recognized by
Rainflow and saved in the count, but the large half-cycles that are part
of closed hysteresis loops are retained. Rainflow counting method also
retains the material memory. Although it is possible to define a fatigue
damaging event as some other feature of a variable amplitude

waveform, a process that recognizes loop closure memory as fatigue


events generally gives the most satisfactory results.
Cycle counting using Rainflow rules is not a perfect process, however.
Although it sustains the material memory, the sequence of the cycles
has been lost when the closed loop maximum and minimum stresses or
strains are stored in the counting matrix.
For scenarios where fatigue damage is dependent upon frequency, the
methodology of Rainflow with Palmgren-Miner damage accumulation
can be questioned since the method ignores the duration of the counted
half cycles. In the riser application the steel is typically in seawater and
cathodically protected. The inner radius of the pipe may be affected by
CO2 or H2S. In these environments the fatigue life of steel is generally
frequency dependent and the S-N curves are generated with constant
amplitude testing at a frequency similar to the cyclic loading frequency
of the riser.
Variable amplitude testing (OTO 199058, 2000) done in marine
environments indicates that the fatigue damage could be correlated to
the properties developed using constant amplitude lab tests via
Rainflow counting and the Palmgren-Miner linear damage
accumulation methods. Presumably the relations still hold because in
the real world marine environment, the peak and valley points of the
large cycles are generally part of the same wave load cycle.
Without full scale testing one cannot justify that material memory is
lost when the riser system is in service, and consequently the Rainflow
counting method is by far the most popular and well recognized cycle
counting method.
DAMAGE CALCULATION AND LIFE PREDICTION
Overall there are three types of approaches to calculate fatigue life: the
stress-life (S-N) approach, the strain-life (-N) approach, and crack
growth approach. Both the S-N and -N approaches are used to predict
life to crack formation or nucleation (SAE Design Handbook, 1997).
The S-N approach is mostly used in fatigue analysis of a riser system
when the load history is in the elastic range. When the load history has
large overloads, for example, when reel-lay is selected as installation
method, or in mechanical connectors where high stress concentrations
exist, -N approach is generally a better approach than S-N approach.
S-N curves used for the evaluation of a riser system are defined by
equations of the following form. These curves can be implemented with
multiple slopes (DNV RP 203, 2011):

(2)
Where,

N
a,m
k
tref

is the stress range including the stress concentration factor


is the number of cycles to failure for the stress range
are the parameters defined the curves
is the wall thickness correction factor
(for riser wall thickness greater than tref)
is the reference thickness of the thickness correction term

A widely used strain-life prediction with mean stress effect is strain-life


prediction with SWT parameter (Smith, Watson and Topper, 1970):

max a E ( 'f ) 2 ( 2 N f ) 2 b 'f 'f ( 2 N f ) bc

(3)

Where,

max

a
E

is the maximum stress


is the strain amplitude

'f

is the Youngs Modulus


is the fatigue strength coefficient

is the fatigue strength exponent

'
f

c
Nf

is the fatigue ductility coefficient


is the fatigue ductility exponent
is the fatigue life for the combination of strain amplitude and
Maximum stress

In the crack growth approach for riser fatigue assessment, the analysis
is conducted based on the Paris Law approach, as defined by the
Equation (4). Just as for the SN method it can be implemented with
multiple slopes and take mean stress into account:
(4)
da /dN = A (K)m
Where:
da/dN
is the fatigue crack growth rate
K
is the stress intensity range
A,m
are the parameters defining the single or multi stage crack
growth relationship
In both S-N and strain-life approaches, after fatigue life N for the stress
range / strain range is calculated for each bin or count, the damage for
each bin can be calculated given the number of cycles in that bin
computed by cycle counting. The cumulative damage can be calculated
by the Palmgren-Miner rule (Palmgren, 1924, Miner, 1945) or a
nonlinear model (Yin and Fatemi, 2010; Yang and Fatemi, 1998). The
Palmgren-Miner rule is expressed in the following equation:

(5)
Although there is experimental and theoretical evidence that indicates
that linear damage accumulation is not always accurate, PalmgrenMiner rule is still the most commonly used approach in S-N and strainlife fatigue damage calculation not only due to its simplicity but also
because all other nonlinear models require testing data to validate the
input parameters. The complexity of each service load makes it hard to
obtain one single nonlinear model that fits all service histories.
The fatigue crack growth approach is usually adopted in riser design --via engineering critical assessment (ECA). For a riser system already in
service, the fatigue crack growth approach is also selected for fitness
for service (FFS) assessment. In general a fatigue crack growth analysis
with variable amplitude loading depends on the load sequence
especially when isolated overloads occur in the loading history. A
tensile overload may induce compressive residual stress fields that tend
to temporarily retard or arrest crack growth. A compressive overload,
on the other hand, may leave tensile residual stress fields that tend to
accelerate crack growth. In ECA or FFS of risers, the load history is
usually generated from the statistics of the expected service life and the
loading is randomized by assuming that a characteristic histogram or
load block is repeated. The block load histogram is obtained from
global analysis and cycle counting. However when severe overload
occurs (this can happen when a hydro-test is performed) or when a load
amplitudes change magnitude in predominantly one direction (high
blocks followed by low blocks or low blocks followed by high blocks),
fatigue crack growth models incorporating load sequence effects need
to be incorporated. When load a history is highly random, the
assumption is usually made that the hi-lo and lo-hi effects cancel out

each other This ability of the FCG method to account for load sequence
effects has its advantages compared to S-N and strain-life methods.
CASE STUDY
As an example of how the cycle counting method affects the predicted
fatigue life, an analysis is conducted with a deepwater steel catenary
riser (SCR) in the Gulf of Mexico (GOM). The loading is derived from
a 3 hour time domain simulation of a 100 year hurricane. The SCR
configuration is shown in Figure 5. Two locations along the SCR are
identified as fatigue critical locations: 1. The first (top most) weld near
hang off; 2. The touch down point (TDP).

Figure 6 Stress Time Traces for First Weld and TDP under 100 Year
Hurricane

Figure 5 SCR Configuration and Fatigue Critical Locations


The SCR pipe data is given in Table 3.
Table 3 SCR Pipe Data
Parameter
Material
Steel Grade
Nominal OD (mm)
Nominal Wall Thickness (mm)
Yield Strength (MPa)
Ultimate Tensile Strength (MPa)
Elastic Modulus (MPa)
Poissons Ratio

Value
Steel
X70
320
38
482
565
207,000
0.3

The three hour time traces are assumed to repeat in fatigue analysis.
Stress histograms obtained with both Rainflow counting and Simple
Range counting methods are shown in Figure 7 for TDP and Figure 8
for first weld. In smaller stress ranges the difference between the
Rainflow counting and Simple Range counting results is not significant
for these two stress time traces. The difference starts to show in larger
stress ranges which contribute higher fatigue damage. It can be seen
from both Figure 7 and Figure 8 that Rainflow counting results in more
large stress range cycles than Simple Range counting, which would
lead one to conclude that Simple Range counting could result in underconservative results in fatigue damage calculation.

A three-hour time domain simulation is conducted with the SCR and


effective tension and bending time histories are extracted from the
simulation for the two critical locations. Stress time histories are shown
in Figure 6. The tension at the first weld near the hang off location is
high so overall the stress time history at the first weld has a higher
mean stress compared to that at TDP, however the higher bending
results in the occasional high stress at the TDP.
Figure 7 Stress Histograms at TDP (Rainflow Counting vs Simple
Range Counting)

html, and shown in Figure 9, is used for fatigue damage assessment.


The stress concentration factor Kt is increased to 4.0 to allow the
calculations to result in finite lives for better comparison. Note the
nominal stresses shown n the stress time traces are amplified by Kt
and then corrected with a Neuber plasticity correction to determine
the local stresses and strains at the fatigue hot spots. The calculated
fatigue lives with SWT mean stress correction parameter as stated in
Equation (3) are given in Table 7 in years and are given in Table 8 in
number of consecutive storms that could be endured.

Figure 8 Stress Histograms at First Weld (Rainflow Counting vs


Simple Range Counting)
To quantify the difference in fatigue damage and fatigue life
calculations, S-N fatigue lives are calculated based on the DNV E
curve (DNV CN 30.2, 1984). The parameters as defined in Equation (2)
are given in Table 4.
Table 4 S-N Fatigue Curve Details
Fatigue
Material
a
Detail
Steel
DNV'84
1.0354E+12
weld
E

Figure 9 Strain-Life Curve from SAE980 Coupon Testing

tref
(mm)

SCF

3.0

22.0

0.25

1.1

Assuming 100 year hurricane loading is continuous, the S-N fatigue


lives calculated with Rainflow counted histograms and Simple Range
counted histograms for both locations are given in Table 5 in years and
are given in Table 6 in number of consecutive storms that could be
endured.
Table 5 S-N Fatigue Life in Years Calculated with Rainflow and
Simple Range Counted Histograms
S-N Fatigue Life (Years)
Percentage
Cycle Counting Method
Difference
Location
Rainflow
Simple Range
(%)
First Weld
0.473
0.763
61.3
TDP
0.379
0.504
33.1
Table 6 S-N Fatigue Life in No. of Consecutive Storms That Could
be Endured, Calculated with Rainflow and Simple Range Counted
Histograms
S-N Fatigue Life (No. of Consecutive
Percentage
Storms Endured)
Difference
Cycle Counting Method
Location
(%)
Rainflow
Simple Range
First Weld
1381
2227
61.3
TDP
1105
1471
33.1
The strain-life curve of axial coupon specimens machined from
SAE980 with modification with periodic overload, as found in raw
form in References:
http://fde.uwaterloo.ca/Fde/Materials/Steel/Hsla/mergedSAE980.html
http://fde.uwaterloo.ca/Fde/Materials/Steel/Hsla/mergedSAE980_fitted.

Table 7 Strain-Life Fatigue Life in Years Calculated with Rainflow


and Simple Range Counted Histograms
Strain-Life Fatigue Life (Years)
Percentage
Cycle Counting Method
Difference
Location
Rainflow
Simple Range
(%)
First Weld
0.304
0.967
218.1
TDP
0.110
0.210
90.3
Table 8 Strain-Life Fatigue Life in No. of Consecutive Storms,
Calculated with Rainflow and Simple Range Counted Histograms
Strain-Life Fatigue Life (No. of
Percentage
Consecutive Storms Endured)
Difference
Cycle Counting Method
Location
(%)
Rainflow Simple Range
First Weld
887
2822
218.1
TDP
322
613
90.3
Fatigue crack growth analysis with short term 100 year hurricane
fatigue is also conducted. Input parameters for the fatigue crack growth
analysis are summarized in Table 9.

Table 9 Input Parameters for Fatigue Crack Growth Analysis


Value for First
Parameters
Value for TDP
Weld
ECA Assessment
Level 2A
Level 2A
Level
Crack Type
External
External
Initial Crack Depth,
1.0
1.0
mm
Initial Crack Length
2.0
2.0
Crack Tip Opening
Displacement
0.5
0.5
(CTOD), mm
SCF
1.1
1.1
End of Life Load,
380.5
445
MPa
Residual Stress, MPa
482
482
Marine + Cathodic
Marine + Cathodic
Paris Curve
Protection
Protection
The fatigue crack growth curve is defined as the Paris law in Equation
(4), in this study the mean + 2 stand deviation laws for R ratio 0.5 (R
ratio = minimum stress / maximum stress) is used in accordance with
the recommendations in BS7910 (2007). The Paris Curve details are
given in Table 10. This crack growth model is applicable for K above
the threshold value K0 defined in Table 11. For K less than K0,
da/dN is assumed to be zero.

Figure 10 Fatigue Crack Growth Curves for the First Weld (Rainflow
Counting vs. Simple Range Counting)

Table 10 Paris Curve (Steels in a Marine Environment with Cathodic


Protection at -850 mV Ag/Agcl)
Transition Point K
R
A
m
(N/mm3/2)
-17
0.5
290
5.10
2.10 x 10
-11
0.5
2.67
2.02 x 10
Table 11 Fatigue Crack Growth Threshold Stress Intensity Factor
Range Value
Material
Environment
K0 (N/mm3/2)
Marine with Cathodic
Steels, Including
Protection, Up to
63
Austenitic
20C
Initial crack sizes are assumed to be 1mm in depth (a) and 2mm in
length (2c). A Level 2A assessment is conducted per BS7910 (2007).
Extreme loads that define the unstable fracture flaw sizes are obtained
from the maximum stress from the 100 year hurricane load histories for
the first weld and TDP locations, respectively. Histograms calculated
from Rainflow counting and Simple Range counting methods are
applied and the fatigue crack growth curves are compared in Figure 10
for the first weld and in Figure 11 for the TDP.

Figure 11 Fatigue Crack Growth Curves for the TDP (Rainflow


Counting vs. Simple Range Counting)
The fatigue crack growth lives calculated with Rainflow counted
histograms and Simple Range counted histograms for both locations are
given in Table 12 in years and are given in Table 13 in number of
consecutive storms that could be endured.
Table 12 Fatigue Crack Growth Life in Years Calculated with
Rainflow Counting and Simple Range Counting Methods
Fatigue Crack Growth Life (Years)
Percentage
Cycle Counting Method
Difference
Location
Rainflow
Simple Range
(%)
First Weld
0.296
0.600
102.7
TDP
0.186
0.250
34.4

Table 13 Fatigue Crack Growth Life in No. of Consecutive Storms,


Calculated with Rainflow and Simple Range Counted Histograms
Fatigue Crack Growth Life (No. of
Percentage
Consecutive Storms Endured)
Difference
Cycle Counting Method
Location
(%)
Rainflow
Simple Range
First Weld
864
1752
102.7
TDP
543
730
34.4

CONCLUSION
Load sequence effects sometimes play an important role in fatigue
estimation in variable amplitude loading. Before conducting the fatigue
assessment for the riser system it is recommended to screen the load
histories either from the time-domain simulation or from the real-time
monitoring data to check if one-sided overloads are expected to
produce sequence effects.
Different cycle counting methods can result in very different load
histograms for fatigue assessment. Caution should be taken to select the
appropriate cycle counting method. Rainflow counting method is the
most widely used cycle count method and is recommended to be used,
especially when no experimental testing data are available to justify the
material memory has been lost during the loading history.
The S-N approach is most widely used in fatigue assessment of riser
system. When local stress concentration is expected (for example in
mechanical connectors), the strain-life approach is recommended. Both
S-N and strain-life approaches are used to predict life to crack
initiation. However when crack growth life is a significant part of the
total fatigue life, fatigue crack growth approach should be used.
In all three fatigue assessment approaches: S-N, strain-life, and fatigue
crack growth, Simple Range counted histogram results in 30% to 200%
higher lives compared to Rainflow counted histograms. Using Simple
Range counting could result in non-conservative riser design.

REFERENCES
Palmgren A.G.: Die Lebensdauer von Kugellagern (Life Length of Roller
Bearings. In German). Zeitschrift des Vereines Deutscher Ingenieure
(VDI Zeitschrift), ISSN 0341-7258, Vol 68, No 14, April 1924, pp
339341.
Miner A.M.: Cumulative Damage in Fatigue, J. of Appl. Mech. Trans
ASME, 12, 159-164, 1945.
ASTM Standard E 1049 - 85 (2005). Standard Practices for Cycle
Counting in Fatigue Analysis.
Matsuishi, M. and T. Endo, "Fatigue of Metals Subjected to Varying
Stress," presented at Japan. Soc. of Mech. Engrg., Fukuoka, Japan,
March 1968.
BS5400: Part 10, 1980, "Cycle counting by the reservoir method,"
Appendix B, pg. 9/22.
Downing, S.D. and D. F. Socie, "Simple rainflow algorithm," Int. J. of
Fatigue, V4, N1, 31-40, 1982.
Conle, A., T.R.Oxland, T.H.Topper, "Computer-Based Prediction of
Cyclic Deformation and Fatigue Behavior," Low Cycle Fatigue ASTM
STP 942, 1988, pp.1218-1236
OTO 199 058, Fatigue Design Curves for Welded Joints in Air &
Seawater Under Variable Amplitude Loading. HSE, July 2000.
Smith K.N., Watson P. and Topper T.H., A Stress-Strain Function for
the Fatigue of Metals, ASTM J. of Materials, V5 N4 Dec. 1970.
SAE Fatigue Design Handbook, 3rd Edition, Editor R.C. Rice, Society

of Automobile Engineers, ISBN 1-56091-917-5, 1997, pp. 141-142.


Crack Propagation Simulation with Material Deformation Memory
Compensation.
http://fde.uwaterloo.ca/Fde/Crackgrowth/pdpropPaper2013.pdf
Strain-Stress-Life Data File for Tests of HSLA SAE980X Steel:
http://fde.uwaterloo.ca/Fde/Materials/Steel/Hsla/mergedSAE980.html
Fitted Strain-Stress-Life Curve for Damage Calculation of SAE980X
http://fde.uwaterloo.ca/Fde/Materials/Steel/Hsla/mergedSAE980_fitted.
html
BS7910: Guide to Methods for Assessing the Acceptability of Flaws in
Metallic Structures, September 2007.
Crack Propagation Simulation with Material Deformation Memory
Compensation
http://fde.uwaterloo.ca/Fde/Crackgrowth/pdpropPaper2013.pdf
DNV CN 30.2, Fatigue Strength Analysis for Mobile Offshore Units,
1984.
DNV RP 203, Fatigue Design of Offshore Steel Structure, 2011.
Yang L. and Fatemi A., Cumulative Fatigue Damage Mechanisms and
Quantifying Parameters: A Literature Review,, J. of Testing and
Evaluation, Vol.26, 1998.
Yin F., Fatemi A., and Bonnen J., Variable Amplitude Fatigue Behavior
and Life Predictions of Case-Hardened Steels, International J. of
Fatigue 32, 2010, pp. 1126 1135.

S-ar putea să vă placă și