Sunteți pe pagina 1din 17

OTC-29534-MS

Strain-Life Fatigue Analysis for HPHT Equipment: Theory to Validation

Alvaro Aguilar, Parth D Pathak, James E Stevens, and Claire R Foley, OneSubsea, a Schlumberger Company

Copyright 2019, Offshore Technology Conference

This paper was prepared for presentation at the Offshore Technology Conference held in Houston, Texas, USA, 6 – 9 May 2019.

This paper was selected for presentation by an OTC program committee following review of information contained in an abstract submitted by the author(s). Contents of
the paper have not been reviewed by the Offshore Technology Conference and are subject to correction by the author(s). The material does not necessarily reflect any
position of the Offshore Technology Conference, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Offshore Technology Conference is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of OTC copyright.

Abstract
Subsea equipment covered by the API Spec 17 subcommittee has had limited focus on assessing fatigue life
because of external environmental loads using traditional analysis methods. With the current trend of high-
pressure, high-temperature (HPHT) development, the industry is migrating to an era of modern analysis
methods with complex material testing programs to assess potential fatigue life impacts due to such high-
pressure and -temperature exposures as well. This paper presents an approach and an example of a multiaxial
strain-life analysis method that meets the provided HPHT design guidelines of API Technical Report 17TR8.
The paper bridges the gap between theory and practicality in strain-life-based fatigue analysis and
presents a robust process developed for HPHT nickel alloy components, which are part of the subsea
20,000-psi vertical monobore subsea tree. The endeavor includes strategizing for required material tests
in environment, actual material testing, followed by material data processing, which includes statistical
corrections and extraction of parameters necessary for efficient fatigue analysis. The components are then
analyzed in finite-element analysis (FEA) with typical loading sequences as seen in its life of field. Finally,
the FEA results are postprocessed using the critical plane approach for all nodes in the model. The governing
equations are presented throughout the analysis to enable readers to develop their own results.
The 20,000-psi vertical monobore tree fatigue analysis depends on the operations forecasted for its life
cycle. Using the expected load histogram, a series of pressure and thermal analyses were executed to produce
cycles to failure. Implementing the Palmgren–Miner's rule enabled obtaining the total damage produced
by factory acceptance tests total field life shut-ins, and flow-in pressure cycles. This not only serves as
verification that the required safety factor is met per API Technical Report 17TR8 but also enables making
engineering assessments of "what-if" operations. In this sense, a change or addition of an operation will lead
to a simple recalculation of fatigue damage without requiring performing the analysis from the ground up.
The method also allows for computation of cycles to failure for a pressure range when the other pressure
ranges and conditions don't change.
In addition to the life cycle calculation, the method evaluates the damage of all nodes, which produces
full-contour plots. The contour plots, in addition to displaying the hot-spot locations, when used with
structural analysis results, enable the engineer to assess areas of improvement and product optimization.
The method proposed gives an effective way to communicate and recommend the design life capabilities
of a product to the operator to predict life expectancy for combinations of expected load scenarios.
2 OTC-29534-MS

Introduction
When developing HPHT subsea equipment covered by API SC 17, the design guidelines provided by API
Technical Report 17TR8 focus on a key step within the verification analysis, which is fatigue assessment.
Due to the high pressure and temperature exposure, HPHT equipment may experience high stresses and
strains at the surfaces which are exposed to these conditions. In addition to the structural design verification,
the fatigue life of the equipment needs to be assessed to ensure that premature failure does not occur during
its service life.
With these conditions in mind, a 20,000-psi vertical monobore subsea tree was developed to
accommodate such pressure and thermal loads. The Tree utilizes specialized materials and experiences
numerous operations during its life. This paper focuses on exposed nickel alloy components, typically used
in HH-trim class. These also happen to be some of the highly stressed components when exposed to 20,000-
psi pressure and are therefore designed with the high strength benefit of CRA materials. The production stab
connects the vertical monobore subsea tree to the tubing hanger to provide a conduit for the produced fluids
and the tubing hanger suspends the production tubing. Both components are used as examples in this paper.

Figure 1—20,000-psi vertical monobore subsea tree assembly that utilizes nickel alloy components.

Based on the component location and barrier philosophy in the assembly, the fatigue assessment method
may vary for the component. The components being reviewed are catalogued as secondary barrier elements.
These are thus analyzed using the strain-life fatigue assessment method in lieu of fracture mechanics.
Plenty of design standards and codes provide data for carbon and alloy steels. Nickel alloy data is limited
especially when it comes to strain-life data. This paper discusses the various material tests performed
to characterize the nickel alloy data utilized to perform strain-based fatigue-life calculations. The paper
also presents details on mean stress correction for material test data and fatigue-life calculations with
nonproportional loading.
OTC-29534-MS 3

Fatigue Analysis Methods


API requires fatigue screening or fatigue analysis to be performed using either stress-life, strain-life, or
fracture mechanics. Although fatigue screening out is the simplest path forward, it is not possible for
all equipment. If fatigue analysis is required, API Technical Report 17TR8 requires that it considers the
environment conditions the equipment will see for its life.

Fatigue Screening
API Technical Report 17TR8 allows performing fatigue screening per ASME Section VIII Div. 2 part 5.5.2,
which is limited to certain materials, can only be used with proportional loading, and uses elastic stress
results at a specific location of the component. Additionally, Method A limits you to a tensile strength of
80,000 psi, or Method B assumes that the component is taken to maximum capacity each cycle in the first
step. Consequently, many components do not satisfy this criterion, making fatigue analysis using material
properties in environment a requirement for most of the components of the permanently installed equipment
which are exposed to the shut-in pressure cycles of the well or external riser loads.

Stress- and Strain-Based Methods


A component's fatigue life is determined by the fluctuation of stress and the component material. Both
stress- and strain-based methods require three main things to perform a fatigue analysis: a material curve, an
alternating stress/strain, and a mean stress/strain. This means that, for a given stress/strain, the fatigue life
can vary based on the alternating stress and the mean stress. The mean stress correction factor is a method
to account for these variances.
API Technical Report 17TR8 allows use of multiple standards for stress-life fatigue analysis, including
ASME Section VIII Div.2 §5.5, ASME Section VIII Div.3 Article KD-3, DNVGL-RP-C203, or BS 7608.
DNVGL-RP-C203 is limited to analysis of steel materials subjected to external environmental loads and
high cycle fatigue. BS 7608 is limited to steels below an ultimate tensile strength of 53,000 psi.
While ASME Section VIII Div. 2 and ASME Section VIII Div. 3 have procedures for performing elastic-
plastic fatigue analysis, they use built-in fatigue curves for testing performed in air. The fatigue curves
are also limited to certain materials and use a modified stress-based approach with built-in mean stress
correction and safety factors. ASME Section VIII Div. 2 methodology assumes the maximum mean stress
to eliminate the need for a mean stress correction factor. This would produce overly conservative results
and does not give guidance to produce the stress-life curve in subsea environments. Methodology similar
to ASME Section VIII Div. 3 could be followed. Although the safety factor not to be replicated, API
Technical Report 17TR8 requires a safety factor of 10. ASME Section VIII Div. 3 uses a modified Goodman
methodology with a mean stress correction factor based on an endurance limit defined by 1E6 cycles. The
testing required to define the endurance limit would be schedule prohibitive for most projects.
In general, stress-life analysis is not accurate when plastic strains are present because this typically results
in a low-cycle fatigue life. The results of the static structural analyses of these HPHT components show that
low cycle life cannot be neglected, and therefore strain-life methodology is preferred.
API Technical Report 17TR8 also mentions methods from API Standard 597-1 Part 14 Level 3 that allows
for strain-based fatigue analysis, which is best for low-cycle fatigue evaluation and used as the basis for
defining the fatigue assessment in this report. API Standard 579-1 Level 3 assessment Method B provides
a new fatigue evaluation technology that includes the effects of nonproportional loading and load sequence
effects on fatigue life [5]. This standard uses coefficients from built-in material properties and modified
Brown–Miller strain-life method with a critical plane approach for methods A and B. Since methods and B
use a built-in curve, it would not be accurate for our environment. Although the Brown–Miller strain-life
equation is suitable for both high and low cycle fatigue, methods A and B also state that "other strain-life
equations may be used depending on the availability of model parameters for the material and temperatures
being evaluated." The widely accepted modified Coffin–Manson equation (1) with Morrow mean stress
4 OTC-29534-MS

correction equation is proposed for analysis of ductile materials and be used in lieu of the Brown–Miller
equation presented in API Standard 579-1. For simplicity in this paper, the equation is simply referred as
the Coffin–Manson equation.

(1)

Material Testing Program


This section discusses the variety of considerations required when developing an effective material testing
program for the development of the nickel alloy fatigue properties.

Environmental Considerations
API Technical Report 17TR8 is clear that material fatigue properties, data, or both must be representative of
operating conditions. Although it is true that testing in air does not represent subsea environments, subsea
components are exposed to a multitude of environments during different operations, such as:

• Air

• Water

• Completion fluids

• Seawater with or without cathodic protection

• Hydrocarbons

• Injected chemicals

In addition to the variety of environments, every environment can have varying concentrations depending
on the field and its operator. With such a wide variety of operating environments, it is crucial to determine the
appropriate environments to test in. Although being conservative is the traditional route, overly conservative
assumptions can lead to reduced material properties, ultimately limiting a component's ability to withstand
fatigue. In the case of this nickel alloy testing program, three different environments were tested in: air,
completion brine, and sour service at 300 degF. These environments were chosen based on the specific
operations performed by the nickel alloy components within the tree system. The chemical composition
of each environment was determined based on previous field experience and recommendations provided
by the end user.

Testing Parameters
Project schedules drive the different programs, ranging from costing, material testing, and engineering
design. Often, a program cannot be fully completed until others are concluded, and some cannot start without
proper assumptions. After environments in which testing will be conducted are selected, a testing regime
must be developed, and the required testing parameters determined. To obtain the necessary properties to
analyze the nickel alloy components in fatigue in compliance with API Technical Report 17TR8, three
separate tests are performed: strain limit damage (SLD), tensile tests at ambient (70 degF) and elevated
temperature (300 degF), strain-life (ε-N) tests in the specified environments, and slow-strain-rate tests
(SSRTs) in the specified environments. In addition, for strain-life tests, API Technical Report 17TR8
requires that two material heats are evaluated with a sample size of three for each at low-, mid-, and high-
cycle range (total of 18 specimens per environment). The same is true for testing as per DNVGL-RP-C203.
Low-, mid-, and high-cycle strain values were determined based on preliminary structural analysis results
performed on the tubing hanger, which was considered the most sensitive to fatigue because of its location
within the tree system and complex loading scenarios. A set of three strain ranges are obtained and used
OTC-29534-MS 5

as the basis of the strain-life materials testing program in the elastic, elastic-plastic, and localized plastic
regions. Figure 2 shows one of the test results at high temperature.

Figure 2—Strain-life tests.

Theory in Practice
Determination of Fatigue-Life Parameters
Strain-life (ε-N) testing is typically performed in a fully reversed setup, where R = −1 and where there are
zero mean stresses. The nickel alloy tests were performed as "zero to tension," where R = 0, meaning each
stress-life point is at a different mean stress. Langer [12] recognized the importance of mean stress in the
performance of reliable fatigue analyses. Langer suggested that the problem could be handled by adjusting
the fatigue curve downward in the high-cycle regime to reduce the sustained number of cycles as shown in
Figure 3. For an infinite number of mean stresses, an infinite number of corrected lines create a surface.

Figure 3—Fatigue curve—API Standard 579-1.


6 OTC-29534-MS

Figure 4—Coffin–Manson surface.

The Coffin–Manson equation (1) contains σ′f, ε′f, E, b, and c, which are unique to a material and
environment. This equation represents a surface from where Nf = f(εa, σm) and can describe the laboratory
test results and predict fatigue life of components of the same material. The test results are plotted as points
lying very close on this surface.

Design Points and Statistical Corrections


API Technical Report 17TR8 presents an S-N design curve that is at M-2S from the test results. Aerospace
literature and other design codes such as in DNVGL-RP-C203 present a method for obtaining a design curve
based on probability of survival and confidence levels. DNVGL recommends that the design curve provide
a 97.7% probability of survival with at least a 75% confidence interval. The design curve is generated when
the test data shows a Gaussian distribution when plotted in a logarithmic format as shown in Figure 5.
DNVGL presents a SMF for calculating the shift for the elastic high-cycle S-N curve. This method, however,
assumes that a constant mean stress (regardless of stress amplitude) occurs between all points and that a
linear regression line can be plotted in the mean of the test data.

Figure 5—Design curve—Lostberg.

Since the nickel alloy ε-N tests performed are data points on different mean stress curves, a SMF cannot
be obtained. An important consideration about this method is that the distribution of each set of strains for a
given strain amplitude also follow a Gaussian distribution in a logarithmic format. A method is developed
where a ε-N design curve can be obtained if the mean strain of each strain range is shifted by a set number
of standard deviations that capture a 97.7% probability of survival and a 75% confidence in the logarithmic
scale format. The Fatigue Design of Marine Structures by Inge Lostberg [7] includes the number of standard
deviations to be subtracted from the mean to derive the design curve as shown in Figure 6. For this case
study, an average of five useful data points is obtained from the nickel alloy tests for an environment and a
OTC-29534-MS 7

strain range. The population standard deviation is unknown; therefore, a value of 2.95 standard deviations
is used for the subsequent calculations.

Figure 6—Standard deviations from mean—Lostberg.

The calculation of the design points can be obtained using the following steps:
1. Obtain the logarithmic value of the cycles to failure for each data point as the , with being
the value for cycles to failure.
2. Compute the average of the logarithmic values of cycles to failure, calculated in step 1, for each strain
range as , with as the first value for cycles to failure for a strain
range, and as the final value of cycles to failure for the same strain range.
3. Compute the sample standard deviation of the logarithmic values for cycles to failure of each strain
range as .
4. Obtain a new cycle to failure design point for each strain range as:

It may be necessary to neglect outliers as they can negatively underestimate the shift of a design point,
as is demonstrated in the figure below. The design points using the modified DNVGL method for these
test data produce more conservative results than the method proposed by API Technical Report 17TR8.
Figure 7 shows the example of design points for Environment 2. The method is repeated for each set of
environments and temperatures.
8 OTC-29534-MS

Figure 7—Test results at Environment 2.

Fatigue Life coefficients


The constants to the Coffin–Manson Equation are obtained using the design points previously derived.
Norman Dowling [8] suggests that parameters related to the material limits such as the fatigue ductility
coefficient and the fatigue strength coefficient are similar to the true strain at fracture and true stress at
fracture respectively such that σ′f ≈ σf and ε′f ≈ εf. The stress, strain at fracture, and elastic modulus E are
obtained from the monotonic tensile tests performed at environment and at temperature. The fatigue ductility
exponent c and the fatigue strength exponent b are also material dependent, but before they can be obtained
using a regression algorithm, the mean stress for each strain range must be obtained. The mean stress for
tests that are not performed with a fully reversed stress state can be cumbersome to determine, because the
stress/strain value is dependent on the cyclic hardening of the material.
API Technical Report 17TR8 provides different methods for the evaluation of the fatigue tests. For this
case ASTM E606 is used (Figure 8). In this standard, the strain is controlled in a feedback control loop by
monitoring the displacement of the neck cross section of the tensile coupon. This displacement is the same
as the engineering strain range.

Figure 8—ASTM E606

For the study of the fatigue curve in Environment 2, an FEA model is created for each strain range set.
Each model is constrained in a way that a boundary displacement produces a uniaxial strain range. When
the mean stress is zero, a kinematic hardening material model for nickel alloy at Environment 2 can be
generated using a four-point correlation method to determine the cyclic strength and cyclic strain hardening
coefficients in the stabilized cyclic strain hysteresis loops using the mean data points. For this case, the
kinematic hardening model is approximated using a combined hardening model with backstresses and half-
OTC-29534-MS 9

cycle from mean data of the monotonic tensile tests at temperature and at environment. It is noted that the
combined hardening model with three backstresses in an FEA modeling software produces shakedown and
stabilization of the hysteresis loop almost immediately. The mean stress is computed observing the stress
state at an arbitrary point in Time Step 1 (TS1) where γ ≈ Δε and at Time Step 2 (TS2) where γ ≈ 0.

(2)

Table 1—Strain Range and Mean Stresses

With the retrieval of the mean stresses through FEA, it is now possible to calculate the missing b and c
coefficients of the Coffin–Manson Equation. Because the design points are not exact solutions to the Coffin–
Manson Equation, a regression algorithm is used to approximate b and c. This algorithm is effectively
approximating the Coffin–Manson surface where these three design points lie in.

Figure 9—Environment 2 design surface.

Verification of the Coffin–Manson Constants


API Standard 579-1 Level 3 assessment method B utilizes an elastic-plastic stress analysis in combination
with a critical plane approach. The critical plane is obtained by rotating the stress and strain tensors from
which normal component produces the smallest fatigue life. This is achieved by postprocessing the FEA
solution at either nodal or integration points. The user can create its own subroutine script, but commercially
available software eliminates the need for this time-consuming task. To verify the calculated Coffin–Manson
constants, the results of the static structural FEA from the tensile coupon previously used to obtain the mean
stresses are imported into the software. The cycles to failure are then calculated. The results are compared
against the cycles to failure of the design points and are shown in Table 2. The small error can be attributed to
the inexact solution to the Coffin–Manson surface and uncertainties in the reported material testing results.
10 OTC-29534-MS

Table 2—Fatigue Life Error

Nickel Alloy Hardening Curves


API Technical Report 17TR8 requires that an adjusted true-stress true-strain material model be used for
the elastic-plastic analysis. Material curves can be obtained from ASME Section VIII Div. 2 Annex 3D,
ASME Div. 3, KD-231.4, or testing. A question existed whether the proprietary nickel alloy performed as
well as the curves defined in ASME. The testing program included two heats of different forging sizes at
room and at temperature with twelve tensile coupons. A correction is performed to account for statistical
variability of the material hardening. The design material hardening curves are created at a M-2S down for
each temperature. The comparison of these curves with ASME are shown in Figure 10. The ASME curve at
elevated temperature is created by applying a derating factor Yr from API 6A Table G4 to both the SMYS
and UTS. It is noted that the testing program demonstrated that the proprietary nickel alloy performs better
than the hardening curves of ASME.

Figure 10—Nickel alloy comparison with ASME.

Case Study
One of the most critical components in the 20,000-psi vertical monobore subsea tree is a production stab
that experiences multiple pressure reversals, particularly during the factory acceptance and subsea testing
phases. Having the thinnest wall thickness required, this component underwent fatigue life verification per
the FMECA and barrier philosophy requirements of API Technical Report 17TR8.

Load Histogram
API Technical Report 17TR8 requires a load histogram that includes loads, pressures, and temperatures for
significant events that are applied to the equipment. An example of the load histogram, which is developed
jointly between the end user and the manufacturer, is displayed below. External pressure is represented as
negative pressure. The histogram includes hydrostatic testing, factory acceptance testing, system integration
OTC-29534-MS 11

testing, subsea landing loads, subsea preproduction testing, startups and shut-ins, pressure drops, flow-
induced vibrations, flow-ins, maintenance rig operations, and emergency shutdowns throughout its lifetime
along with a conservative number of repetitions for each expected event denoted as n# in the load histogram
of Figure 11. This load histogram is used in the static structural FEA for producing the histogram of stresses
and evaluating the Palmgren–Miner's rule calculation for fatigue damage. The sequence of events is mostly
critical in the load step sequencing of the static structural FEA due to the kinematic shifts and Baunschinger
effects that occur during load reversals. Therefore, a review of historical equipment performance and good
engineering judgment is required when defining the load sequence.

Figure 11—Production Stab Load Histogram

Static Structural FEA


The production stab case study is a component that is not subjected to any external bending loads but
exposed mainly to the pressure and temperature cycles. The production stab is analyzed as an axisymmetric
model using quadratic elements with full integration. A fine mesh is applied in fillets and chamfers, and
no defeaturing in critical loaded areas occur. A combined kinematic hardening model is used that includes
the nickel alloy M-2S half-cycle isotropic material hardening curves. A field output is requested for all
timesteps which is used in subsequent fatigue life postprocessing. A total of 28 load steps are created to
produce strain ranges between stressed and relaxed states for the 14 blocks produced in the load histogram.
The results for each load step are reviewed to ensure that all loads are applied properly.

Figure 12—Production stab static FEA.


12 OTC-29534-MS

Surface Finish Correction Factor


Microscopic imperfections in material homogeneity and machining increases the possibility of crack
initiations. These imperfections are mainly accounted in during the machining stage of the component and
often indicated in the engineering drawing with a surface finish. The correlation that shows the effect of
machining finish to fatigue life is best described by the surface finish correction factor Kt. This correction
factor varies slightly between numerous literatures. The FEA software offers built-in correction factors that
are dependent of surface roughness (Ra) and material ultimate strengths (UTS). Often, different areas of the
component will have a different surface finish, and the hot-spot location is yet to be determined. A method
is presented here to obtain the surface finish which yields a conservative value of the component fatigue
life shown in the steps below.
1. Obtain preliminary damage results without a surface finish penalty factor Kt.
2. Observe the location of the lowest life/largest Df (hot spot).
3. Compare the hot-spot location against the machining drawing and choose the corresponding surface
finish.
4. Apply the new Kt to the whole model and reanalyze.

Fatigue Analysis Results


For components that cannot be inspected, the calculated life should be a minimum of 10 times the service
life, according to API Technical Report 17TR8. Such is the case for this production stab. The interpretation
for this statement is that, in the calculation of fatigue damage within the Palmgren–Miner's rule, a safety
factor of 10 must be included. This reduces limits the accumulated damage to 0.1 as is shown in the Equation
3.

(3)

The results from the static structural FEA at each integration point in the model are postprocessed using
the modified Coffin–Manson Equation with the corresponding environment constants (σ′f, ε′f, E, b and c) to
obtain the sustained number of cycles Nf for each block in the rainflow cycle counting. Then, the Palmgren–
Miner's rule is applied to each block with the minimum expected number of cycles for each block n#. The
critical plane is obtained as the plane where the normal stress and strain at each integration point yields
the smallest fatigue life.

(4)

The total accumulated damage at each integration point is displayed as a contour plot for the whole
production stab body in Figure 13. This allows the visualization of the hot-spot location. The location of the
hot spot determines the selection of the surface finish correction factor to be applied to the whole model.
The fatigue analysis is then iterated one more time to produce the results. If the total accumulated damage
is less than 0.1, the fatigue analysis is satisfied as per API Technical Report 17TR8.
OTC-29534-MS 13

Figure 13—Production stab accumulated fatigue damage.

Generating Preliminary Results in Variating Histograms


In many cases, the load histogram varies between wells, and preliminary calculations are necessary to
determine the feasibility of the system in a new well or a variation of the load histogram. Monitoring and
recording of a subsea tree is more common, and the real-life histogram may differentiate from the initial
expected load histogram. An unexpected load could reduce the remaining life in the same way that an
unexpected larger number of cycles has, and lower loads could increase the remaining life in the same way
that a smaller number of cycles has. Any variation from the histogram would invalidate the fatigue analysis.
Then the question arises: "What if?" With an infinite number of plausible histograms, a method is proposed
that allows a rapid estimation of fatigue life without the need to reevaluate the static structural FEA or the
postprocessing of the data as part of the fatigue analysis, which can be used to calculate preliminary results
every time there is a change.
It is noted that, in the case of the production stab, the biggest damage per repetition occurs during the
first six blocks of the equipment testing. Such as in the case of hydrostatic testing and autofrettage, the high
loads would generate plastic sets that have the most influence in the subsequence loads. This leads to the
idea of a creation of tables that can be used to input directly into the Palmgren–Miner's rule for calculation
of fatigue that also includes a safety factor of 10:

(5)

Where

(6)

Where X7 is the sustainable number of cycles for a particular block. The first term is containing the
accumulated damage for the first 6 blocks and is kept as a constant. The calculated damage D7 for each
block P7 is simply the inverse of the number of the Coffin–Manson Equation solution Nf,7 for block 7. A
static structural FEA is created with a compact load histogram. In this analysis example the loads steps for
blocks P1 to P6 are solved only one time. A restart solution allows to solve for block P7 as shown in Figure
14. An array of varying differential pressures is created for as many block P7 as are wanted. In this example,
block P7 is ranged from Δ20,000 to Δ1,000 psi.
14 OTC-29534-MS

Figure 14—Load Histogram for ΔPs

The solutions to the sustainable cycles are recorded in Table 3. Note that Table 3 is shown as an example
and is not representative of the actual component analysis.

Table 3—Maximum Damage for ΔPs at an Integration Point

The new modified Palmgren–Miner's rule approximation for a given load histogram is:

(7)

Where φ…ψ represents a choice of combinations in a ΔP in a certain environment of block 7. For example,
let's assume that a histogram includes testing events, 111 repetitions of a ΔP = 19,500 psi in Environment 2,
OTC-29534-MS 15

222 repetitions of a ΔP = 15,00 psi in Environment 1, and 333 repetitions of a ΔP = 5,000 psi in Environment
3. The modified Palmgren–Miner's rule approximation would be:

(8)

(9)

(10)
This solution is used as an indication that this component is likely to sustain a load histogram with this
combination. It is agreed that the Palmgren–Miner's rule should be applied at each integration point/node
individually. However, this method looks at the maximum damage for an integration point and assumes
that the hot-spot location does not change. This is mostly true for nonreciprocating loading conditions.
A comparison study shows that the results using this approach and a full-load histogram-sequenced FEA
vary within 4% of each other for this component, which gave confidence in the use of this method only as
a tool for producing preliminary results. A full-sequenced fatigue analysis will be required to satisfy the
requirements of API Technical Report 17TR8.

Visualizing Histogram Damage


To fully answer the "what-if" question, a visual aid is often valued in a decision matrix. Each block damage
is plotted as a percentage of the total damage. Plotting this in a pie chart allows rapid assessments of the
influence of load conditions plus repetition combinations which can help determine the criticality of added
subsea or rig operations, such as the one shown in Figure 15. In this example, the combination of 111
repetitions with a ΔP = 19,500 psi in Environment 2 cause the biggest life damage.

Figure 15—Percentage of total damage.

Conclusion and Recommendations


Historically, equipment designed to API Specs 6A and 17D has had limited focus on fatigue life assessment
because of environmental loads. HPHT equipment is exposed to conditions of temperature and pressure
that present stresses beyond the traditional allowables in typical analysis methods. Material properties and
environmental performance now require more rigorous testing programs within the HPHT design guidelines
of API Technical Report 17TR8. The modern analysis methods evaluate multiaxial stress-strain conditions
which require advanced understanding of material behaviors and complex multiaxial FEA. The methods
16 OTC-29534-MS

presented here served as a basis for verification of safety factors of multiple HPHT equipment, which
ranged from a tree cap, tubing hanger, production stabs, and hydraulic couplers made from nickel alloy.
The formulation and visualization of rapid preliminary results provided guidance to the project team and
end user in the assessment of foreseeable operations. The methods presented in this paper will allow future
projects to deliver fatigue feasibility studies at an early age for this equipment.
A series of recommendations follow the conclusion which hope to facilitate future material testing
programs and analysis of HPHT equipment. On the material testing size, it is recommended to use a mean-
minus-two standard deviations as a statistical correction of true stress-true strain curves, which is not
mentioned in the guidelines of API Technical Report 17TR8. To obtain fatigue curves that have a higher
degree of confidence, it is recommended to carry on more than the three minimum strain ranges required by
API Technical Report 17TR8. The span of stress ranges should vary from values of magnitude much lower
to the yield strength and approximate the ultimate tensile strength.
ASME Section VIII Div. 2 and Div.3 and DNVGL-RP-C203 fatigue analysis methods contain inherent
limitations within the design philosophy of API Technical Report 17TR8 due to the limitations of built-in
fatigue curves in air based on high-cycle fatigue methods. The provisions of API Standard 579-1 FFS offer
a more comprehensive assessment to the fatigue analysis for ductile materials in both the low- and high-
cycle regions and is therefore the recommended method. In situations where the strain tests are not fully
reversed, the effect of the mean stress do not allow a simple regression line to pass through the mean of
all data points. If the data follows a Gaussian distribution, a 97.7% probability of survival with at least a
75% confidence level in many cases produce more conservative results than M-2S recommended by API
Technical Report 17TR8 for each strain range set.

Abbreviations and Acronyms


API American Petroleum Institute
ASME American Society of Mechanical Engineers
ASTM American Society for Testing and Materials
CRA Corrosion Resistant Alloy
DNVGL Det Norske Veritas Germanischer Lloyd
FEA Finite Element Analysis
FFS Fitness-For-Service
FMECA Failure Mode Effects and Criticality Analysis
HPHT High Pressure High Temperature
NACE National Association of Corrosion Engineers
SC Subcommittee
SLD Strain Limit Damage per ASME Section VIII Div.2/3
SMYS Specified Minimum Yield Strength
SMF Stress Modification Factor
SSRT Slow Strain Rate Testing per NACE TM0198
UTS Ultimate Tensile Strength

Nomenclature
Avg Average
b Elastic strain slope of measured on the logarithmic plot of vs Nf
c Plastic strain slope of measured on the logarithmic plot of vs Nf
Df Accumulated Fatigue Damage
ε True Strain
OTC-29534-MS 17

ε-N Strain-Life
εa True Strain Amplitude
εe True Elastic Strain
εp True Plastic Strain
ε′f Fatigue Ductility Coefficient
εf True Strain at Fracture
Δε Total True Strain Range
E Elastic Modulus
γ Engineering Strain
Kt Surface Finish Correction Factor
Log Logarithm base 10
M Number of load blocks for service histogram
M-2S Mean Minus 2 Standard Deviations
Nf Fatigue Life
nk Number of cycles of block from service histogram
Nf,k Fatigue Life for a load block
σ Stress
σ′f Fatigue Strength Coefficient
σf True Stress at Fracture
σm Mean Stress
R Ratio of minimum stress to maximum stress
Ra Arithmetic Mean Deviation of Assessed Profile
S-N Strain-Life
Stdev Sample Standard Deviation

References
1. API Technical Report 17TR8 2nd Ed. High-Pressure High-Temperature (HPHT) Design
Guidelines
2. ASME BPVC VIII, Div. 2 (2017) ASME Boiler and Pressure Vessel Code, Section VIII, Division
2
3. ASME BPVC VIII, Div. 3 (2017) ASME Boiler and Pressure Vessel Code, Section VIII, Division
3
4. DNVGL-RP-C203 (April 2016) Fatigue Strength Analysis of Offshore Steel Structures
5. API Standard 579-1 (June 2016) Fitness for Service
6. ASTM E606-12 Standard Test Method for Strain-Controlled Fatigue Testing
7. Cambridge University Press, March 2016. Fatigue Design of Marine Structures, by Inge Lostberg
8. Mechanical Behavior of Materials 2nd Edition, by Normal Dowling
9. IIW-XIII-WG1-114-03 Best Practice Guide on Statistical Analysis of Fatigue Data
10. NASA Contractor Report 3697 Statistical Summaries of Fatigue Data for Design Purposes
11. Springer Netherlands, 2009 Fatigue of Structures and Materials, by Jaap Schijve
12. Design of Pressure Vessels for Low Cycle Fatigue, J. Basic Eng., Vol 84 (No3), by F. B. Langer
13. Metal Fatigue Analysis Handbook, 2012. Strain-Based Fatigue Analysis and Design by. Yung-li,
et.al

S-ar putea să vă placă și