You are on page 1of 49

Platform, Pipeline and Subsea Technology

Fatigue Design

Platform, Pipeline and Subsea Technology– Fatigue Design


1
1 INTRODUCTION ....................................................................................................... 3

2 STRUCTURAL FATIGUE......................................................................................... 3
2.1 THE NATURE OF FATIGUE LOADING ........................................................................ 3
2.2 FATIGUE BEHAVIOUR OF STRUCTURAL DETAILS ..................................................... 8
2.3 FATIGUE OF WELDED STRUCTURES ....................................................................... 12
2.3.1 Stress Concentrations..................................................................................... 12
2.3.2 Weld Defects................................................................................................... 15
2.3.3 Residual Stress................................................................................................ 16
2.4 FURTHER FACTORS AFFECTING WELD FATIGUE .................................................... 19
2.4.1 Effect of Steel Strength ................................................................................... 19
2.4.2 Effect of Plate Thickness ................................................................................ 20
2.4.3 Effect of Environment ..................................................................................... 20
2.4.4 Effect of Welding Method ............................................................................... 20
2.5 PREDICTING FATIGUE LIFE UNDER VARIABLE AMPLITUDE LOADING.................... 21
2.6 FATIGUE OF TUBULAR JOINTS ................................................................................ 24
2.7 FATIGUE OF OTHER STRUCTURAL DETAILS ........................................................... 34
2.8 WELD FATIGUE LIFE IMPROVEMENT TECHNIQUES................................................. 34
3 FATIGUE ANALYSIS .............................................................................................. 36
3.1 DETERMINISTIC FATIGUE ANALYSIS ...................................................................... 36
3.2 SPECTRAL FATIGUE ANALYSIS............................................................................... 37
3.3 RELATIVE MERITS OF DETERMINISTIC AND SPECTRAL FATIGUE ANALYSIS .......... 40
4 RISK BASED FATIGUE ANALYSIS AND INSPECTION PLANNING ........... 43

5 REFERENCES .......................................................................................................... 47
5.1 BOOKS ................................................................................................................... 47
5.2 STANDARDS ........................................................................................................... 47
5.3 RESEARCH REPORTS AND PAPERS .......................................................................... 47

Platform, Pipeline and Subsea Technology– Fatigue Design


2
1 Introduction

This section of the unit describes the phenomenon of structural fatigue, and discusses the
approaches used to design and maintain structures against fatigue failure. It is organized
into three sections: the first provides a general background to structural fatigue, the second
discusses methods of fatigue analysis used in the design of offshore structures, and the
third provides an introduction to risk-based approaches and inspection philosophies.

2 Structural Fatigue

In the context of engineering, fatigue is the process by which a crack can form and then
grow under repeated or fluctuating loading.

The magnitude of the fluctuating loading required to induce fatigue cracking may be much
less than that required to cause failure under a single application of the load. In
particular, it may be much less than the load corresponding to the allowable static
design stress.

While some types of structure may be able to tolerate extensive cracking without
compromising their load-carrying capacity, other structures may be prone to sudden
collapse or excessive deflection if a crack is allowed to progress to some critical size.

The initiation and growth of fatigue cracks depends on the number of load cycles,
regardless of whether these cycles occur in quick succession, or there are significant
periods of time between them. Fatigue performance is therefore often expressed as a life
(or period of time before failure) under a particular loading regime. An important
corollary of this is that fatigue is a cumulative process.

2.1 The Nature of Fatigue Loading

The definitions commonly used to describing individual fatigue loading cycles are shown
in Figure 2-1. Note carefully the definition of stress range which is important in the
fatigue analysis of welded structures.

Another parameter which is sometimes used to describe fatigue loading cycles is the
Stress Ratio, which is defined commonly as the lower limit stress/upper limit stress, or
simply as:

Minimum Stress
Stress Ratio  (1)
Maximum Stress

Figure 2-1 shows constant amplitude loading, however, in practice the majority of
engineering structures are subjected to some type of variable amplitude, or random
loading. Some general examples are shown in Figure 2-2.

Platform, Pipeline and Subsea Technology– Fatigue Design


3
Figure 2-1: Definition of parameters describing constant amplitude fatigue
loading cycles.

Figure 2-2: Some general examples of random loading on engineering


structures.

Platform, Pipeline and Subsea Technology– Fatigue Design


4
Some examples of loads which contribute to fatigue damage may be:

 Loads induced during fabrication or construction.


 Loads induced during transportation.
 Loads induced during installation (e.g. pile driving)
 In-place loads induced by waves, current and wind.
 Pressure variations (for pipelines and pressure vessels).
 Temperature variations.
 Weight variations (or live loads).
 Vortex induced vibrations.
 Machinery induced vibrations.

In the case of fixed and floating offshore platforms it is usually the in-place environmental
loads due to waves which contribute mostly to fatigue damage.

Figure 2-3 shows the stress variation in the joint of an offshore platform due to wave
loading. Unsurprisingly the stress variation follows the water surface elevation, i.e. the
stress fluctuations are associated directly with waves.

Figure 2-3: Example of stress variation in a joint of a fixed offshore platform.

Figure 2-2 and Figure 2-3 show examples of random fatigue loading described in the time
domain. It is also possible to represent a time series of water surface elevations (and
associated fatigue loads) as the sum of a number of sinusoidal wave components, by a
Fourier series:

N
x(t )   ai cos( i t   i ) (2)
i 1

Where ai is the amplitude and i is the frequency of the ith wave component. i represents
a phase angle

Platform, Pipeline and Subsea Technology– Fatigue Design


5
This concept is shown in Figure 2-4. This permits the random loads to be described in the
frequency domain by an energy spectrum, S() which is defined as:

1
ai2
S(ω i )  2
(2)
ω

where i is a discrete frequency, and ai is the amplitude of the cycles in the frequency
band . These definitions are further illustrated in Figure 2-5. This definition of a load
spectrum can be useful in spectral fatigue analysis, which will be discussed in a later
section of these notes.

Figure 2-4: The relationship between time domain and frequency domain
representation of ocean waves.

Figure 2-5: A random load x(t) and its energy spectrum, S().

Platform, Pipeline and Subsea Technology– Fatigue Design


6
Spectra are sometimes described as either broad-banded or narrow-banded and these
concepts are shown in Figure 2-6. In a broad-banded process, a number of smaller cycles
are superimposed on larger cycles, and these smaller cycles may not pass through the
mean level. By contrast, in a narrow-banded process most cycles pass through the mean
value.

Figure 2-6: Definition of narrow-band and broad-band load spectra.

A further important point is that fatigue loads reflect the dynamic behaviour of the
structure. If a structure is excited at close to its natural frequency, dynamic amplification
of the loads must be accounted for. A simple way that is sometimes used to correct the
results of a quasi-static structural analysis to account for dynamics is to assume that the
structure behaves as a single degree of freedom system. A Dynamic Amplification Factor
(DAF) is simply applied to the stresses calculated in the quasi-static analysis. The DAF is
a function of the system damping and the closeness of the frequency of the excitation
force and the natural frequency of the structure. It may be calculated according to the
following equation, and the relationship between the DAF and damping, and the excitation
frequency is shown in Figure 2-7.

Figure 2-7: Relationship between Dynamic Amplification Factor (DAF), damping


ratio() and /N.

Platform, Pipeline and Subsea Technology– Fatigue Design


7
1
DAF  (3)
1     2  ζ  
2 2 2

where   ω ωN

and  is the damping ratio,


 is the frequency of the excitation force,
N is the natural frequency of the structure (assuming a single degree of freedom
system).

For sway of fixed offshore platforms, a commonly used damping ratio in fatigue analysis
is  = 0.02.

For dynamically sensitive structures, or sections of structures, a more rigorous approach is


often required. An excellent reference is Barltrop and Adams (1991).

2.2 Fatigue Behaviour of Structural Details

In describing the fatigue behaviour of engineering structures, the focus is frequently on


structural details. This is because fatigue cracking is highly localized. Therefore, in
engineering design the fatigue behaviour of each detail must be considered under the
action of local loads.

Fortunately, there is now a considerable body of test data covering a wide range of
structural details. Figure 2-8 shows a butt welded specimen being tested under constant
amplitude loading in a hydraulically powered fatigue test machine. Data from these types
of test form the basis for fatigue analysis. The data is usually presented as an S-N curve,
which plots Stress Range on the ordinate (y axis) and Number of Cycles to Failure (on a
logarithmic scale) on the abscissa (x axis). An example is given in Figure 2-9.

It is worthwhile noting the wide range of scatter in the test results (an order of magnitude
in cycles to failure in some cases!). Therefore, Design S-N curves usually represent the
mean line minus 2 standard deviations to obtain a suitably conservative estimate of the
number of cycles to fatigue failure.

Some S-N curves include an endurance limit, or stress level below which fatigue failure
will not occur irrespective of the number of loading cycles applied to the specimen (see
Figure 2-10). To a large extent this is a consequence of the method of fatigue testing. The
correct method of dealing with small amplitude load cycles in the fatigue analysis of
welded structures remains the subject of ongoing research. However, there is increasing
evidence that endurance limits should be ignored, particularly for structures
subjected to variable amplitude loading in corrosive environments. However, some
engineering guidelines make a compromise by suggesting S-N curves with different slope
in the low stress region. Some examples of these are provided later in these notes.

Platform, Pipeline and Subsea Technology– Fatigue Design


8
Figure 2-8: Constant amplitude fatigue test of a butt welded specimen.

Platform, Pipeline and Subsea Technology– Fatigue Design


9
Figure 2-9: Fatigue test data for butt welded specimens.

Platform, Pipeline and Subsea Technology– Fatigue Design


10
Stress

Endurance Limit

S0

Number of Cycles

Figure 2-10: The endurance limit represents the stress level below which
fatigue failure will not occur in a constant amplitude fatigue test.

One important shortcoming in describing fatigue behaviour using an S-N curve is that the
point of “failure” is not always clearly defined. For example, it may be the point at which
a detectable crack initiates, the point at which a crack proceeds through the full thickness
of a plate, or the point at which a structural specimen breaks in half.

A more physically correct approach is to describe the propagation of a crack under cyclic
loading using fracture mechanics techniques. Unfortunately, a discussion of fracture
mechanics techniques is beyond the scope of this course, however, there are numerous
texts on this subject (a good introduction is the book by Broek (1986)). Furthermore,
British Standard 7910 provides an excellent engineering guidance for assessing flaws and
cracks in engineering structures using fracture mechanics techniques (also known as
Engineering Critical Analysis).

Platform, Pipeline and Subsea Technology– Fatigue Design


11
2.3 Fatigue of Welded Structures

Welded joints are particularly susceptible to fatigue failure. Figure 2-11 compares the
fatigue performance of welded specimens with plain steel specimens and specimens
containing holes.

Figure 2-11: S-N Curves comparing the fatigue behaviour of welded specimens
with plane plate and notched specimens.

There are three major factors which contribute to the susceptibility of welded joints to
fatigue failure:

1. As joints, welds are subjected to both the stress concentration caused by their
location at structural discontinuities, and the stress raising effect of the weld shape
itself.

2. Due to the nature of the welding process they are likely to contain defects which
act as fatigue crack initiators.

3. High tensile residual stresses frequently exist in the vicinity of the weld as a result
of shrinkage during solidification and cooling of the weld metal.

These reasons are all important, and influence the way in which engineers carry out
fatigue analysis of welded structures, and ultimately design against fatigue failure. They
are therefore discussed in more detail in the following sections.

2.3.1 Stress Concentrations

There are two important sources of stress concentration which contribute to the
susceptibility of welded joints to fatigue cracking.

Platform, Pipeline and Subsea Technology– Fatigue Design


12
The first source of stress concentration is due to the location of welds at structural
discontinuities where there are changes in geometry and stiffness. This leads to a local
increase in stress at locations within the joint, and these are sometimes referred to as “hot
spots”. Misalignment of joints can also lead to secondary bending moments which can
increase stresses locally.

The concept of hot spot stress concentration is illustrated in Figure 2-12.

Figure 2-12: Example of hot spot stresses in a tubular nodal joint.

Platform, Pipeline and Subsea Technology– Fatigue Design


13
The second source of stress concentration is the shape of the weld itself. A notch is
formed by the toe of the weld and this location is by far the most common initiation site
for fatigue cracking in welded structures. However, the root of partial penetration welds
and defects introduced during the welding process also represent notches, and therefore
potential fatigue initiation sites. Some examples of fatigue crack initiation sites in welds
are shown in Figure 2-13 while the influence of the toe radius and flank angle on the
fatigue performance of butt welds is shown in Figure 2-14.

(a) (b)

(c)

Figure 2-13: Fatigue crack initiation sites: (a) at the toe of a butt weld, (b) at the
toe of a butt weld containing porosity, (c) at a lack of penetration defect.

Percent reduction
in fatigue strength

Figure 2-14: Influence of flank angle and toe radius on the reduction in fatigue
strength of butt welded joints.

Platform, Pipeline and Subsea Technology– Fatigue Design


14
The influence of weld shape and defects on fatigue behaviour emphasizes the importance
of weld quality in achieving adequate fatigue performance of welded joints. Particular
care should be taken with weld quality when joints are subjected to fatigue loading.

Another important point is that S-N curves for welded joints are based on experimental
data from fatigue tests on welded structural details. Therefore the stress concentration due
to the weld itself is included implicitly in the S-N curve and does not need to be
considered separately. The stress range on the ordinate (y axis) of the S-N curves of
welded details refers to the local, or hot spot stress. This stress must be calculated as part
of any fatigue analysis.

2.3.2 Weld Defects

The complexity of the welding process often leads to defects occurring in, or adjacent to,
the weld metal. These may be macroscopic defects (such as lack of fusion between weld
metal and parent plate, cracking, inclusions, or porosity) which are commonly detected
using conventional non-destructive examination techniques. It is common practice to
examine critical welds in structures to ensure that any defects present are below an
acceptable level.

However, even in “sound” welds a range of microscopic defects may occur. These are
commonly encountered at the weld toe where melted weld metal meets unmelted parent
metal and surface oxides. The rapid cooling of this zone promotes material
heterogeneities and the formation of a range of microscopic defects which may include
porosity, slag inclusions and sharp undercuts. These defects coincide with the stress
concentration of the weld and may ultimately become the initiation sites of fatigue cracks.
The nature of these defects is shown diagrammatically in Figure 2-15.

Intrusions at weld toe


approx 0.1-0.15 mm
deep

Figure 2-15: Microscopic intrusions at the toe of a “sound” weld.

Platform, Pipeline and Subsea Technology– Fatigue Design


15
2.3.3 Residual Stress

The contraction of weld metal during solidification and cooling is responsible for the
introduction of “locked-in” residual stresses in welded structures. These stresses exist
independent of external loading. Therefore they will be balanced within the structure; in
other words there is a system of tensile and compressive components of stress which is in
equilibrium.

Two systems of residual stress may be produced in a welded structure: global reaction
stresses which affect members as a whole, and localized residual stresses in the vicinity of
joints.

The concept of global reaction stresses is shown in Figure 2-16. Here the assembly
procedure may introduce an overall distribution of stresses within the structure. In the
simplest cases, tension in some members will be balanced by compression in others.

Tensile load in brace


balanced by compression
load in adjacent
members

Girth weld
completed after
adjacent welds

Figure 2-16: Concept of global reaction stresses.

The concept of localized residual stresses is shown in Figure 2-17 and Figure 2-18. Most
welded joints have sufficient external restraint to lead to tensile residual stresses of similar
magnitude to the material yield stress in the vicinity of the joint.

Platform, Pipeline and Subsea Technology– Fatigue Design


16
Figure 2-17: Formation of residual stress as a result of welding: (a) expected
longitudinal shrinkage of “unrestrained” weld; (b) longitudinal shrinkage of
restrained weld.

Figure 2-18: Typical residual stress distribution in a welded joint.

From the viewpoint of the analysis of welded structures the most important residual
stresses are the localized stresses in the joint. To understand the implications of high local

Platform, Pipeline and Subsea Technology– Fatigue Design


17
tensile residual stress on fatigue, consider Figure 2-19 which shows an alternating fatigue
load applied to a welded joint. It is well established that, all other things being equal,
compressive stress does not contribute to fatigue crack initiation or propagation.
Therefore only the tensile part of the loading cycle contributes to fatigue failure.
However, if a tensile residual stress of yield stress magnitude is present at the weld, this
combines with the fatigue load to give the stress cycle shown in Figure 2-19(b). Here an
elastic-perfectly plastic material behaviour has been assumed so that the stress cycle varies
from the yield stress downwards. In this instance the entire stress range is tensile and
therefore damaging from the viewpoint of fatigue.

Figure 2-19: When the residual stress is equivalent to a tensile yield stress the
actual stress range will vary from yield stress downwards, regardless of the
nominal stress ratio.

Therefore, the significance of welding residual stress is that even compressive loads
applied to a structure may lead to a net tensile fatigue stress in the vicinity of the welds.
For this reason it is the stress range which is usually most important in determining
the fatigue behaviour of welded joints.

Localized residual stresses can be relieved to some extent by Post Weld Heat Treatment
(PWHT). This involves heating the joint in an oven to approximately 600C for a period
of some hours. The yield stress of the steel is reduced by the heating and this allows
relaxation of the locked-in stresses. Some fatigue analysis guidelines make allowance for
a limited benefit from PWHT. If in doubt, a conservative approach is to ignore any
beneficial effects which may arise from PWHT.

Platform, Pipeline and Subsea Technology– Fatigue Design


18
2.4 Further Factors Affecting Weld Fatigue

There are a number of other factors which have some influence on the fatigue behaviour
of welded joints and are usually taken into account during fatigue analysis. These are
described in this section.

2.4.1 Effect of Steel Strength

One of the most useful features of steel is that its mechanical properties can be
significantly altered through alloying and heat treatment. Up to a certain limit, the fatigue
strength of smooth steel specimens increases proportionally with tensile strength.
However, this relationship does not hold for welded joints. Figure 2-20 shows fatigue test
results for welded specimens made from steels ranging in tensile strength from 438 to 753
MPa. The test results fall within a relatively narrow scatter band and there is no
correlation between ultimate tensile strength and fatigue strength.

Therefore, if an engineering design is limited by its fatigue performance, using a


higher strength steel will not improve the situation.

Figure 2-20: Effect of steel ultimate tensile strength on fatigue behaviour (note:
1 tons/in2 = 15.44 MPa).

Platform, Pipeline and Subsea Technology– Fatigue Design


19
2.4.2 Effect of Plate Thickness

It has generally been observed that, all other things being equal, increasing plate thickness
results in a decrease in fatigue performance. Figure 2-21 shows the results of fatigue tests
on welded specimens in bending. Although the reason for the thickness effect is not
universally agreed, one suggestion has been that thinner plates have a higher stress
gradient, and therefore the driving force behind fatigue crack growth decreases more
rapidly than in thick plates. Most design codes and standards which deal with fatigue
design include a correction for the thickness effect.

Figure 2-21: Fatigue test results showing the influence of plate thickness.

2.4.3 Effect of Environment

A corrosive environment, such as seawater, may have the effect of accelerating the growth
of fatigue cracks, and therefore reducing overall fatigue performance. The presence of a
corrosive environment also effectively removes the endurance limit on the S-N curve.

Cathodic protection reduces the impact of this detrimental effect to some extent.

In the case of offshore structures different S-N curves are usually specified depending on
the prevailing corrosion conditions. It is important to note at this point that cathodic
protection is ineffective for joints in the splash zone where electrolyte presence is not
continuous. These joints exist under effectively free corrosion conditions.

2.4.4 Effect of Welding Method

There is very little evidence to suggest that, among common welding techniques, different
welding methods produce intrinsically different fatigue strengths in welded joints. If

Platform, Pipeline and Subsea Technology– Fatigue Design


20
anything, manual welding techniques produce welds with marginally better fatigue
performance compared with automatic welds.

2.5 Predicting Fatigue Life Under Variable Amplitude Loading

It has already been discussed how S-N curves are generated from laboratory tests carried
out under constant amplitude fatigue conditions. However, it has also been pointed out
that most engineering structures are subject to random, variable amplitude loading. It is
therefore necessary to have a method of estimating the fatigue life under a variable
amplitude loading regime, but still use S-N curves generated under constant amplitude
loading.

The method that has achieved the broadest acceptance is the Palmgren-Miner linear
cumulative damage hypothesis, commonly known as Miner’s Rule. This states that

ni n1 n2 n3
N   
N1 N 3 N 3
 ... D (4)
i

where n1, n2,… are the number of cycles that stresses 1, 2,… are applied to the joint, and
N1, N2,… are the corresponding numbers of cycles to failure of a similar joint under
constant amplitude loading at those stresses. D is a constant that represents the
accumulated fatigue “damage” of the joint. If D = 1.0, the rule may be restated that,
under variable amplitude loading, the basic damage fraction caused by each separate
loading cycle is equal to that caused by a single cycle of the corresponding constant
amplitude loading, and that failure occurs when the sum of these basic damage fractions
reaches unity.

This concept is shown diagrammatically in Figure 2-22.

Figure 2-22: Definition of Miner’s rule.

Platform, Pipeline and Subsea Technology– Fatigue Design


21
Figure 2-23 shows a comparison of Miner’s Rule with typical experimental fatigue test
results under variable amplitude loading. It can be seen that a Miner’s summation (or
damage accumulation, D) =1.0 is the most likely result corresponding to failure, however,
there is considerable variation in the results.

0.35
Frequency of Occurrence

0.3

0.25

0.2
Corresponding lognormal distribution
0.15

0.1

0.05

0
0 1 2 3 4 5 6 7 8 9
Miner's Sum

Figure 2-23: Comparison of calculated Miner’s summation and variable


amplitude test results.

There may be instances where a method of turning a time series of stresses into a number
of discrete stress cycles is required. A complete discussion of the various methods
available is beyond the scope of this course, however it is worth mentioning that the
Rainflow, or Reservoir method of cycle counting has gained the broadest acceptance. A
discussion of this technique can be found in British Standard BS 7608:1993. In many
instances, the fatigue loads in offshore structures are already grouped in a manner which is
conducive to fatigue analysis using Miner’s Rule. An example of this is the wave
occurrence table shown in Figure 2-24.

An alternative is the wave height exceedance diagram of Figure 2-25 which shows Wave
height on the ordinate (y axis), and the number of waves exceeding a particular wave
height on the abscissa (x axis).

An important corollary of the Miner’s Rule calculation is that a large number of small
stress cycles can be as equally damaging as a few large cycles. It is therefore important to
be able to carry out the Miner’s Rule calculation in order to determine just how damaging
a given loading spectrum really is.

Platform, Pipeline and Subsea Technology– Fatigue Design


22
Figure 2-24: Wave occurrence table for fatigue analysis (25 year period).

Platform, Pipeline and Subsea Technology– Fatigue Design


23
16.0
North West Shelf (typical)
14.0 Central North Sea (typical)
Gulf of Mexico (design)
Significant Wave Height, H sig (m)

12.0

10.0
Central North Sea

8.0

6.0
North West Shelf
4.0 (shallow water)

2.0

0.0
1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07 1.E+08
Number of Waves Exceeding H sig per year

Figure 2-25: Wave height exceedance diagram.

2.6 Fatigue of Tubular Joints

The vast majority of fixed offshore platforms are constructed from a space frame of
tubular joints. Considerable research effort has therefore been directed toward
characterizing the fatigue performance of this particular class of welded joint.

Where tubulars are butt jointed to form members, any eccentricity between adjacent
tubulars can lead to the development of secondary bending moments in response to axial
loads in the member. For the purposes of fatigue analysis this may be accounted for by
assigning a Stress Concentration Factor (SCF) which, when multiplied by the nominal
axial stress in the member, describes the increased local stress in the vicinity of the welded
joint, i.e.

local = SCFnominal (5)

The various sources of geometric stress concentration which may be encountered in a


tubular butt weld are shown in Figure 2-26. For butt welded tubular members, the
following formula for eccentricities in flat plates provides a conservative estimate of the
SCF at the joints:

3( m )
SCF  1  (6)
t

Platform, Pipeline and Subsea Technology– Fatigue Design


24
A more sophisticated and less conservative method of evaluating the stress concentration
factor in butt jointed members can be found in DNV-Recommended Practice C203, along
with guidance for joints in pipelines.

Figure 2-26: Sources of geometric stress concentration in butt welded tubular


members.

Where tubular members such as jacket braces and legs intersect, a complex geometry is
created. In these cases the stress distribution around the joint is not uniform due to the
complexity of the localized geometry. Considerable effort has been spent investigating
stress concentration factors to be used with tubular joints such as those shown in Figure
2-27. The methods used have included strain gauging and load testing scale models,
photoelastic testing, and detailed finite element modelling. These activities have resulted
in the availability of parametric equations which are suitable for the analysis of a range of
commonly encountered tubular joints. A comprehensive discussion of this subject may be
found in HSE (1997).

Appendix A provides the parametric equations due to Efthymiou (1988), which are
probably now the most widely used of their type. The SCFs for T and Y joints are defined
only at the “crown” and “saddle” locations (refer back to Figure 2-12 for the definition of
these locations). Therefore there must be some method of interpolation so that the SCFs
at the intermediate points can be calculated. The conventional method is to linearly
interpolate the SCFs for axial loading, while the SCFs associated with in-plane and out-of-
plane loading are sinusoidally distributed. This concept is shown in Figure 2-28.

Platform, Pipeline and Subsea Technology– Fatigue Design


25
Figure 2-27: Geometric parameters describing welded tubular joints.

When carrying out the interpolation described above it is important to understand that the
SCFs calculated using parametric equations for tubular joints should be used only with the
nominal member stresses. In the case of axial SCFs, the nominal member axial stress
should be used in conjunction with the calculated SCF, while in the case of bending SCFs,
the maximum (outer fibre) bending stress should be used in conjunction with the relevant
bending SCF.

3
2 4

1 5

8 6
7

Figure 2-28: Recommended distribution factors for SCFs in T & Y joints.

Platform, Pipeline and Subsea Technology– Fatigue Design


26
Stresses due to axial, in-plane bending and out-of-plane bending loads may be added using
the principle of superposition. The following formula is useful:

HSS  SCFaxS  SCFaxC  SCFaxS   θ 90σ ax 

 
 SCFipC  sin θ σ ip  (7)


 SCFopS  cos θ σ op  

where HSS is the hot spot stress


SCFaxS is the axial SCF at the saddle point
SCFaxC is the axial SCF at the crown point
SCFipC is the in-plane bending SCF at the crown point
SCFopS is the out-of-plane bending SCF at the saddle point
ax is the nominal axial stress
 ip is the nominal (maximum) in-plane bending stress
 op is the nominal (maximum) out-of-plane bending stress
 is the angle around the joint measured from the saddle (see diagram below).

Crown

90


Saddle 0

Figure 2-29: Definition of  in hot spot stress calculation equation.

Note that this method of interpolating the axial SCFs only works in the first quadrant (i.e.
for  between 0 and 90 degrees) Symmetry should be used to determine the interpolated
axial SCFs in the other quadrants.

The fatigue behaviour of tubular joints has been characterized by numerous fatigue tests,
and typical results are shown in Figure 2-30. From tests such as these, suitably

Platform, Pipeline and Subsea Technology– Fatigue Design


27
conservative design S-N curves can be derived. As an example, Figures 231 to 233
show the design S-N curves for tubular joints published in the UK Health and Safety
Executive’s (HSE) guidance on the design, construction and certification of offshore
installations. Different curves are shown for fatigue of joints in air, under free corrosion
conditions in seawater, and in seawater with cathodic protection. These S-N curves may
also be expressed mathematically as:

log10 ( N )  log10 ( K1 )  m log10 ( S B ) (8)

where N is the predicted number of cycles to failure under stress range SB, K1 is a constant
and m is the inverse slope of the S-N curve. Details of the basic design S-N curves are
given in Table 1 below. S0 and N0 represent the point where the slope of the S-N curve
changes, see Figure 2-34.

The HSE design S-N curves apply to all joints with plate thicknesses of 16 mm or less.
For thicker joints, a thickness correction applies to take account of the thickness effect on
welded joints discussed previously. The thickness correction can be incorporated into the
mathematical expression of the S-N curves in the following manner:

 SB 
log10 ( N )  log10 ( K1 )  m log10  
q 
(9)
 t B t  

where tB is the reference thickness, = 16 mm


t is the thickness of the member under consideration
q is the thickness exponent factor, = 0.3
and all other constants are defined as before.

Platform, Pipeline and Subsea Technology– Fatigue Design


28
Figure 2-30: Experimental results for 16mm thick tubular joints tested in air.

Table 1: Details of thee basic S-N curves in HSE Guidance.

Welded
Plates

Tubular
Joints

Platform, Pipeline and Subsea Technology– Fatigue Design


29
Figure 2-31: Basic HSE design S-N curves for welded tubular joints and plates
in air.

Platform, Pipeline and Subsea Technology– Fatigue Design


30
Figure 2-32: Basic HSE design S-N curves for welded tubular joints and plates
in seawater – free corrosion.

Platform, Pipeline and Subsea Technology– Fatigue Design


31
Figure 2-33: Basic HSE design S-N curves for welded tubular joints and plates
in seawater – cathodic protection.

Platform, Pipeline and Subsea Technology– Fatigue Design


32
Stress

Point where slope of


S-N curve changes

S0

N0 Number of Cycles

Figure 2-34: Definition of S0 and N0 on dual slope S-N curve.

It is important to note that different offshore design codes have subtle differences in the
elements of fatigue analysis. There are slight differences in S-N curves, thickness
corrections, and the required fatigue design life. Once a particular fatigue design code
is adopted, it should be followed consistently, and in its entirety.

A good example of the differences between design codes is the required fatigue design
life. The API offshore design code, RP2A, requires that the calculated fatigue life be at
least twice the design life of the structure, whereas the HSE guidance requires that the
calculate fatigue life be at least equal to the design life (however, a more conservative S-N
curve is used in the latter case). The Norwegian offshore structural design code (Norsok
Standard N-004, 1998) suggests an even more sophisticated approach where the calculated
design lives also depend on the accessibility of the joint for inspection and repair, and this
is described in Table 2. Note that a design fatigue factor of 10 means that the calculated
fatigue life must be at least 10 times the design life of the structure.

Table 2: Design fatigue factors (from Norsok Standard N-004, 1998).

Access for Inspection and Repair


Classification of structural
components based on No access, or in the Accessible
damage consequence splash zone
Below Splash Zone Above Splash Zone

Substantial Consequences 10 3 2
Without Substantial
3 2 1
Consequences

Platform, Pipeline and Subsea Technology– Fatigue Design


33
2.7 Fatigue of Other Structural Details

In addition to tubular joints there are, of course, a wide range of other types of welded
structural details which occur in offshore structures. These include attachments such as
ladders and boat landings, standard weld details in the topsides structure, and ship details
in FPSOs and other floating production systems. Where these joints are subjected to
fluctuating loading, their fatigue performance must also be assessed.

Fortunately, there are a range of design standards and guidelines which describe the S-N
curves and analysis methodologies for a range of commonly encountered structural details.
These guidelines include:

 British Standard 7608:1993 and DNV Recommended Practice RP-C203 (2001) which
provide guidance for the analysis of a range of common structural details, some of
which might typically be found in platform topside structures.

 DNV Class Note 30.7 (1998) which provides guidance on stress concentration factors
and design S-N curves for a range of ship structural details which might typically be
found in FPSOs and other offshore structures made up from stiffened panels.

2.8 Weld Fatigue Life Improvement Techniques

Because of the susceptibility of welded joints to fatigue failure, there have been numerous
techniques developed to improve their behaviour. Gurney (1979) provides a
comprehensive coverage of the different methods which have been tested.

However, in practice, the most widely used methods for improving the fatigue behaviour
of welded joints are toe grinding and peening.

Toe grinding involves using either a disc, or rotary burr grinder to reduce the stress
concentration at the toe of the weld as shown in Figure 2-35. To be most effective, toe
grinding should extend 0.5–1.0 mm into the parent plate to remove any small undercuts or
intrusions which may act as fatigue crack initiators.

Peening involves impacting the surface of the weld with a pneumatic hammer fitted with a
rounded tool. This induces a compressive residual stress which retards fatigue crack
initiation. Peening is usually directed towards the toe of the weld and this may have the
added benefit of also improving the shape of the weld toe.

The relative benefits of toe grinding and peening are shown in Figure 2-36. Despite the
significant benefits which can be achieved through these techniques, they are time
consuming and costly, and skill is required to achieve consistently good results. It is for
these reasons that these techniques are rarely considered in the design of welded joints
(that is, joints are usually designed so that improvement techniques are not relied upon to
achieve satisfactory performance). However, they are sometimes used as an added
defence against fatigue failure in critical welds.

Platform, Pipeline and Subsea Technology– Fatigue Design


34
Methods of improving the fatigue performance of a joint at the design stage include
modifying the joint geometry to reduce local stresses, or replacing the welded tubular joint
with a cast node. Using a casting is an expensive option, but may be cost effective for
particularly complex joints.

0.5 mm min

Figure 2-35: It is recommended that toe grinding extend below the plate surface
to remove weld defects.

Figure 2-36: Comparison of grinding, peening and some other weld


improvement techniques on the fatigue performance of fillet welded specimens.

Platform, Pipeline and Subsea Technology– Fatigue Design


35
3 Fatigue Analysis

There are two common approaches in the fatigue analysis of offshore structures. Perhaps
the most popular is analysis in the time domain, commonly known as deterministic
fatigue analysis. However, as will be discussed in a later section, there can be some
advantages carrying out fatigue analysis in the frequency domain, and this type of analysis
is called a stochastic, or spectral fatigue analysis.

The deterministic and spectral approaches to fatigue analysis are described in the
following sections, and the relative advantages of each are compared.

3.1 Deterministic Fatigue Analysis

Deterministic fatigue analysis requires fatigue loads to be expressed as a finite number of


discrete events. An example of this was provided in Figure 2-24 where a discrete numbers
of waves of varying height and period are provided over a 25 year period.

The subsequent deterministic fatigue analysis is best summarised as a series of steps:

1. All the physical phenomena which are likely to contribute to fatigue of a structure
over its entire life need to be identified. In the case of offshore structures these may
include loads imposed during construction, transportation, live loads due to machinery,
and, perhaps most importantly, in-place environmental loads due to waves, current and
wind. It must be recognized that the stresses at any potential fatigue failure site in the
structure will be different for waves coming from different directions and this must be
accounted for in the analysis.
2. The physical phenomena need to be translated into loads on structural members. The
use of specialized offshore structural computer analysis packages can often be used for
this step. The effect of dynamics should also be accounted for.
3. Loads in structural members need to be translated into localized joint stresses.
Member loads must firstly be translated into member stresses and, in the case of
tubular joints, hot spot stress concentration factors can be used to determine the
localized stresses at various points around the joint. Typically 8 locations around the
joint on both the chord and brace would be considered. Remember that it is the stress
range which is important in determining fatigue behaviour, and so the maximum and
minimum hot spot stresses need to be determined.
4. A S-N curve describing the relevant structural detail must be chosen. Appropriate
codes and standards provide a range of S-N curves.
5. A Fatigue Damage calculation must be carried out using Miner’s Rule. The
calculation must be carried out for expected loads over the design lifetime of the
structure. Remember that Miner’s Rule is cumulative, therefore the fatigue damage
due to one set of loads can be added directly to the fatigue damage caused by other
loads. In particular the fatigue damage may be summed for each wave direction.

Platform, Pipeline and Subsea Technology– Fatigue Design


36
6. The fatigue life of each detail (or hot spot) of interest can be compared with the design
life after appropriate factors of safety have been accounted for.

3.2 Spectral Fatigue Analysis

Spectral fatigue analysis recognizes the underlying random nature of the wave height
occurrence table (as given in Figure 2-24) and models the process statistically.

It has already been discussed in S ection 2.1 of these notes that a random series (of for
example wave heights) may be expressed as an energy spectrum. Figure 2-5 is
reproduced below:

The energy of a harmonic wave is proportional to the square of its amplitude, and the
energy in each frequency band is given by the following equation:

1
ai2
S (ω i )  2
(10)
ω

Using Fourier analysis techniques, a random process such as a wave history may be
represented by a superposition of a large number of sinusoidal components. If one
component of the excitation process is given by

x(t )  ai  cos(ω i t   i ) (11)

then the component of the response at the same frequency is given by

y (t )  Tω  x(t ) (12)

Platform, Pipeline and Subsea Technology– Fatigue Design


37
This provides the definition of the transfer function, T(). The excitation process is
typically the wave height spectrum. The response function is typically the hot spot stress
range at a particular joint location.

The energy spectrum of the response may therefore be related to the energy spectrum of
the excitation by using the relationship

S y (ω)  Tω  S x (ω)


2
(13)

The relationship between the excitation spectrum, the response spectrum and the transfer
function is illustrated in Figure 3-1.

Figure 3-1: The transfer function T(f) relates the excitation spectrum Sx(f) and
the response spectrum Sy(f).

The excitation and response functions, as well as the transfer functions may equally be
expressed as functions of frequency in Hz, instead of radians/sec. The definition of the
transfer function would then be given by:

S y ( f )  T f   S x ( f )
2
(14)

where f represents frequency expressed in Hz. The remainder of this section of the notes
will assume that the functions are expressed in Hz. In passing, it is important to note that
care should be taken to properly define the shape of the loading spectrum and transfer
function, particularly around the natural frequencies of the structure.

Once a fatigue stress spectrum has been developed (i.e. a plot of Sy( f )) then the fatigue
damage may be calculated by firstly calculating the zero and second order moments of the
stress spectrum.

The moments of an energy spectrum are defined as:

Platform, Pipeline and Subsea Technology– Fatigue Design


38

m n   S x  f   f n df (15)
0

Therefore the zero order moment is:


m 0   S x  f df (16)
0

and this represents the area under the spectral curve.

The second order moment is:


m 2   S x  f   f 2 df (17)
0

In practice, these integrals can be evaluated by numerical integration using for example
the trapezoid rule. This concept is shown in Figure 3-2.

Figure 3-2: Definition of zero order and second order moments of response
spectrum.

For a narrow-banded process, and where the frequency in the response spectrum is
expressed in Hz, the mean zero upcrossing period may be approximated as:

m0
Tz  (18)
m2

Platform, Pipeline and Subsea Technology– Fatigue Design


39
[ Note that this equation will be slightly different if the input, output and stress transfer
functions are expressed in rad/sec. rather than Hz. ]

The number of stress cycles, n, in a total time of T seconds is therefore given by:

T
n (19)
Tz

The complete derivation of spectral fatigue damage under a narrow banded process is
beyond the scope of this course. Interested readers are referred to Barltrop and Adams
(1990). However, assuming that the stress range within each short term seastate can be
described by the Rayleigh distribution, the fatigue damage for each seastate may be
calculated as:

m 2 (8m 0 ) m 2  2  m 
D T      (20)
m0 K  2 

where T is the time period in seconds


m0, m2 are the zero and second order spectral moments (defined above)
m and K are constants describing the S-N curve (see Section 2.6 of these
notes)

 is the incomplete gamma function: ( g )   x ( g 1) e x dx
0

The fatigue damage may be summed linearly over all seastates and all wave directions. If
T is the number of seconds in 1 year, then the fatigue life of the hot spot in years will be
1/D. The spectral fatigue analysis approach is illustrated schematically in Figure 3-3.

3.3 Relative Merits of Deterministic and Spectral Fatigue Analysis

An important aspect of spectral fatigue analysis is that, via the transfer function, it can
account for dynamic effects more completely than deterministic fatigue analysis.
Provided that appropriate software is available to develop the transfer function between
the excitation spectrum and the response spectrum spectral analysis is usually more
computationally efficient than deterministic analysis.

However, an important implicit assumption in the development of the transfer function is


that it applies to each wave component irrespective of its amplitude. This means that there
must be a linear relationship between the excitation (wave height) and the response (hot
spot stress range). In practice, this is a reasonable approximation for large jacket type
structures in moderate to deep water depth, although effort is sometimes put toward
modifying the transfer functions to account for the non-linear nature of the drag force in
Morison’s equation and the non-linear kinematics associated with higher order wave

Platform, Pipeline and Subsea Technology– Fatigue Design


40
theories. These effects may be particularly important for smaller structures in shallow
water, and spectral fatigue analysis may not be appropriate.

It is also important to recognize that deterministic and spectral fatigue analyses require the
wave height distribution to be expressed differently. Deterministic analyses require a
table describing the occurrence rate of waves of a specific height and period (such as that
shown in Figure 2-24). Spectral analyses, on the other hand require a table describing the
occurrence rate of individual seastate conditions (described typically by the significant
wave height and zero upcrossing period).

It should be apparent from the previous sections that both deterministic and spectral
fatigue analysis are computationally intensive. In practice, it would be commonplace to
carry out a damage summation for 8 locations on both the brace and chord side of all
critical tubular joints for a range of range of wave heights and periods over 8 directions for
a fixed offshore structure. For this reason, specialised offshore analysis packages such as
SESAM, SACS and StruCAD*3D all have fatigue analysis modules which carry out the
numerous repetitive calculations required.

Despite the convenience of analysis software, a sound understanding of the


fundamentals of fatigue remains an essential prerequisite for obtaining reliable
results.

Platform, Pipeline and Subsea Technology– Fatigue Design


41
Figure 3-3: Frequency domain spectral fatigue analysis.

Platform, Pipeline and Subsea Technology– Fatigue Design


42
4 Risk based fatigue Analysis and Inspection Planning

It is apparent from the previous discussion that there are several sources of uncertainty, or
variability in fatigue analysis. The practice adopted by most engineers when faced with
such uncertainty is to adopt conservative approaches (for example using the mean S-N
curve minus two standard deviations for design). While there is nothing wrong with this,
an alternative approach is to consider the uncertainties explicitly along with the inherent
conservatisms, and carry out a reliability analysis. The concept behind fatigue reliability
analysis is illustrated schematically in Figure 4-1. Here, Tmean represents the best estimate
of fatigue life, T represents the standard deviation (or variability) associated with the
estimate of fatigue life, and Ts represents the required service life. The shaded area
therefore represents the probability of fatigue failure.

Ts Tmean
Probability Density

T

Fatigue Life (Years)

Figure 4-1: Concept of fatigue reliability analysis: Tmean and T represent the
calculated distribution of fatigue lives, Ts represents the required service life,
and the shaded area represents the probability of fatigue failure.

A crucial aspect of any engineering reliability assessment is quantifying the uncertainties


associated with the analysis. In estimating the fatigue life of welded structures there are
three principal sources of uncertainty:

1. Uncertainty in estimating the fatigue loads and hot spot stresses on the structure.
2. Uncertainty associated with fatigue test data which is apparent as scatter on the S-N
curve (refer back to Figure 2-9).
3. Uncertainty associated with the exact value of Miner’s summation which will result in
fatigue failure of the joint (refer back to Figure 2-23).

Platform, Pipeline and Subsea Technology– Fatigue Design


43
A detailed description of the calculation of probability of fatigue failure taking into
account the uncertainties listed above is beyond the scope of this course, however,
Appendix B provides a general introduction. The calculation can be carried out using
either S-N data, or by using a fracture mechanics approach.

Although Figure 4-1 illustrates the concept of the probability of fatigue failure at the end
of a service lifetime, Ts, it is not difficult to imagine that the probability of fatigue failure
can be calculated at other times. This enables the probability of fatigue failure to be
plotted as a function of time, an example is shown in Figure 4-2. This figure shows how
the probability of fatigue failure increases with the amount of time that a joint is subjected
to a certain loading environment.

An extension of Figure 4-2 also demonstrates one of the important uses of probabilistic
fatigue analysis. The results of inspections can be used to update the probability of fatigue
failure using Bayesian statistics. Figure 4-3 illustrates schematically the effect of
inspection after 10 years with no crack being detected. In this way, inspection can be used
to maintain the probability of failure below some acceptable level. The results of the
analysis can also be used to determine optimum inspection intervals on a risk basis. This
is known as Risk Based Inspection (RBI) planning.

1.0E+00
Probability of Fatigue Failure, PF

1.0E-01

1.0E-02

1.0E-03

1.0E-04

1.0E-05

1.0E-06

1.0E-07
0 5 10 15 20
Service Life (years)

Figure 4-2: Typical results of a probabilistic fatigue analysis: Time in Service


vs. Probability of Fatigue Failure.

Platform, Pipeline and Subsea Technology– Fatigue Design


44
1.0E+00

Probability of Fatigue Failure, PF


1.0E-01 Effect of Inspection
at 10 years -
no crack detected
1.0E-02

1.0E-03
-4
P F = 2.2 x 10

1.0E-04

1.0E-05

1.0E-06

1.0E-07
0 5 10 15 20
Service Life (years)

Figure 4-3: The probability of fatigue failure can be updated through inspection.

Another use for probabilistic fatigue analysis is in design optimization studies. Here, the
total cost of a structure may be defined as:

Total Cost = CAPEX + OPEX + RISKEX (21)

where CAPEX is capital expenditure, including the cost of fabricating and installing a
structure or structural element.
OPEX is operating expenditure, including fatigue inspection.
RISKEX is risk related expenditure, defined as the probability of failure
multiplied by the cost of failure.

In many engineering designs there is a trade-off between these types of expenditure.


Figure 4-4 shows the results of probabilistic fatigue analysis on a joint in an offshore
structure. Increasing the thickness of the chord in the joint will increase CAPEX,
however, this will be offset by decreased OPEX (because the inspection intervals have
been increased) and decreased RISKEX (because the probability of failure has been
reduced).

Platform, Pipeline and Subsea Technology– Fatigue Design


45
1.0E-01

Minimum Joint
Probability of Fatigue Failure Design Thickness
1.0E-02

1.0E-03

Increasing Joint
1.0E-04 Thickness =
Increasing CAPEX

1.0E-05 Increasing Inspection Interval 41mm


= Decreasing OPEX 44mm
47mm
50mm
1.0E-06
0 5 10 15 20
Service Life (years)

Figure 4-4: Increasing joint thickness in an offshore structure may increase


CAPEX, but may also decrease OPEX and RISKEX.

Platform, Pipeline and Subsea Technology– Fatigue Design


46
5 References

5.1 Books

Almar-Naess A. (1985) Fatigue Handbook – Offshore Steel Structures, Tapir, Trondheim.


Barltrop N.D.P. and Adams A.J. (1991) Dynamics of Fixed Marine Structures, 3rd Ed.,
Butterworth-Heinemann, Oxford.
Broek D.B. (1986) Elementary Engineering Fracture Mechanics, 4th Ed., Kluwer,
Dordrecht.
Gurney T.R. (1979) Fatigue of Welded Structures, 2nd Ed., Cambridge University Press,
Cambridge.
Maddox S.J. (1991) Fatigue Strength of Welded Structures, 2nd Ed., Abington Publishing,
Cambridge.
UEG (1985) Design of Tubular Joints for Offshore Structures, UEG.

5.2 Standards

API (2000) Recommended Practice for Planning, Designing and Constructing Fixed
Offshore Platforms – Working Stress Design, 21st Ed., American Petroleum Institute.
BS 7608:1993 Fatigue Design and Assessment of Steel Structures, British Standards
Institute.
BS 7910:1999 Guide on Methods for Assessing the Acceptability of Flaws in Metallic
Structures, British Standards Institute.
DNV RP C203 (2001) Fatigue Strength Analysis of Offshore Steel Structures, Det Norske
Veritas.
DNV Classification Note 30.7 (1998) Fatigue Assessment of Ship Structures, Det Norske
Veritas.
HSE (1990) Offshore Installations: Guidance on Design, Construction and Certification,
4th Ed., United Kingdom Department of Energy / Health and Safety Executive.
Norsok (1998) Design of Steel Structures, Rev.1, Norsok Standard N-004, Norwegian
Petroleum Directorate.

5.3 Research Reports and Papers

Efthymiou (1988) Development of SCF Formulae and Generalized Influence Functions


for use in Fatigue Analysis, Offshore Tubular Joints Conference, Surrey UK.
HSE (1997) Stress Concentration Factors for Simple Tubular Joints, Offshore
Technology Report OTH 354, United Kingdom Health and Safety Executive.
HSE (1999) Background to New Fatigue Guidance for Steel Joints and Connections in
Offshore Structures, Offshore Technology Report OTH 92 390, United Kingdom Health
and Safety Executive.

Platform, Pipeline and Subsea Technology– Fatigue Design


47
APPENDIX A

Platform, Pipeline and Subsea Technology– Fatigue Design


48
APPENDIX B

Platform, Pipeline and Subsea Technology– Fatigue Design


49

You might also like