EPRI - Guided for piping stress analysis

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 90

Guide for Piping Stress Analysis and Assessment

1019632

10371024
10371024
Guide for Piping Stress Analysis and Assessment
1019632

Technical Update, December 2010

EPRI Project Manager

K. Coleman

ELECTRIC POWER RESEARCH INSTITUTE


3420 Hillview Avenue, Palo Alto, California 94304-1338  PO Box 10412, Palo Alto, California 94303-0813  USA
10371024800.313.3774  650.855.2121  askepri@epri.com  www.epri.com
DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN ACCOUNT OF
WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH INSTITUTE, INC. (EPRI).
NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE ORGANIZATION(S) BELOW, NOR ANY
PERSON ACTING ON BEHALF OF ANY OF THEM:

(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I) WITH
RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM
DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS FOR A PARTICULAR
PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR INTERFERE WITH PRIVATELY OWNED
RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL PROPERTY, OR (III) THAT THIS DOCUMENT IS
SUITABLE TO ANY PARTICULAR USER'S CIRCUMSTANCE; OR

(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER (INCLUDING
ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE HAS BEEN ADVISED
OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR SELECTION OR USE OF THIS
DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED IN
THIS DOCUMENT.

THE FOLLOWING ORGANIZATION, UNDER CONTRACT TO EPRI, PREPARED THIS REPORT:

Stress Engineering Services

This is an EPRI Technical Update report. A Technical Update report is intended as an informal report of
continuing research, a meeting, or a topical study. It is not a final EPRI technical report.

NOTE
For further information about EPRI, call the EPRI Customer Assistance Center at 800.313.3774 or
e-mail askepri@epri.com.

Electric Power Research Institute, EPRI, and TOGETHERSHAPING THE FUTURE OF ELECTRICITY
are registered service marks of the Electric Power Research Institute, Inc.

Copyright © 2010 Electric Power Research Institute, Inc. All rights reserved.

10371024
ACKNOWLEDGMENTS
The following organization, under contract to the Electric Power Research Institute (EPRI),
prepared this report:
Stress Engineering Services
Cincinnati, OH
Principal Investigator
P. Carter

This report describes research sponsored by EPRI.

This publication is a corporate document that should be cited in the literature in the following
manner:

Guide for Piping Stress Analysis and Assessment. EPRI, Palo Alto, CA: 2010. 1019632.
10371024 iii
10371024
ABSTRACT
This report gives recommendations for piping analysis methods and integration with inspection
and operating data to give estimates of risk and remaining life. The scope of the document is as
follows:
 Objectives of piping assessment
 Selected literature review of assessment practices for piping and other relevant components
 Failure modes
 Inspection and operating data
 Piping analysis: methods and data
 Risk assessment
 Integration of analysis, inspection, and operating data for assessments and recommendations
for the following:
 Repairs
 Inspection
 Sample testing
 Operating conditions and effects of changes and derates
 Re-assessment
 Applications to hot reheat and main steam examples

The use of an accurate stress analysis and life assessment piping model as part of ongoing piping
risk management is the key recommendation. The significance of changes such as varying
hanger performance, operating conditions, and new inspection data should be understood with
reference to the piping model. The justification for inspection intervals, sample testing, or a
decision to run at a reduced temperature with a damaged component should be based on a life
and risk assessment using the updated model.
A comprehensive high-energy piping model is therefore an investment aimed at optimizing the
life and performance of a piping system.
A range of options in terms of model and element types and analysis methods is presented. This
covers shell, elbow, and pipe models and creep, limit, and shakedown analysis.
These are used to estimate remaining life and risk and to evaluate the following:
 Creep rupture
 Effects of girth and seam weld
 Cyclic loading and accelerated creep due to thermal shock and locked hangers
 Use of risk and remaining life approaches

10371024 v
In summary, this report sets out practical analysis methods that provide a defensible basis for
decision making, taking into account design, history of operating conditions, inspection, and
testing.

Keywords
Hangers
Hot reheat
Main steam
Piping
Stress analysis

10371024 vi
CONTENTS
1 INTRODUCTION ....................................................................................................................1-1
2 OBJECTIVES OF PIPING ASSESSMENTS ..........................................................................2-1
3 SELECTED LITERATURE .....................................................................................................3-1
3.1 Piping Assessment...................................................................................................3-1
3.2 Creep Stress and Damage Analysis ........................................................................3-2
3.3 CDM Methods ..........................................................................................................3-2
3.4 Decoupled Creep and Damage Analysis .................................................................3-3
3.5 Simplified Analysis Methods.....................................................................................3-4
3.6 Multiaxiality and Constraint ......................................................................................3-6
3.7 Defect Assessment ..................................................................................................3-7
4 PIPING ANALYSIS AND ASSESSMENT ..............................................................................4-1
4.1 Introduction...............................................................................................................4-1
4.2 Differences Between Analysis for Design and Life Assessment ..............................4-1
4.3 Analysis for Risk and Life Assessment ....................................................................4-2
4.4 Sources of Data for Stress Analysis and Life Assessment ......................................4-3
4.5 Multiaxiality...............................................................................................................4-3
4.6 Weldments ...............................................................................................................4-4
4.7 Peaking and Out-of-Round Piping............................................................................4-4
4.8 Aspects of a Piping Model........................................................................................4-5
4.9 Identification of Hanger Problems ............................................................................4-6
4.10 Typical Data Requirements for Piping Assessment .................................................4-7
4.11 Piping Model Element Comparison ..........................................................................4-7
4.11.1 Conventional Beam Elements with Independent Pressure Calculations..........4-7
4.11.2 Shell Elements .................................................................................................4-8
4.11.3 Elbow Elements................................................................................................4-9
4.12 Analysis Methods .....................................................................................................4-9
4.12.1 Fully Coupled CDM Modeling.........................................................................4-10
4.12.2 Decoupled Stress and Damage Analysis .......................................................4-10
4.12.3 Isochronous Elastic-Plastic Analysis ..............................................................4-10
4.12.4 Steady Load Life Calculation: Limit Analysis..................................................4-11
4.12.5 Cyclic Load Creep Life Calculation: Shakedown Analysis .............................4-12
4.13 Piping Example ......................................................................................................4-14
4.13.1 Elastic Analysis ..............................................................................................4-16
4.13.2 Prediction of Creep Deflection........................................................................4-17
4.13.3 Prediction of Creep Rupture and Girth Weld Cracking...................................4-19
4.13.4 Cold Pull .........................................................................................................4-20
4.13.5 Summary of Constant Load Analyses ............................................................4-20

10371024 vii
4.13.6 Cyclic Analysis ...............................................................................................4-21
5 HOT AND COLD WALKDOWN AND INSPECTION..............................................................5-1
6 DESIGN AND OPERATING CONDITIONS ...........................................................................6-1
7 ASSESSMENT AND RISK ANALYSIS..................................................................................7-1
7.1 Risk-Based Assessment ..........................................................................................7-1
7.1.1 Background ........................................................................................................7-1
7.1.2 Data....................................................................................................................7-1
7.1.3 Derivation of Failure Probability .........................................................................7-2
7.1.4 Use of Probability to Manage Component Risk..................................................7-3
8 HEAVY SECTION WYES AND TEES....................................................................................8-1
8.1 Rupture Life Prediction.............................................................................................8-1
8.2 Transient Thermal Stress and Creep Fatigue ..........................................................8-1
8.3 Sustained Stress and Creep Rupture.......................................................................8-2
8.4 Creep-Fatigue Damage............................................................................................8-3
9 REFERENCES .......................................................................................................................9-1
A APPENDIX 1: CASE STUDIES: HOT REHEAT AND MAIN STEAM PIPING ..................... A-1
A.1 History ..................................................................................................................... A-1
A.2 Conclusions from Reports ....................................................................................... A-2
A.3 HR Piping ................................................................................................................ A-2
A.3.1 Analysis Cases ................................................................................................. A-3
A.3.2 Analysis 1: Elastic-Plastic Thermal-Mechanical Loading.................................. A-4
A.3.3 Analysis 2: Isochronous Elastic-Plastic Analysis .............................................. A-5
A.3.4 Analysis 3: Creep and Damage Analysis.......................................................... A-7
A.3.5 Analysis 4: Limit Analysis for Reference Stress Creep Damage Calculation ... A-8
A.3.6 Analysis 5: Shakedown Analysis to Identify Critical Hangers ......................... A-10
A.4 Main Steam Piping ................................................................................................ A-11
A.4.1 Design Data .................................................................................................... A-11
A.4.2 Analysis Cases ............................................................................................... A-12
A.4.3 Comparison of Reinforced and Unreinforced Joints Between Turbine
Leads and Piping ............................................................................................ A-12
A.4.4 Elastic-Plastic Isochronous Analysis for Deflection and Strain ....................... A-14
A.4.5 Creep and Damage Analysis .......................................................................... A-14
A.4.6 Limit Analysis and Reference Stress Calculations of Life............................... A-14
A.4.7 Calculation of Probability of Failure ................................................................ A-15
A.4.8 Conclusions .................................................................................................... A-19
B APPENDIX 1: ISOCHRONOUS AND CREEP DATA: GRADE 22 AND GRADE 91........... B-1

10371024 viii
LIST OF FIGURES
Figure 3-1 MPS and von Mises limit surfaces for piping and shell elements .............................3-7
Figure 4-1 Yield and minimum rupture data for grade 22 material ..........................................4-14
Figure 4-2 Shell elastic analysis ..............................................................................................4-15
Figure 4-3 Elastic vertical deflection = 5 in. (127 mm) .............................................................4-17
Figure 4-4 Shell creep deflection after minimum life (1.1e6 hours) = 48 in. (1219.2 mm) .......4-18
Figure 4-5 105-hour isochronous stress-inelastic strain data for grade 22 at 1000°F
(537.8°C)..................................................................................................................................4-18
Figure 4-6 Creep strain prediction (0.35%) after 1  105 hours using isochronous stress-
strain data for grade 22 at 1000°F (537.8°C) (The deflection in Figure 4-2 does not require
large creep strains.) .................................................................................................................4-19
Figure 7-1 Distribution of grade 22 rupture data ........................................................................7-2
Figure 7-2 Mean and minimum life, damage accumulation with cumulative failure probability
(cdf), and probability rate per year (pdf) for grade 22 design stress at 1000°F (537.8°C) .........7-4
Figure 7-3 Effect of temperature derate on risk approaching minimum life ...............................7-4
Figure 7-4 Results of sample testing at 20 and 35 years showing decreasing mean
properties, with scatter in sample tests less than nominal, but unaffected by time ...................7-5
Figure 7-5 Conclusions for piping risk assessment from sample tests at 20 and 35 years .......7-5
Figure 8-1 (a) Wye section with MPS contours and (b) Wye section with stress histories ........8-3
Figure A-1 HR line with constant load hangers......................................................................... A-3
Figure A-2 HR line: elastic-plastic analysis............................................................................... A-4
Figure A-3 Isochronous elastic-plastic analysis. Strain (0.7%) and vertical displacement
(−19 in. [482.6 mm] at horizontal bend): 1e5 hours ............................................................ A-5
Figure A-4 Isochronous elastic-plastic analysis. Strain and vertical displacement:
5e5 hours The horizontal bend has not dropped significantly after 100,000 hours............. A-6
Figure A-5 Creep analysis showing mean creep life calculated from von Mises stress
limited by top wye and outlets to the turbine at 1.4e6 hours. Max creep strain = 1.2%. ..... A-7
Figure A-6 Use of yield, 1 × 105 and 5 × 105 hour rupture data to define “yield” stresses
for shakedown analysis..................................................................................................... A-10
Figure A-7 HR line with constant load hangers, identifying the critical hanger when locked .. A-10
Figure A-8 Inlets to upper wye are the predicted regions for creep-fatigue damage
when the critical hanger is locked ..................................................................................... A-11
Figure A-9 Predicted minimum life of unreinforced turbine lead joints = 25,000 hours........... A-13
Figure A-10 Temperature plot showing how the turbine leads unreinforced joints are
eliminated from creep calculations.................................................................................... A-14
Figure A-11 Predicted 100,000-hour vertical deflections for original hanger loads ................ A-15
Figure A-12 Isochronous elastic-plastic analysis .................................................................... A-16
Figure A-13 Damage plots identifying the critical turbine lead with minimum life =
1.0 × 105 hours.................................................................................................................. A-17
Figure A-14 Comparative risks: turbine leads with and without thermal stress and grade
22 material design stress .................................................................................................. A-19

10371024 ix
10371024
LIST OF TABLES
Table 4-1 Example piping data ................................................................................................4-15
Table 4-2 Constant load hangers.............................................................................................4-16
Table 4-3 Summary of analyses: example piping model .........................................................4-21
Table A-1 Design data .............................................................................................................. A-2
Table A-2 Interpretation of limit analysis for creep life prediction ............................................. A-9
Table A-3 Design data ............................................................................................................ A-11
Table A-4 Main steam preloads and constant loads (in kips) ................................................. A-13
Table A-5 Reference stress life calculations........................................................................... A-18
Table B-1 Grade 22 at 1000°F (537.8°C).................................................................................. B-1
Table B-2 Grade 22 at 1050°F (565.6°C).................................................................................. B-1
Table B-3 Grade 91 at 1000°F/537.8°C .................................................................................... B-2
Table B-4 Grade 91 at 1050°F (565.6°C).................................................................................. B-2
Table B-5 Grade 91 at 1100°F (/593.3°C)................................................................................. B-3

10371024 xi
10371024
1
INTRODUCTION
The complexity and features of high-energy piping systems mean that special analysis and
assessment methods are required compared with B31.1 design methods [1] and fitness-for-
service methods, as in the ASME/API standard [2]. This document aims to provide guidelines on
analysis methods and use of inspection and other data to meet the specific objectives of high-
energy piping assessment. Recently, the B31.1 Code Committee developed draft requirements
for inspection and maintenance of covered piping systems. With a systematic program for
material assessment and piping remaining life and risk assessment, this will provide a sound and
defensible basis for operational and risk management.
Fitness for service refers to methods for managing the risks in an aging or damaged plant. In the
context of high-energy piping, this generally means creep and fatigue of welded pressure parts
with additional factors. The use of constant load hangers over extended periods is likely to lead
to deflections greater than the hanger limits. The B31 design procedures are unlikely to prevent
the possibility of excessive deflections with constant load hangers. Therefore, in addition to
problems of corrosion and inaccessibility leading to neglect and deterioration in performance,
some constant load hangers in a piping system are likely to “bottom” or “top” out over time.
When this happens, the effects of cyclic loading can change. Increased failure risks due to such
changes are likely to be in girth weldments:
 Unidentified seam welds and multiple girth welds.
 Water/steam hammer leading to distortion and possible girth-weld damage.
 100% inspection of welds is usually impractical.

As with conventional high-temperature welded pressure parts, the questions and issues to be
answered are typically the following:
 Use of design and operating data to determine “safe remaining life”
 Sources of material data
 Methods, extent, and frequency of inspection
 Sampling and testing to determine remaining life
 The risk of weld failure
 Significance of overtemperature events
 Effectiveness of down-rating to extend remaining life.
 Significance of “cold pull”

10371024 1-1
This report focuses on structural aspects of high energy piping life, particularly the following:
 Girth weld cracking
 Excessive deflections leading to hanger difficulties
 Reduction in piping life due to cyclic loading

These failure modes have a common basis in yielding, creep deformation, and rupture.
Assessment of defects is mentioned but is not a significant part of the present scope.
The data and analysis methods were selected to make them accessible to as wide a group of users
as possible. The stress analysis methods may be performed with any general commercial
nonlinear finite element program. The methods were selected to make use of generally available
data in the ASME/API standard [2] and ASME III Subsection NH [3].

10371024 1-2
2
OBJECTIVES OF PIPING ASSESSMENTS
This report is intended to provide guidance on the scope and methods of piping assessment to
provide the piping owner or operator with a sound and defensible basis for decisions on
operation, inspection, and testing.
The scope of this report is therefore as wide as necessary to capture possible piping failure
modes and possible contributors to failure. It is assumed that the design complies with the
relevant version of the B31.1 Code [1]. If some design discrepancy associated with a low and
anomalous remaining life estimate did exist, it is expected that the analyses described in this
report would identify it.
Integration of inspection and monitoring data with analysis results is necessary for a complete
assessment. From the assessment, recommendations for future inspection and monitoring may be
made that reflect the most comprehensive view of risk that is possible. There is therefore a
process of data acquisition and other activities (for example, operating conditions, inspection,
monitoring, assessment, repairs) that need to be integrated and managed.
The failure modes and contributions to failure considered in this report include the following:
 Creep rupture of piping and welds
 Effects of weld misalignment and peaking
 Reduction in life due to cyclic effects (creep fatigue)
 Effect of thermal shock on heavy-walled components such as forged wyes and tees
 Dysfunctional hangers due to bottoming or topping out or other problems
 Piping approaching the end of nominal minimum life

10371024 2-1
10371024
3
SELECTED LITERATURE
3.1 Piping Assessment
The starting point for piping assessment is the B31.1 design code [1]. Stresses from sustained
loads are limited by the allowable stress, modified by a joint factor. Thermal expansion stresses
are limited to a combination of hot and cold allowable stresses, which take advantage of the cold
values. Assuming that the expansion stress limits are effective, piping minimum life should
therefore be at least that of the code-allowable stress. In other words, for correctly designed
piping, cyclic loading should not reduce the life to less than calculated from the design
temperature and allowable stress. Operating conditions are usually less severe than design
conditions. Further, because beneficial stress redistribution and relaxation occur due to creep,
piping life should be greater than that implied by design conditions. The ASME/API Fitness for
Service Standard [2] has data and methods for levels 1, 2, and 3 assessments. Level 1
assessments use materials-based screening techniques, and levels 2 and 3 assessments make use
of a range of stress analysis methods from hand calculations to full creep stress analysis.
The methods in this report fall into the category of level 3 assessments because they are detailed
inelastic analyses of creep deformation and damage and consider steady and cyclic loading.
The literature on piping life assessment is focused overwhelmingly on problems of cracks and
defects, secondly on inspection and monitoring, as seen in the International Journal of Pressure
Vessels and Piping.
Papers on analysis and assessment have typically followed the general approach of Balaschak
and Wray [4]. Here the need for a systematic approach is emphasized, which would integrate the
following:
 Documentation review
 Walkdown information
 Simplified (elastic) flexibility or stress analysis
 Nonedestructive evaluation (NDE) and field metallography
 Metallurgical assessment
 Remaining life assessment
 Monitoring plans

More recently, Hahn et al. [5] note that piping is often neglected compared with other power
equipment and that its dynamic and complex nature is often underestimated. Since at least the
early 1990s, there has been a steady number of conference papers describing hanger problems
and the need for regular inspection and maintenance to prevent the associated accelerated creep
and fatigue of piping. References 5 through 10 provide typical examples. The repeated themes

10371024 3-1
are constant load hangers reaching their expansion limits (even after adjustments), corrosion and
broken springs, and the need for frequent hanger walkdowns and maintenance. The results are in
some cases observable piping damage. A bottomed or topped out or broken spring usually means
significantly different loading on the pipe compared with design and therefore the possibility of
accelerated creep and fatigue and failure.
We will see that problems with well-maintained constant load hangers in high-energy piping can
be due to the sensitivity of piping deflection to relatively small creep strains. Constant load
hangers do not have the self-correcting properties that variable spring hangers have. This may
lead to excessive deflections over time.

3.2 Creep Stress and Damage Analysis


This report focuses on methods for high-temperature piping assessment, and so a description of
the modeling methods for creep stress analysis and damage calculation is helpful. The gap
between the linear elastic design analysis and the detailed modeling of creep damage will be
evident. As stated, there is no evidence that current design methods themselves are responsible
for faulty designs, other than possibly a reliance on constant load hangers. The objective here is
to demonstrate effective practical methods that are intermediate in complexity between elastic
piping analysis and the more research-oriented continuum damage mechanics (CDM) methods.

3.3 CDM Methods


The need for creep damage modeling was significantly motivated by high-temperature weld
failures that, although apparently understood in principle, had not been convincingly modeled
until the 1990s. The basis for the current methods of creep damage modeling are the classic
models for creep damage rates and rupture life prediction associated with Kachanov and
Rabotnov and reviewed in Reference 11. In the finite element models, conventional elastic
properties are modified in a similar way to the steady-state creep equations to take damage into
account. The first such comprehensive finite element model of a welded joint was by Hall and
Hayhurst [12]. This model used a basic CDM model for weld and parent material, which has
been used with variations and developments for all detailed finite element models intending to
represent weld and heat-affected zone (HAZ) properties and failure as accurately as possible. A
typical form of the model in terms of stress and strain invariants is as follows:

  A. exp(Q1 / RT ){ e /(1  D ) /  D}n


c
Eq. 3-1
.
D  B. exp(  Q 2 / RT ){ rupt /(1  D ) /  R} Eq. 3-2

 e  E (   c ) /(1  D) Eq. 3-3

 rupt   I  (  1) e Eq. 3-4

 rupt   e exp[C ( J 1 /  e  1)] Eq. 3-5

10371024 3-2
In these equations A, B, Q1, Q2, E, D, R, α, n, and σ are material property constants. T is
temperature. D is the damage parameter, which varies from D = 0 initially to D = 1 when the
material test sample has ruptured. σe and εc are von Mises effective stress and effective creep
strain invariants, respectively. J1 is the first stress invariant (3  hydrostatic stress). rupt is the
rupture stress. Multiaxiality and constraint are described in the equations for σrupt. Alternative
forms are a linear combination of effective stress and maximum principal stress (MPS) and an
exponential function of hydrostatic stress and effective stress. = 1 gives an MPS criterion in
the first case. In the second case, the constant C = 0.2 is typical and is consistent with the MPS
criterion over a limited range of multiaxial conditions.
It can be seen that the model consists of a traditional (Norton) equation for creep strain rate
modified by a quotient with a “damage” parameter D giving failure (infinite strain rate) in a
finite time. The damage rate is given by a similar equation, with a different exponent χ. The
elastic modulus E is similarly modified, so that elastic strain cannot support load when D = 1.
Typically, for a piping system, the use of CDM analysis would show how stresses relax over
time, whereas damage rates correspondingly decrease. When damage in some “hot spot” increase
to more than ~0.8, the analysis starts to cut back with smaller and smaller time steps and
becomes significantly slower. Special measures are required when damage = 1 to track
progression of damage through a section until predicted failure.
The API 579/ASME FFS document [2] uses the omega creep material model. This is a combined
deformation and damage model using a specific form for the creep curve defining the material
creep strain model as follows:
dεc/dt = secondary creep rate  exp(εc Ω).
Rupture life = 1/(secondary creep rate  Ω).
The material property data in Reference 2 define secondary creep rate and Ω for given stress and
temperature.
Multiaxiality is included in the formulation so that deformation, multiaxiality, and damage are
unified. Connecting multiaxiality, rupture, and strain accumulation is a significant difference
from traditional CDM models. The key to understanding low-ductility rupture in high-constraint
regions is the combination of low creep strain rate with high damage rate. This is not possible
with the multiaxial omega model.

3.4 Decoupled Creep and Damage Analysis


Full CDM models have been used since the early 1990s primarily for high-temperature weld
assessment. A number of authors noted that for realistic conditions of life, stress, and
temperature, weldment life is reasonably well characterized by the time to damage initiation.
Molinieux et al. [13] find that identifiable creep damage appears at a life fraction of 80% to 90%
in notched specimens. Similarly et al. [14] found that a life assessment based on steady-state
creep analysis predicts the failure location and 60% to 80% of weld life. Payten [15] and Hyde
et al. [16] use a conventional creep analysis and a decoupled damage calculation conservatively
to assess weld creep life. Similarly, Hyde et al. [17] used full CDM and steady-state analyses to
evaluate narrow gap and conventional pipe welds.

10371024 3-3
A decoupled model could typically be based on the following equations.

  A. exp(Q1 / RT ){ e /  D}n


c
Eq. 3-6

.
D  B. exp( Q / RT ){ rupt /  R}  Eq. 3-7

 e  E (   c ) Eq. 3-8

 rupt   I  (  1) e Eq. 3-9

The creep stress analysis depends on the von Mises stress, and rupture depends on multiaxial
stress, typically MPS. In practice, it is recommended to use α = 1 in Equation 3-9 due to the
uncertainty of less conservative options. Other models for strain rate such as the API 579
correlations may be used, with caution, because these polynomials may have unexpected
behavior outside certain stress limits. These limits are not practical because nonlinear analysis
may use stresses considerably different from those expected while iterating.
It can be seen that these equations are reasonable for the prediction of first significant creep
damage because they give a reasonable estimate of the evolving stress distribution up to the time
of first significant creep damage.
For the purposes of piping life prediction, both the omega and traditional decoupled methods
have been used in the analysis in this report and are both recommended if the stress analysis
software and data are available. The only caveat is that the multiaxial approach in the omega
analysis can lead to difficulties, which can be avoided by reverting to a conventional- (effective
stress) based creep stress analysis, with a decoupled multiaxial damage calculation based on the
omega rupture data.
For steady loading, a decoupled creep stress and damage analysis provides the best estimate of
creep life.

3.5 Simplified Analysis Methods


Simplified methods require less data and analysis capability than creep stress and damage
analysis. These are elastic-plastic limit analysis methods (reference stress methods) that have
been in the literature on high-temperature analysis stress and life analysis since the early 1970s,
[11, 18]. More recently, cyclic loading has been included in simplified methods in the form of
shakedown analysis. Limit and shakedown analyses for high-temperature life assessment are
used in the high-temperature nuclear assessment Code R5 [19].
As described in Reference 19, the limit load reference stress is defined by:
ref = Py/PL
where
P = load on structure
y = nominal or assumed yield stress
PL = limit load

10371024 3-4
An alternate definition for the reference stress is the smallest value of the yield stress for which
the structure does not collapse.
Reference stress methods were originally developed for prediction of strain and deformation. In
R5, a modification is used to account for ductility and multiaxial effects for calculation of creep
rupture stress σR, defined as follows:


 R   ref 1 
1
  1
 n 

where
ref = limit load reference stress
n = creep exponent
 = (max. elastic stress)/ref = limit load/load to first yield
For cyclic loading, the shakedown reference stress may be defined as the lowest yield stress for
which the structure shakes down, that is, behaves elastically over the cycle [19].
The use of a limit load reference stress for rupture prediction requires factors to account for
limited material ability to accommodate stress redistribution before rupture and for multiaxial
rupture. As noted, the R5 correction is intended to cover both. In Section 4, it will be seen that
for piping, a multiaxiality correction is required in addition to a potentially more conservative
version of the R5 correction. In Section 4, to explain the relationship between piping reference
stress and calculated creep rupture life, corrections for multiaxiality and strain accumulation are
found to be necessary.
The multiaxial rupture correction may be calculated as described in Section 3.6; however, this is
more likely to be helpful in a creep stress analysis than in a reference stress calculation. An
alternative is to define the MPS reference stress in terms of the MPS from the limit analysis
directly as follows:
MPSref = PMPS/PL
where
P = load on structure
MPS = value from limit analysis
PL = limit load
If MPS/y > 1.15, then the contour plot is misleading, and a factor of 1.15 (for pipes, elbows, and
shells) on the conventional reference should be used as a multiaxiality factor. Note that y is the
yield stress used in the limit analysis and should not be read from a contour plot.

10371024 3-5
3.6 Multiaxiality and Constraint
Cane [20] showed that high temperature creep rupture of tubes was driven by two mechanisms.
For relatively high stress and short life, rupture is defined by the von Mises stress, that is, the von
Mises stress correlates with uniaxial rupture data. For lower stress and longer life, there is a
transition to rupture defined by hoop stress or MPS. This is associated with the transition from
transgranular to intergranular creep failure. Equations such as 3-4 and 3-5 capture the possibility
of rupture due to a combination of von Mises (effective) stress and MPS, but do not capture the
possibility of a stress-based transition from ductile to brittle behavior.
High-temperature codes [2, 3] have different approaches to creep rupture under multiaxial
conditions, which can be shown to predict divergent and unlikely behavior under various stress
combinations. The Omega approach [2] uses a multiaxial correction that can be particularly
severe in triaxial conditions. It also does not capture the distinction between high-stress von
Mises and low-stress MPS rupture. It is connected to the theory of Rice and Tracey [21].
Resolving the differences between the MPS and Rice-Tracey multiaxial rupture approaches is
outside the scope of this report. Experiments do not provide a clear answer due to the difficulty
and time scales required for testing. It may be argued that the use of MPS is a reasonable
approach based on simple arguments of the local stress acting normal to grain boundaries, where
intergranular failure occurs. For combinations of tension and compression stress, the divergence
between MPS and von Mises criteria becomes very large, as shown in Figure 3-1. It is doubtful
that full credit should be taken for the MPS criterion for all biaxial stress states. For realistic
problems, any compression loading is probably associated with more severe tension loading in
the section. The maximum difference between the von Mises and MPS stresses in the tension-
tension quadrant is 15%. For typical design conditions, this means a difference in piping
minimum life of ~200%.
In this report, the MPS rupture criterion is used in the tension-tension stress quadrant where it is
the more conservative and the von Mises elsewhere. It may be argued that this is also a
reasonable triaxial approach, with the use of MPS in the tension-tension-tension region and the
von Mises stress elsewhere. For plain piping, triaxial stresses are not usually a factor. Weldments
are dealt with using a strength reduction factor. Triaxial stresses are encountered in heavy section
wyes and tees, which are typically dealt with separately from piping, as described in Section 8.
Figure 3-1 shows the von Mises and MPS limit surfaces. These define the multiaxiality factor R
that must be applied to the von Mises stress obtained from creep stress analysis to calculate
rupture life. R = 1.0 everywhere except the positive-positive quadrant. Within this quadrant, R
varies between 1 and 1.15.
The use of the multiaxiality factor with creep stress and limit load reference stress rupture
calculations is discussed in Section 4.5.

10371024 3-6
Figure 3-1
MPS and von Mises limit surfaces for piping and shell elements

3.7 Defect Assessment


Defect assessment for creep and fatigue crack growth is dealt with in API-579/ASME-FFS [2]
and is based on the failure assessment diagram (FAD). The FAD defines structural failure in the
presence of a crack and is defined by the ratios (L/Llim, K/K1c), where L = load, Llim = limit load,
K = crack tip stress intensity, and K1c = critical stress intensity. These quantities are defined to
provide safety factors that may be chosen to reflect particular levels of risk. Reference 2 contains
details of these calculations. It is clear that L/Llim = reference stress/yield stress and that a factor
of safety may be included in this definition. Other sources are available that provide more
straightforward calculations of the K term. In this report, the limit load is calculated in a limit
analysis. For a pipe, it is usually possible to calculate a simple and conservative reference stress
correction factor to account for the defect. Note that the collapse term L/Llim should include the
strain and multiaxiality correction factors required for creep rupture calculations.
The effect of time and cycles is defined by crack growth, which for high-energy piping is the
sum of creep and fatigue crack growth. These are also given in Reference 2. An economical
creep crack calculation method is given in References 20 and 21 as follows:
Creep crack growth rate is
da/dt = A/e0(C*)0.85
where
C* = ref   ref R

10371024 3-7

 = material strain rate as a function of reference stress

R = (K/ref)2
A ~100 for piping
e0 = creep ductility
Units in these equations are as follows:
da/dt = mm/h

 = strain/h
C* = MPa-m/h
R=m
References 20 and 21 contain recommendations for e0 and A.
Defect assessment is not the focus of this report. For defect assessment in piping, the only
additional question, compared with a pressure vessel, is the piping loads that must be considered
in addition to pressure. The analysis methods in this report give predictions for the evolution of
piping stress over time, which may be used in the creep crack growth rate calculation. The
fatigue crack growth depends on stress range, which is given by an elastic analysis.

10371024 3-8
4
PIPING ANALYSIS AND ASSESSMENT
4.1 Introduction
Piping analysis has a number of features that make it more complex than the analysis of
conventional pressure vessels, as shown in the following:
 Number and types of supports. Multiple constant and variable load hangers, rigid
hangers/supports, and snubbers are found on most piping systems.
 Even relatively low values of creep strain accumulation over time can lead to constant load
hanger malfunctions by reaching movement limits and bottoming or topping out.
 When hangers bottom out or otherwise malfunction, a significant change in piping loading
can occur.
 Cyclic loading may accelerate deflection and damage rates, particularly after such changes.

These features are difficult at best to deal with on the basis of elastic analysis. This is reflected
by industry experience with typical ongoing problems of malfunctioning hangers and the often
unknown consequences for piping integrity. To assess the risk of high-energy piping, the piping
analysis should be able to capture real piping behavior.

4.2 Differences Between Analysis for Design and Life Assessment


Piping Design. Design requirements are defined in terms of temperature-dependent, material-
allowable stresses for different categories of stress and loading. They address both steady and
cyclic loading. The design of a piping system is usually an iterative process, with a linear elastic
simplified (Caesar, commercially available software [http://www.coade.com/products/casesarii])
stress analyses performed for each trial design. The selection and positioning of hangers to
satisfy hot and cold design requirements usually require trial and error. Piping flexibility to
accommodate thermal expansion is assessed realistically.
Hanger Deflections. A well-designed (ideal) piping system would have a life equal to that of the
plain pipe under pressure. In this case, axial stresses (corrected for the girth weld strength) due to
self-weight, pressure, and other sustained loading would not be greater than the hoop stress due
to pressure. The problem with constant load hanger deflections arises because even for such a
well-designed system, satisfying allowable stresses does not prevent piping distortion, leading to
hanger positions outside the original specifications.
Elastic and Inelastic Analysis. Piping design is based on elastic analysis [1], and it is well-
known that prediction of piping deflection and stress as a result of creep requires more general
analysis techniques. One of the objectives of this report is to investigate both creep and a range
of limit and simplified elastic-plastic analysis techniques for this purpose. These simplified
elastic-plastic methods have significant advantages in terms of data requirements and ease of
analysis.

10371024 4-1
Design and Operating Conditions. For a life assessment, there is a choice between use of
design or operating data. There is a considerable benefit to having a sound basis for a realistic set
of operating conditions. This will typically consist of combinations of pressure and temperature,
each with a time fraction. This is the main contribution to risk assessment.

4.3 Analysis for Risk and Life Assessment


The objective of analysis in this project is to determine and predict failure based on a specific set
of data. We attempt to define analysis methods to address the following:
 Girth weld cracking
 Excessive deflections leading to hanger problems
 Reduction in piping life due to cyclic loading
 Use of a risk-based approach for ageing piping

The focus is on efficient analysis methods and data to predict failure and assess risk.
Analysis Objectives. A detailed structural analysis seeks to determine the changes in stress and
deflection over time and accumulation of creep strain. From the stress history, a creep damage
calculation can be performed. Excessive strains and deflections are forms of failure, and
particular attention must be paid to localized inelastic strain accumulation and follow-up.
Follow-up refers to the concentration and accumulation of inelastic strain at points of weakness
and flexibility. It is a particular possibility in piping due to long, straight sections causing
inelastic strain in bends. Indeed, Balaschak and Wray [4] identified an elbow in grade 11, which
accumulated an estimated 5% strain and ruptured after the assessment was performed. An
example of this is given in Appendix A.
Cyclic Loading. The effect of cyclic loading on creep damage accumulation needs to be
checked, and this is most efficiently done with a shakedown analysis technique. It would not be
expected that piping operating under as-designed conditions would have cyclic effects. Cyclic
effects are more likely to be associated with faulty conditions such as bottomed out or stalled
hangers. It appears that where use of constant load hangers is most likely to lead to excessive
deflections and hanger problems (in the middle of long horizontal sections and away from
downcomers) that there may not be serious consequences in terms of stress and creep fatigue.
Hanger locking in horizontal runs near vertical sections is more likely to lead to such problems.
When analysis shows that cyclic loading is causing creep damage rates higher than the steady-
state damage rates due to such problems, then urgent remedial action is probably necessary.
Analysis and Data Options. The life assessment methods for steady and cyclic loading include
reference stress techniques that require only rupture data for failure prediction. Because the
resources available for assessment (data and analysis software) are likely to be very variable, as
many options as possible are described in this report. The methods in this report are as follows:
 Time-dependent creep and damage analysis
 Inelastic stress analysis based on isochronous stress-strain data
 Reference stress limit analysis for life assessment
 Shakedown analysis for cyclic loading and creep fatigue

10371024 4-2
Life assessment will require a number of analyses, typically the following:
 Elastic analysis for design temperature and pressure
 Calculation of “follow-up” strains resulting from redistribution of thermal stress and effect of
“cold pull” if any
 Analysis of creep deflections
 Life assessment based on design and/or operating history if it is available
 Cyclic analysis reflecting any nonfunctioning hangers or restraints to determine whether
cyclic loading affects high-temperature life

4.4 Sources of Data for Stress Analysis and Life Assessment


Typical high-energy piping materials are the following:
 Grade 11
 Grade 22
 Grade 91

The sources of creep and rupture data for typical high-energy piping materials are primary and
secondary creep (ASME III NH [3]) and secondary and tertiary creep (ASME-FFS/API-579 [2]).
Mean and minimum rupture data included Larsen Miller (API 530 and ASME-FFS/API-579) and
Omega data (ASME-FFS/API-579).
The different analysis methods require data in different forms. For a creep stress analysis, creep
data are required in the form of strain rate equations as functions of stress, temperature, and
possibly strain or time. Elastic-plastic isochronous stress analysis uses isochronous data from
Reference 3. This and the limit analysis method are required for life assessment without creep
stress analysis.

4.5 Multiaxiality
B31.1 piping design is based on acceptable hoop stress and axial stress. For piping, two-
dimensional stresses are adequate, which is reflected in piping, elbow, and shell finite elements.
As discussed in Section 3.6, a combination of MPS and von Mises stress are used to define the
multiaxiality correction factor R.
For thick sections with penetrations and welds where hydrostatic stress and MPS can be greatest
within a section (triaxial constraint), care must be taken to find the critical elements and obtain
representative or MPS values. For typical piping, triaxial constraint is only a potential issue in
heavy section wyes and tees (see Section 8).
In Section 3.6, it was stated that for piping, the multiaxiality correction factor R is between 1 and
1.15, depending on the combinations of axial and hoop stress. This may be used directly with
creep stress analysis. For creep stress analysis, the factor may be programmed into the user
routine calculating creep strain and creep damage. The parameters available in the Abaqus creep
user routine are von Mises stress (Q) and pressure stress (P) = −σ(direct stresses)/3.

10371024 4-3
For piping, elbow and shell (biaxial) elements, the factor R is accurately represented by the
function:
R = 1 + {√(4/3) − 1}SIN{p [(3(−P/Q − 1/3)]2.24} if −P/Q > 1/3
R = 1 otherwise.
This expression is used in the creep subroutine to the rupture stress for creep damage. For a
reference stress calculation, it is straightforward to calculate a reference MPS. A strain correction
factor should then be applied to define the reference rupture stress R. This may be used in a
creep subroutine to calculate creep damage.

4.6 Weldments
Weld metal and HAZs are usually the sources of first and most severe creep and creep-fatigue
damage to the pressure boundary. Weld design factors as in the B31.1 Code [1] are intended to
reflect this, but the use of strength factors (E = 1) based on inspection, for high-temperature
piping, is not generally realistic. Guidance on weld strength reduction factors (WSRFs) is
available from EPRI and recent editions of the B31.1 Code.
Without explicit representation of seam and girth weldments in a piping model, distinguishing
between them is not really possible. Therefore, a life assessment of seamless pipe is not possible,
without explicitly including girth welds in the model. This is certainly possible, and it is only
necessary to model a few critical girth welds, which may be readily identified. This would be
done in a shell model by having a ring of weak elements the width of the weldment, with
reduced properties.
An assessment with a homogeneous material would have to be based on either nominal
properties, which could be nonconservative, or on a weakened material, which represents and
implies both girth and seam welds.
For seamwelded piping, if the piping life obtained from a homogeneous model is significantly
less than that of a plain seamwelded pipe under pressure, then a girth weld rather than a
seamweld is likely to be the location of the damage.

4.7 Peaking and Out-of-Round Piping


A seam-welded pipe with a peak due to difficulties with achieving a constant radius all the way
to the welded joint will add to the nominal pressure stress, exactly in the weld material and in
the weak HAZ. If this situation is encountered, it is very likely that stress will be higher than
the nominal (circular) pipe value, and creep life considerably less than the nominal minimum
design life.

10371024 4-4
4.8 Aspects of a Piping Model
Typical structural components in a high-energy piping system are the following:
 Straight and elbow or sweep sections
 Heavy section wyes and tees at piping intersections
 Restraints and hangers, rigid constraints, snubbers

Scope of Piping Model. Although wall thickness is usually nearly optimally chosen for straight
and curved sections, wyes and tees are usually conservatively designed by area replacement
methods, resulting in stiff, strong components. Their likely failure modes are creep fatigue and
cracking due to thermal shock rather than creep damage due to pressure and the piping system
loads. Unless there is a particular reason to include them in detail in the piping model, such
components are best assessed on their own with relevant pressure and thermal loading. They may
be represented by a point in the piping model, where three converging lines are connected.
Element Types. The complexity of piping systems makes the use of line (beam or elbow)
elements an appealing analysis option. It is well-known that pipe elements (beam elements with
an independent pressure calculation) cannot accurately model piping sweeps, elbows, and bends.
Elbow elements are line elements derived from shell theory that do claim to represent bends
realistically. They have variables allowing selection of deformation modes and integration
points. Abaqus has default values used in the example in this report. One of the objectives is to
compare the use of traditional pipe elements, elbow elements, and shell elements for piping
analysis. Second-order pipe, elbow, and shell elements were used in the examples in this report.
The basic options for piping models are as follows:
 Line element model with elbow elements.

Each hanger, support, or restraint is connected to a single node.


 Intersections consist of three line elements connected at a single node. This may require
some care not to create out-of-balance pressure forces. It may be necessary to apply a
balancing force. This can be checked with a simple model representing the intersection of
three lines and a check on out-of-balance forces.

Shell elements for straight and curved sections.


 Hangers, supports, and restraints may be connected to a lug or bracket on the shell or to a
node on the pipe centerline. Unless the hanger is attached to a lug welded to the pipe and
there is a need to understand the local strength of the pipe shell and attachment, the use of a
node or reference point on the centerline is recommended. If the possibility of distortion of
the pipe wall due to a trunion exists, the trunion load may be applied over an area of the
shell.

10371024 4-5
Intersections (wyes and tees) may be represented by the following:
 A local accurate detail with solid or shell elements

or
 A common node or rigid connections between the centerline nodes at the ends of the pipe
shell sections. Care must be taken to account for pressure loads correctly at these joints. This
is usually a more efficient approach, unless there is a specific reason to include the detailed
intersection in the piping model. As noted previously, the intersection may require a separate
assessment. The main steam example in Appendix A has a junction or header connecting the
main pipe to a number of leads. This is will almost certainly be reinforced. If it is, then a 3-
dimensional solid model of the part is necessary for the stress analysis and life assessment of
the part. In this case, the solid model should be included in the system model, so that piping
loads coming in to the part are accurately modeled. In such a case, it will be important to use
a mesh density that is carefully chosen to be economical and adequate for the stress analysis.
This may involve trying out different mesh densities on the solid model on its own before an
analysis of the system.
- If elbow elements are used, it is usually satisfactory to ignore an intersection such as a
wye and simply connect the elbow elements at a common point.
- Hangers and restraints will be connected to a single node. Snubbers would not have any
role in a static model because they only generate forces with dynamic loading.
The three dominant types of hanger are easily modeled. The model will generally have hangers
and supports attached to nodes on the pipe centerline.
 Constant load hangers are represented by a constant point load. They will generally ramp
up during the first analysis step.
 Rigid hangers are represented by a fixed displacement.
 Variable hangers with a preload (cold load) are represented by a spring to ground with the
hanger stiffness and a preload. The preload is applied during the first analysis step. This is a
simple operation in a standard analysis package.

4.9 Identification of Hanger Problems


In addition to any general observations about hanger condition, distortion, or damage, detailed
data on hanger positions under hot and cold conditions are necessary for the assessment. It has
been mentioned that hanger positions are likely to change over time. Tracking these trends based
on yearly hot and cold walkdowns is recommended. Photographs of the hanger showing the
position indicator are the best form of data to collect. The position indicator is not always a direct
measure (in inches [millimeters], for example) of deflection. The hanger catalogue may therefore
be required to translate a pointer position into deflection. Deflections in inches or in millimeters
are required for assessment.
Changes in constant load hanger positions over time are most likely due (a) the hanger load
changing or hanger movement restricted or (b) the hanger load being wrong for creep conditions.
As noted, it is difficult to design against creep deflections with constant load hangers based on
elastic analysis.

10371024 4-6
A steady change in hanger deflection over time with the hot-cold deflection range nearly
constant would indicate (b). A significant change in deflection range could mean a change in
hanger spring stiffness due to corrosion and damage or excessive compression leading to
contacting spring coils. A significant change in position could mean a change in hanger preload.
Misalignment of hangers, support structure, and piping will generally increase hanger load, and
in constant load, hangers will probably lead to bottoming out.
Adjacent hangers could be affected by such changes, so although it is likely that a change in
hanger performance would be most obvious at the hanger in question, it should be remembered
that the piping system is more complicated than the individual elements.
A change in average hanger position without bottoming or topping out or significantly changing
the displacement range does not necessarily require any action or response.
A hanger that has a displacement range within about 80% to 120% of the design value and does
not show any signs of distress probably does not need immediate attention.
Monitoring of hanger load changes with strain gauges or other instrumentation will allow the
identification of hanger problems in real time.
A shakedown analysis of the line with a locked hanger as described in Section 4.11.5 will show
whether a significant life reduction is expected.

4.10 Typical Data Requirements for Piping Assessment


There is a core set of requirements without which assessment is not viable. Beyond that, the
value of the assessment increases with more information.
The core requirements are as follows:
 Piping general arrangement and detailed drawings defining geometry and materials of the
piping system
 Hangers and supports with constant loads, preloads, stiffness, and travel limits
 Design and operating conditions, operating hours
 Inspection and assessment reports
 History of modifications and repairs

4.11 Piping Model Element Comparison


We compare the behavior of these three element types for piping analysis.

4.11.1 Conventional Beam Elements with Independent Pressure Calculations


These are line elements and represent a segment of pipe as a beam in space with length L,
section area A, second moment of area I, and polar moment of area J defining the geometry.
These quantities with the material elastic modulus E and Poisson’s ratio n define the axial,
bending, and torsional stiffness of the element. As noted in the following, the relationship

10371024 4-7
between bending moment and stress in this approach is not correct for bends. In this case,
flexibility and stress factors for bends are provided in the design code [2].
The stiffness properties of the element are combined with those of other elements to define the
piping system structural stiffness matrix. The effects of gravity and thermal loads are dealt with
in this structural system, whereas pressure stresses are calculated on a pointwise (local) basis.
Hoop pressure stress = pr/t, axial stress = pr/2t, where p = pressure and r = mean wall radius.
This forms the basis of typical piping design calculations.
There are difficulties with using beam elements to represent a piping system. The equilibrium
bending stress distribution in a simple beam or straight pipe element cannot exist in a pipe bend.
The double curvature of the pipe bend wall means that bending through the pipe wall is
developed in a bend, leading to generally higher stress than in a straight section under the same
pressure and system loads. Shell elements (or solid elements for very high pressure and heavy
sections) are required to model this effect. The B31.1 [1] design codes provides flexibility and
stress factors for pipe bends in beam element modes to deal with this problem. It is standard
practice to use beam elements to represent a piping system for design, and it is believed that this
produces adequate designs or rather that the problems with current designs are not due to the use
of pipe elements. However, the use of strength and flexibility factors for these elements for
design makes them suspect for the nonlinear analysis used for life assessment.

4.11.2 Shell Elements


Shell elements with shear (so-called thick shell elements) in a pipe model are the best generally
available and efficient method of analysis. They provide direct and bending stress. The ability to
generate realistic contour plots of stresses, strains, and displacements is helpful. Pressure loading
in a straight section of standard shell elements gives direct hoop stress equal to the mean radius
formula s = pr/t, where r = mean radius and t = thickness. For predicting yielding failure, this is
accurate to less than 2% for r/t > 2. Around the pipe circumference, 12 second-order elements
appear to be necessary to avoid warnings about aspect ratio and curvature and high discontinuity
stresses. The element axial dimension can be a significant multiple of the hoop dimension. For
straight sections, long, high aspect ratio elements may be used provided that standard mesh
checks are performed. For sweeps and bends, 10 second-order elements per 90° are
recommended. For the reference stress analyses used in this report, coarser meshes appear to be
acceptable. In the example, eight elements around the circumference and around each 90° bend
gave reasonable results. If mesh density can be easily changed, it is recommended to start with a
coarse mesh and refine it until the results are mesh insensitive.

10371024 4-8
4.11.3 Elbow Elements
These are really high-order shell elements, but are handled for purposes of input data as line
elements. They therefore model pipe bends correctly and have the advantage of ease of
assembly, modeling supports, and use compared with conventional shell elements. They generate
axial pressure stress automatically. These are available in Reference 26. Due to the efficiency of
these elements, there is usually no reason to compromise accuracy with a coarse mesh. As with
shells, it is usually straightforward to begin with a coarse mesh and refine it until the analysis
results show no further improvement.
Due to a recent change in the Abaqus elbow elements [24], there is no longer the option to
choose between “open” and “closed” elements. This means elbow and shell elements cannot be
easily combined to give the best of both worlds, namely, an economic model with straight
sections easily able to locate supports and accurate models of bends. This strengthens the
arguments for shell models.

4.12 Analysis Methods


Requirements for the following inelastic analysis methods are discussed:
 CDM: creep analysis with damage and failure prediction
 Fully coupled creep damage mechanics approaches
 Decoupled creep stress analysis and creep damage analysis
 Elastic-plastic analysis using isochronous creep data, with cold pull if applicable
 Limit analyses with cold pull if applicable
 Shakedown analysis with cyclic pressure and thermal loading

For life assessment, elastic analysis is helpful as a preliminary analysis and as a baseline for
other analyses. Elastic analysis is widely used to check hanger loads and deflections and to make
recommendations for hanger modifications. This is often performed by imposing observed
displacement cycles on hanger positions to calculate reactions. Redesign of hangers to replace
constant load with other designs to avoid continual bottoming or topping out is also performed.
Due to creep, it is is often difficult or impossible to match field observations with analysis. Its
use for life and risk assessment is very limited.
Minimum Use of Contour Plots
The assessment methods in this section are intended, as far as possible, not to depend on
interpretation of stress contour plots. Use of contour plot stresses for a subsequent life
calculation should be avoided wherever possible. There are often variability and a lack of rigor in
interpretation, which gets considerably magnified in life calculations. Choices of the critical
elements and the number of them to be assessed to characterize the piping system are judgments
that we would like to avoid if possible.

10371024 4-9
4.12.1 Fully Coupled CDM Modeling
This method is included for completeness, but is unlikely to be a practical tool for piping
assessment. It requires a full UMAT user routine, which can take a significant effort to develop
and use. The data requirements that have to be coded into the UMAT are the following:
 Density- and temperature-dependent modulus and expansion properties
 Temperature-dependent creep and yield properties
 Minimum rupture data

Further information on full CDM finite element models can be found in the documentation for
nonlinear analysis software [22].
Details of Omega modelling and methods can be found in ASME-FFS/API-579 [2]. It can be
implemented in a creep user routine, which is easier to use than a UMAT. The material rupture
data are sufficiently different from the Larsen-Miller rupture data, so be aware of both options in
life assessment, in addition to the difference between mean and minimum data.

4.12.2 Decoupled Stress and Damage Analysis


As note previously, this approach greatly simplifies the coding and analysis effort and is
reasonably accurate and conservative. The following data requirements are the same as for CDM
models:
 Density and temperature dependent modulus and expansion properties
 Temperature-dependent creep and yield properties
 Mean and minimum rupture data

If the ability to perform creep and decoupled damage analysis is available, then this method is
not difficult to perform and provides accurate but conservative answers without the need for
further post-processing.
The damage calculation in this method should be based on MPS, which is calculated as described
in Section 4.5. The strain factor is not required because the interaction between stress
redistribution and damage accumulation is dealt with in the analysis. A weld factor is required.
This may be applied to a weld region if it is modeled explicitly in a shell model. Otherwise, it is
applied to the pipe material as a whole. Clearly, the life predicted on this basis refers to a
weldment in the region of the maximum damage/minimum life. If the life is significantly less
than that of a plain pipe under pressure, then a girth weld rather than a seam weld is likely to be
the location of the damage.

4.12.3 Isochronous Elastic-Plastic Analysis


This is straightforward. It requires the data in the ASME III NH [3] isochronous curve for a
specific time (say 105 hours) to be represented as strain hardening yielding data. Tables
containing these data for grade 22 are given in Appendix A. Use of nonlinear geometry options
are not expected to be necessary. The results are predictions of deflection, stress, and strain. The
strain is required for reference stress strain correction factor.

10371024 4-10
Perform isochronous analyses for a range of times to estimate strain and deflection at a given
time.
If cold pull or variable hangers with preload are used, then the first step will contain these
effects. The second step will contain all other loads including thermal loads.

4.12.4 Steady Load Life Calculation: Limit Analysis


Strain, MPS, and Weld Factors. The following procedure is recommended to obtain reference
stresses that may be used in spreadsheets for life and risk calculations. Some background on the
use of the limit analysis for creep problems was mentioned previously. Its use for damage and
rupture prediction is based on tests, numerical simulations, and limited experience. As discussed
previously, creep life is a competition between stress redistribution and damage accumulation. A
correction factor that reflects the strain required for redistribution must be calculated in addition
to the correction for MPS. The effect of weldments may be addressed by a third factor, the
WSRF, in the life calculations. This method does not distinguish between girth welds and seam
welds. For seamless piping, if the piping system reference stress is greater than the plain pipe
pressure reference stress, then the resulting life prediction will correctly reflect the girth welds.
Otherwise, it will reflect a seam-welded pipe.
The following reference stress calculation allows the effect of cold pull to be determined:
1. Define the design or operating conditions as appropriate.
2. The mechanical properties required are density, a temperature-dependent modulus, and
expansion coefficients.
3. Define a single yield stress such that the limit analysis will not be able to complete. (Slightly
less than the maximum hoop stress is recommended.) Boundary conditions include any fixed
or rigid supports.
4. Load step 1. The first step is used to apply thermal stress, any cold pull, and preload on
variable hangers or restraints.
5. Load step 2. All other loading (pressure, gravity, constant hanger loads) ramp over time from
0 to 1.
6. Perform the analysis without the nonlinear geometry option, with and without yield stress,
with a moderate limiting time increment, say 10−5.
7. The yielding analysis will fail to converge before it reaches the end of the second time step.

The reference von Mises or effective stressref = yield stress/maximum step time.
Calculate rupture stress:
R = MPSref  strain factor  weld factor

The strain factor is a version of the R5 factor in Section 3.2. It was found when analyzing the
examples in Section 4.12 that this factor was unconservative compared with detailed creep stress
analysis and needs to reflect the Monkman-Grant strain (1/Omega). The physical meaning of it is

10371024 4-11
that the stress redistribution toward the reference stress needs to happen fast enough or creep
damage will appear before it is complete or nearly complete. The Monkman-Grant (or Omega)
material parameter is the key property in this respect. In the case of the example in Section 4.12,
Omega = ~80, and so Monkman-Grant strain = ~1.3%. The plastic strains in the limit analysis
are greater than 100% by the time the analysis fails to converge, and it is expected that this will
be typical of piping systems. We therefore have to consider the steps earlier in the limit analysis,
when maximum plastic strain is comparable with the piping inelastic strain as calculated in the
isochronous stress analysis in Section 4.11.3. For any given increment in the limit analysis with
its associated plastic strain and load, the strain correction factor is:

Strain factor = 1 + [1 − exp(−const.p)] ( − 1)]

where const. ~0.5, =limit load/yield load.


Basing the calculation on a reference MPS avoids the need for a multiaxiality factor. A weld
factor is also generally necessary.
The strain factor in the example in Table 4-3 of Section 4.12 is 1.17 compared with the R5
correction for this case of 1.08.
The value of the constant provides the calibration with the creep stress and life analysis in
Section 4.13. It is notable that the value that makes the reference stress prediction consistent with
the creep analysis (when p is approximately equal to ) is equivalent to the R5 criterion with a
value of n such that 2 < n < creep exponent.

4.12.5 Cyclic Load Creep Life Calculation: Shakedown Analysis


The objective of cyclic analysis is to determine whether cyclic loading reduces life compared
with steady loading and, if so, to provide an estimate of reduced life.
There are two different aspects to the thermal cyclic problem: overall temperature cycle of the
piping system and temperature rate-dependent thermal shock, which affects local thick sections
most severely.
In both cases, spreadsheet calculations may be used to give an estimate of the effect on creep
life. The approach is to compare the elastic stress range with an allowable defined for a selected
rupture life as follows:
Allowable stress range = cold yield stress + rupture stress.
For creep-fatigue calculations, this should be based on minimum data.
As a first step, select the rupture life as the constant load life, calculated using one of the
methods described in this report. If the elastic stress range everywhere is clearly less than this
limit, then cyclic loading is not a factor in the piping life assesssment. If it is close to or more
than the limit, then detailed finite element methods are recommended as follows.
As with the limit analysis method, shakedown analysis finds the stress history and distribution,
which is the most efficient for maximizing creep life. The method uses a fictitious temperature-
dependent yield property defined as the lesser of the yield stress and stress to rupture in a given

10371024 4-12
time. The analysis represents changes in temperature and loading over the cycle accurately. If the
structure shakes down, that is, behaves elastically after a number of cycles, then it is possible for
cyclic stresses to exist, which will allow creep life equal to the rupture time defining the
fictitious yield curve. If this time is more than about 90% of the steady load life prediction, then
cyclic loading does not significantly affect piping life.
Abaqus has a direct cyclic routine that allows cycles to be analyzed in a single step. If this is not
available, then it is straightforward to run five or six steps to check for shakedown. If shakedown
has not occurred after this number, then it is unlikely to occur in more steps. Trial and error is
required to find a reasonably narrow rupture time window, above which shakedown does not
occur and below which it does occur.
Typically, cyclic analysis has the following steps:
1. The first step applies any cold pull, gravity loads, and constant hanger loads. Cold pull does
not change shakedown one way or another, but it could make the analysis quicker. A
malfunctioning constant load hanger may be represented by a fixed displacement or a defined
displacement cycle.
2. The pressure and temperature cyclic loads and any known displacement cycles are applied in
the second and subsequent steps. The cyclic steps are over an arbitrary time from 0 to 1 and
have pressure and temperature varying in phase.
3. Transient thermal stress has to be represented in real time and in a cycle.

This method of cyclic analysis requires an assumed life to define the fictitious yield property.
Figure 4-1 is an example.
The fictitious temperature-dependent yield property used in the shakedown analysis is the lesser
of yield and rupture stress. It is recommended that the first analysis is based on a short time, say
10,000 hours, to make shakedown, and therefore an easy analysis, likely. Working up to the
minimum life for steady loading as calculated previously is straightforward.
It would be unlikely that cyclic loading has any effect on steady load life for piping, unless there
are, for example, serious hanger problems.

10371024 4-13
Figure 4-1
Yield and minimum rupture data for grade 22 material

4.13 Piping Example


An example will be used to evaluate the life assessment methods. Comparisons will be made
among the following:
 Shell, elbow, and pipe elements
 Elastic, limit, and creep analysis for life estimates
 Use of minimum and mean rupture data

The analyses will demonstrate the effect of cold pull and the effect of cycling and hanger
malfunctions.

10371024 4-14
Consider the piping system shown in Figure 4-2. Table 4-1 summarizes the data.

Figure 4-2
Shell elastic analysis

Bend radii are 150 in. (3810 mm), vertical leg is 2500 in. (63,500 mm), and horizontal legs
are 5000 in. (127,000 mm) and 2500 in. (63,500 mm), respectively. Constant load supports,
indicated in Table 4-2, are numbered from downcomer section. Maximum contour von Mises
stress = 8.1 ksi (55,847.534.1 Pa) and maximum integration point stress = 7.7 ksi
(53,089,631.2 Pa). Only one hanger reduces creep life if it is locked.

Table 4-1
Example piping data

Material Temperature Pressure Outer Inner t Density


(°F/°C) Diameter Diameter
1000/ 1.8 ksi/ 30 in./ 20 in./ 0.00028/
Grade 22 537.8 12,410,563.1 Pa 762 mm 508 mm 5 0.007.7 kg/cm3

10371024 4-15
Table 4-2
Constant load hangers

Constant load hangers Load (kips)


1 151
2 128.5
3 133.8
4 103.5
5 62.3
6 74.4
7 61
8 106.3
9 125.2
10 105.1

1 kip = 453.6 kilogram-force.

The piping is fully anchored at the ends and is constrained at the bottom of the straight section of
the downcomer to have zero horizontal deflection.

4.13.1 Elastic Analysis


Figure 4-2 shows elastic stress contours and a schematic indication of positions of constant load
hangers. The loads were calculated by restraining the support points vertically and obtaining
reactions. These were then used as constant loads at the support points.
Figure 4-3 shows the elastic deflected shape from pressure and self-weight with a maximum
deflection of 7 in. (177.8 mm) vertically.

10371024 4-16
Figure 4-3
Elastic vertical deflection = 5 in. (127 mm)

4.13.2 Prediction of Creep Deflection


The generally available sources of creep data to predict deflections are API-579/ASME-FFS
Standard [2] and the ASME Section III Subsection NH nuclear code [3]. The former has had
primary creep eliminated, and the latter has data in the form of isochronous curves up to creep
strains of 3% for grade 22 material. The piping deflection in Figure 4-4 was obtained using
isochronous data for grade 22 at 1000°F (537.8°C). The analysis was a conventional time-
independent elastic-plastic analysis with conditions from Table 4-1. The data are in the form of a
hardening yield stress-strain curve shown in Figure 4-5, and given in Appendix A. Figure 4-6
shows corresponding creep strain contours. It may be seen that large deflections can result from
relatively minor creep strains. This shows the vulnerability of constant load hangers used over
long horizontal sections.

10371024 4-17
Figure 4-4
Shell creep deflection after minimum life (1.1e6 hours) = 48 in. (1219.2 mm)

Figure 4-5
105-hour isochronous stress-inelastic strain data for grade 22 at 1000°F (537.8°C)

10371024 4-18
Figure 4-6
Creep strain prediction (0.35%) after 1  105 hours using isochronous stress-strain data for grade
22 at 1000°F (537.8°C) (The deflection in Figure 4-2 does not require large creep strains.)

4.13.3 Prediction of Creep Rupture and Girth Weld Cracking


There is a limited set of possible causes of piping creep failure: failure of plain pipe or seam-
welded pipe by pressure hoop stress and failure of girth weldments driven by piping system
loads, pressure, and the reduced creep strength of the girth weldment.
To predict changes in piping stresses due to creep deformation over time requires one of the
following:
 Creep stress analysis
 Elastic-plastic analysis using isochronous stress-strain data (Figure 4-3)
 Limit (reference stress) analysis

Option 1 is limited by the difficulty of obtaining relevant comprehensive creep data in a form
that is reasonably straightforward to use in finite element stress analysis. To overcome this
would require an amalgamation of the primary and secondary creep data in Reference 3 and the
secondary and tertiary creep data in Reference 2, in a form that is efficient for stress analysis.
The systems of polynomial expressions used for creep data in Reference 2 can have significant
difficulties when used in a general finite element program.

10371024 4-19
Options 2 and 3 are so-called simplified methods that are more straightforward to execute and
should not give significantly different or unconservative answers. The isochronous analysis is
straightforward and requires little or no interpretation. However, it will require iteration of the
time for the isochronous data if time to achieve some strain or deflection limit is required. The
limit analysis method is the most efficient in that estimates of piping life may be made with one
time-independent analysis.
The limit analysis is most conveniently performed with realistic loading and a fictitious (low)
yield stress (sy) so that the analysis will fail to converge at some load (time) factor t < 1. The
piping reference (von Mises) stress σref is then:
Sy
 ref 
t

The limit analysis can develop considerable plastic strain before it reaches the limit load and fails
to converge. A reference stress σref may be defined for any value of t up to the maximum value
before the analysis terminates. The maximum plastic strain may also be determined as a function
of t. Therefore, the reference stress solution may be defined as a function of maximum plastic
strain σref (  p ). A plot of reference stress against plastic strain allows a value of σref to be
identified, which characterizes piping creep life.

4.13.4 Cold Pull


Cold pull is a controversial topic. The benefits of cold pull if they exist, will be seen from the
limit analysis and reference stress calculations with and without cold pull. If there is a difference
in the ratio (limit load/yield load) for the two cases, then there will be difference in rupture life.

4.13.5 Summary of Constant Load Analyses


Table 4-3 is a summary of results from analysis methods and model types. The shell and elbow
element results are in good agreement, except for the creep analysis, where the pipe elements did
not work. The shell, elbow, and pipe reference stress and life prediction results are identical. The
deflections show that the isochronous data are more conservative than the creep data. (This is
probably due to the existence of primary creep in the isochronous data.) The pipe and elbow
elements deflections are about 50% higher than the shell deflections.

10371024 4-20
Table 4-3
Summary of analyses: example piping model

Elastic
Reference Stress Limit Analysis
Analysis
Maximum Maximum
von
von Mises Principal Deflection Principal
Mises Mean life Minimum life
(ksi) Stress (in.) Stress
(ksi)
(ksi) (ksi)
Pipe 5.5 6.1 −6.7 5.2 6.0 4.17E+06 9.91E+05
Elbow 7.7 8.5 −5.8 5.2 6.0 4.33E+06 1.03E+06
Shell 8.1 9.4 −5.1 5.2 6.0 4.47E+06 1.07E+06
1e5 Hour Isochronous Creep Analysis
Analysis 1e5 Hours
Maximum Maximum
Von
von Mises Principal Deflection Principal
Mises Deflection Minimum life
(ksi) Stress (in.) Stress
(ksi)
(ksi) (ksi)
Pipe 5.6 6.3 −76 None None None 1.30E+06
Elbow 6.7 7.3 −70 5.9 6.5 −19 1.10E+06
Shell 7 8.1 −48 6.5 7.5 −7.0 1.20E+06

Mean life
Pipe 5.80E+06
Elbow 4.10E+06
Shell 4.40E+06

4.13.6 Cyclic Analysis


The effect of vertically locked hangers was checked using the shakedown analysis method
described in Section 4.11.5. It was found that the only locked hanger that reduced creep life was
hanger 1 in Table 4-2, the closest to the vertical leg. In this case, it was found that cyclic life was
greater than 100,000 hours. Further trial and error would probably show a less conservative
value.

10371024 4-21
10371024
5
HOT AND COLD WALKDOWN AND INSPECTION
In addition to any general observations about hanger condition, distortion, or damage, detailed
data on hanger positions under hot and cold conditions are necessary for the assessment. It has
been mentioned that hanger positions are likely to change over time. Tracking these trends based
on yearly hot and cold walkdowns is recommended. Photographs of the hanger showing the
position indicator are the best form of data to collect. The position indicator is not always a direct
measure (in inches [millimeters], for example) of deflection. The hanger catalogue may therefore
be required to translate a pointer position into deflection. Deflections in inches or millimeters are
required for assessment.
Changes in constant load hanger positions over time are most likely due to (a) the hanger load
changing or hanger movement restricted or (b) the hanger load being wrong for creep conditions.
As noted, it is difficult to design against creep deflections with constant load hangers based on
elastic analysis.
A steady change in hanger deflection over time with the hot-cold deflection range nearly
constant would indicate (b). A significant change in deflection range could mean a change in
hanger spring stiffness due to corrosion and damage or excessive compression leading to
contacting spring coils. A significant change in position could mean a change in hanger preload.
Misalignment of hangers, support structure, and piping will generally increase hanger load, and
in constant load, hangers will probably lead to bottoming out.
Adjacent hangers could be affected by such changes, so although it is likely that a change in
hanger performance would be most obvious at the hanger in question, it should be remembered
that the piping system is more complicated than the individual elements.
A change in average hanger position without bottoming or topping out or significantly changing
the displacement range does not necessarily require any action or response.
A hanger that has a displacement range within about 80% to 120% of the design value and does
not show any signs of distress probably does not need immediate attention.
A shakedown analysis of the line with a locked hanger as described in Section 4.11.5 will show
whether a significant life reduction is expected.

10371024 5-1
10371024
6
DESIGN AND OPERATING CONDITIONS
When operating histories of steam pressure and temperature are known, considerable benefit
(usually in terms of reduced conservatism and increased accuracy) can follow for life
assessment. Advantage can also be taken of changes in conditions, so that times at different
combinations of temperature and pressure can be factored in to the life usage calculation.
Design practices have a significant effect on the problems likely to be encountered over the
expected life of the piping. In particular, the use of constant load hangers for high-energy piping
is almost certain to lead to excessive deflections in the medium to long term. The simple
example in Section 4 showed how complying with code-allowable stress was no guarantee that
deflection would remain within expected limits. The use of variable and rigid hangers, where
possible, should be explored. This would address a significant problem caused by the belief that
constant load hangers are necessary.

10371024 6-1
10371024
7
ASSESSMENT AND RISK ANALYSIS
For high-temperature pressure parts, the estimate of remaining creep life is based on methods
and data in ASME-FFS1/API-579 Fitness for Service Standard [2]. This provides mean and
minimum rupture and strain accumulation data. The high-temperature basis for acceptability of
some nondesign or damaged condition is usually defined by fractional life usage, based on creep
rupture data. This is a discontinuous criterion in the sense that, over time, the component
switches suddenly from acceptable to unacceptable on the basis of the creep or creep-fatigue life
usage calculation. Whether to use minimum or mean data for the calculation is not clearly
specified.
Other options are available in Reference 2, such as creep testing of material samples to determine
remaining life. This provides a basis for remaining life material properties, which may be
different from minimum.

7.1 Risk-Based Assessment

7.1.1 Background
There are limitations to the use of minimum life calculations as a basis for decision making on
the options for aging piping such as inspection, derating, testing, and replacement. For piping
that has reached the end of minimum life, the options are limited by the fact that there is little to
be done to change the life calculation.
The following are three main reasons for suggesting a risk-based approach:
 Risk-based methods are well-known and widely used in conventional applications.
 A risk-based approach makes use of the full set of creep rupture data rather than just the
minimum trend line.
 The calculations that form the basis for all high-temperature continuum (noncrack)
assessments give life usage typically based on minimum rupture data. This is a well-defined
and reasonable approach, but it lacks the ability to calculate or estimate risk. For example, if
a component has reached the end of life on this basis, there would be no way that the merits
of a pressure or temperature derate could be assessed. The fact of reaching the end of
calculated minimum life is unaffected by such changes.

7.1.2 Data
Figure 7-1 [25] shows rupture data for grade 22. The so-called maximum line is the basis for
minimum life calculations. Typically, ASME and API minimum data correspond to about the
lower 5% of the data, which is −1.645 standard deviations. The normal distribution placed over
the data shows how the log(time) – temperature Larsen-Miller parameter is reasonably well
characterized by a normal distribution.

10371024 7-1
Figure 7-1
Distribution of grade 22 rupture data

Figure 7-2 shows the information available for steady conditions from mean and minimum
rupture data. For this case, temperature = 1000°F (537.8°C), allowable stress = 8 ksi
(55,158,058.3 Pa). The cumulative failure probability is 5% at minumum life and 50% at mean
life by definition. The probability rate is skewed with a maximum at 250,000 hours. This is
because the rupture distribution in Figure 7-1 is normal in log(time), but power plants operate in
real time. The probability rate at the end of minimum life is 6  10−7, corresponding to about
0.5% per typical plant year. This should be interpreted as referring to a single pipe or spool in a
piping system.

7.1.3 Derivation of Failure Probability


The mean and minimum rupture lines define minimum and mean Larsen-Miller functions of
stress and temperature: LMPmin(s,T) and LMPmean(s,T).
These functions are defined in Reference 2. If the minimum data may be associated with a
percentage of the data or with a number of standard deviations, then the Larsen-Miller parameter
for any standard deviation (stdev) may be defined by:
stdev
LMPs ,T , stdev   LMPmeans ,T   LMPmeans ,T   LMP mins ,T  Eq. 7-1
 1.645

10371024 7-2
Clearly, if some value other than −1.645 is appropriate for the minimum, then it can be used in
this definition.
From Equation 7-1, rupture life may be calculated as follows:
LMP
life s ,T , stdev   10  LMConst Eq. 7-2
T
where
LMConst = 20 for trade 22 material, for example.
Conversely, if stress, temperature, and time (s,T,t) are known, then the number of standard
deviations may be calculated from

 t 
log  
stdevs ,T ,t    lifes ,T ,0   log  lifes ,T ,1 
 lifes ,T ,0  
 

log Damage   lifes ,T ,1 


= log   Eq. 7-3
 lifes ,T ,0 
The cumulative probability of failure (cdf) and the probability density function (pdf) are given by
the standard expressions for a Gaussian distribution. The expected failure rate is the time-
dependent density function prate(t) defined by:

 1   stdev 2 
  exp    stdevs ,T ,t 
 2   2 
pratet   Eq. 7-4
t

7.1.4 Use of Probability to Manage Component Risk


Temperature Derate. Figure 7-3 shows the effect of a continuous derate at the end of minumum
life. The assumptions in this calculation are the simplest possible, namely, the risk of a material
sample is defined by the accumulated damage and the current state of temperature and stress.
They show that the risk for end of minimum life components can be managed effectively with
modest reductions in temperature (and/or stress).
The use of a damage calculation on its own is of little value for this problem because of the
following:
There is no basis for accepting or rejecting a value >1.
There is no change in its value due to the derate.
There is only an imperceptible change in the damage rate after the derate.

10371024 7-3
There is no basis for accepting or rejecting a value >1.
There is no change in its value due to the derate.
There is only an imperceptible change in the damage rate after the derate.
Material Sampling and Testing. If samples are removed and tested for creep properties, and the
results characterized by a mean and standard deviation of the nominal data, then the probability
of failure may be recalculated. Figures 7-4 and 7-5 show how sample testing changes the
calculated failure probability. The value before the 20-year test was 0.005. As a result of the test,
the probablility is reset to zero and increases to 0.005 ten years later. Similarly, after 35 years,
the risk is reset to zero and increases to 0.005 five years later. Note that there is nothing special
about the value 0.005, other than it is the predicted risk at minimum life.

8.0E‐03
1.4
pdf
Expected failure rate per year

7.0E‐03 cdf
1.2
6.0E‐03 min life
1.0 mean life

cdf, damage
5.0E‐03
damage
0.8
4.0E‐03
0.6
3.0E‐03

2.0E‐03 0.4

1.0E‐03 0.2

0.0E+00 0.0

0 20 40 60 80 100
Operating years

Figure 7-2
Mean and minimum life, damage accumulation with cumulative failure probability (cdf), and
probability rate per year (pdf) for grade 22 design stress at 1000°F (537.8°C)

Figure 7-3
Effect of temperature derate on risk approaching minimum life

10371024 7-4
Figure 7-4
Results of sample testing at 20 and 35 years showing decreasing mean properties, with scatter in
sample tests less than nominal, but unaffected by time

Figure 7-5
Conclusions for piping risk assessment from sample tests at 20 and 35 years

10371024 7-5
10371024
8
HEAVY SECTION WYES AND TEES
Heavy section components can be assessed on their own. Girth weld joints to plain pipe sections
will be covered in the piping system analysis.
The failure modes to be considered are rupture in high constraint regions and creep-fatigue
cracking initiating on inside surfaces.
The area-replacement design methods are expected to be conservative for yielding and
conventional (low temperature) problems, but, as will be shown, in areas of high constraint,
cracking can occur in less than the nominal design life.
Creep-fatigue cracking is driven primarily by transient thermal stresses that are generally higher
for thicker sections. Spreadsheet calculations can be used to identify likely limiting conditions in
terms of steam temperature rate of change and wall thickness.
The methods described in this section can be applied to plain pipe or other geometries.

8.1 Rupture Life Prediction


Figure 8-1 shows a wye joint, which is as strong as the plain pipes that it connects but is
predicted to have a shorter rupture life due to high constraint.
For a thick pipe, the sustained pressure stress may be conservatively estimated by the Tresca
reference stress:
TP/ln(ro/ri).
For a more complex geometry, a finite element limit analysis is required to obtain the sustained
MPS.

8.2 Transient Thermal Stress and Creep Fatigue


The following data are required:
Steam temperature rate Tdot
Wall thickness d
Material properties
Density 
Specific heat C
Thermal conductivity k
Material diffusivity C/k

10371024 8-1
Modulus E
Expansion coefficient 
Poisson’s ratio 
Provided the temperature ramp continues for at least time = d2/then thermal stress is
approximately
T Tdot d2/[2]
The temperature ramp up and ramp down rates are likely to be different. The stress range for a
complete cycle is the sum of thermal stresses calculated for both cases. Note that if the stresses
are calculated with temperature rates of different signs, then the total thermal stress range is the
difference between the positive and negative stresses or the sum of the absolute values.
Total stress range. The total stress range includes the maximum pressure stress, which should
be added to the thermal stress range. The hoop pressure stress in the bore of a thick pipe (radii ri
and ro) is obtained from the Lame solution:
 PP(ro2 + ri2)/(ro2  ri2).
For a more complex geometry, a finite element stress analysis is required.

8.3 Sustained Stress and Creep Rupture


For a thick pipe, the sustained pressure Tresca and maximum principal stress may be
conservatively estimated by the Tresca reference stress:
TP/ln(ro/ri).
For a more complex geometry, a finite element limit analysis is required to obtain the sustained
MPS, which defines rupture life.
Figure 8-1a shows MPS contours in the wall of a heavy wye, obtained from a limit analysis.
Figure 8-1b shows reference stress histories of von Mises, MPS, and the plain pipe design stress
as creep strain develops. In spite of the conservative wye design, there are higher MPS regions
than in the plain pipe, which are potential sources for cracking and rupture. A limit analysis was
performed, which used an arbitrary yield stress.

10371024 8-2
(a) (b)

Figure 8-1 (a) Wye section with MPS contours and (b) Wye section with stress histories

8.4 Creep-Fatigue Damage


The effect of cyclic loading on creep damage may be calculated using a shakedown argument as
described in Section 4. If:
Total stress range < cold yield stress + minimum rupture stress for an assumed life, then:
Cyclic life > assumed life.
It is straightforward to iterate and calculate cyclic life and creep damage with this method.
Fatigue damage may be calculated using the data from Reference 2. With a creep-fatigue
interaction rule also given in Reference 2, the combined creep-fatigue damage for the service
time and cycles may be calculated.
In practice, it is likely that the cyclic creep calculation will make it clear whether the thermal
transient has any significant effect on the steady load life.
For example, consider a grade 22 pipe with a section 6 in. (152.4 mm) thick. The allowable
stress at 1000°F (537.8°C) (8 ksi [55,158,058.3 Pa]) has a minimum rupture life of 1.8  105
hours. Using the previous calculations:
Consider pressure hoop stress = 8 ksi (55,158,058.3 Pa)
Effective stress (pipe) = 6.9 ksi (47,573,825.3 Pa)
Modulus E = 24,570 ksi (1.694041867e+11 Pa)
Expansion coefficient  = 7.8e-6 F−1

10371024 8-3
Diffusivity  = 34.3 in.2/h (871.2 mm2/h)
Yield (100°F [37.8°C]) = 29.4 ksi (202,705,864.4 Pa)
Assume equal heating and cooling temperature rates of 100°F/h (37.8°C/h). (This is probably an
underestimate of typical maximum rates.)
Thermal stress for heating and cooling (ignoring the difference in sign):
 TTdotd2/[2−] = 14.4 ksi (99,284,505 Pa).
Thermal stress ranges: hoop stress = axial stress = 28.8 ksi (198,569,010 Pa).
Axial stress range = (pressure + thermal) stress range = 4 + 2  14.4 = 32.8 ksi (226,148,039.2
Pa).
For 1.8  105-hour life, allowable effective stresses: cold 29.4 ksi (202,705,864.4 Pa), hot 6.9 ksi
(47,573,825.3 Pa).
For hot hoop stress = 8 ksi (55,158,058.3 Pa), hot axial stress = 4 ksi (27,579,029.2 Pa), hot
effective stress = 6.9 ksi (47,573,825.3 Pa).
The steady state has no favorable residual stress as far as thermal cycles are concerned, so the
range is 2  calculated thermal stress, with different signs for tube outer diameter and inner
diameter.
Calculate Tmax such that hoop stress 8 − Tmax, axial stress = 4 − Tmax lie on the cold yield
surface with effective stress = 29.4 ksi (202,705,864.4 Pa). Result: Tlim = −23.2 ksi
(159,958,369.2 Pa).
Conclusion: The thermal stress due to heating and cooling rates of 100°F/h (37.8°C/h) do not
give shakedown from the high-temperature steady-state stress.
Therefore, cyclic minimum life is less than steady minimum life = 1.8  105 hours.

10371024 8-4
9
REFERENCES
1. ASME B31.1a- 2005, Power Piping, American Society of Mechanical Engineers, New York,
2005.
2. Fitness for Service API 579-1/ASME FFS-1, ASME, New York, 2007.
3. ASME BPV Code Section III Subsection NH. ASME New York, 2007.
4. J. J. Balaschak and R. Wray, “Pipe Support Hanger Evaluation in Operating Fossil Power
Plants.” ASME Pressure Vessels and Piping Conference, Pittsburgh, PA (June 1988).
5. B. Hahn, G. Buhl, and D. Nerger, “In-Service Condition Monitoring of Piping Systems,”
OMMI. Vol. 1, No. 1 (April 2002).
6. J. Thurley and H.Thielsch, “Critical Component Failures Associated with Pipe Supports:
Causes and Prevention.” PVP Vol. 236, Valves, Bolted Joints, Pipe Supports, and Restraints.
ASME (1992).
7. J. P. Gephart and L. E. Kimball, “Case Study: Critical Steam Piping Constant Support
Hanger Testing Systems.” PVP Vol. 236, Valves, Bolted Joints, Pipe Supports, and
Restraints. ASME (1992).
8. L. E. Kimball, “Aging Pipe Supports–A Photographic Study.” Power 2008-60091, Proc.
ASME Power 2008 Conference, Orlando, Florida (July 2008).
9. G. May, “The Need for In-Situ Pipe Support Testing.” Proc. ASME Power 2004, Baltimore
(2004).
10. L. E. Kimball and M. Johnson, “Interpreting Constant Support Hanger Readings.” Power
2009-81170, Proc. Power 2009, ASME Power 2009, Albuquerque, NM (2009).
11. R. K. Penny and D. L. Marriott, Design for Creep, 2nd ed. Chapman and Hall 1995.
12. F. R. Hall and D. R. Hayhurst, “Continuum Damage Mechanics Modeling of High
Temperature Deformation and Failure in a Pipe Weldment,” Proc. R. Soc. Lond. (A). Vol.
433, p. 383-403 (1991).
13. E. Molinieux, P. Martel, C. Duquenoy, P. Dupas, and V. Prunier, “Creep Behavior of Seam-
Welded Reheat Steam Pipes in Thermal Fossil Power Plant: Feedback Analysis and Life
Assessment Methodology,” Materials at High Temperatures. Vol. 15, p. 375-384.
14. T. H. Hyde, W. Sun, and J. A. Williams, “Creep Analysis of Pressurized Circumferential
Pipe Weldments–A Review,” J Strain Analysis. Vol. 38, No. 1 (2003).
15. W. Payten, “Large Scale Multi-zone Creep Finite Element Modeling of a Main Steam Line
Branch Intersection,” Int. J. Press. Vessels and Piping. Vol. 83, p. 359-364 (2006).
16. T. H. Hyde, W. Sun, and A. A. Becker, “Life Assessment of Weld Repairs in ½Cr½Mo¼V
Main Steam Pipes Using the Finite Element Method,” J. Strain Analysis. Vol. 35, No. 5, p.
359-372 (2000).

10371024 9-1
17. T. H. Hyde, J. A. Williams, and W. Sun, “Assessment of Creep Behavior of a Narrow Weld,”
Int. J. Press. Vessels and Piping, Vol. 76, p. 515-525 (1999).
18. R. G. Sim, “Reference Stress Concepts in the Analysis of Structures During Creep,” Int. J.
Mech. Sci. Vol. 12, p. 561-573 (1970).
19. R5 Assessment Procedure for the High Temperature Response of Structures, Issue 3, British
Energy Generation Ltd. (2003).
20. B. J. Cane, “Present Status of Predictive Methods for Remnant Life Assessment and Future
Developments,” ERA report, published in the Residual Life Prediction issue of Metals
Forum, the journal of Australasian Institute of Metals (1985).
21. J. R. Rice and D. M. Tracey, “On the Ductile Enlargement of Voids in Triaxial Stress
Fields,” J. Mech. Phys. Solids. Vol. 17, p. 210-217 (1969).
22. G. A. Webster and R. A. Ainsworth, High Temperature Component Life Assessment.
Chapman and Hall 1994.
23. S. Malek1, Y.-H. Zhang, and K. Nibkin. “Validation of Creep Crack Growth NSW Model in
Exreapolating Short Term Laboratory Test Results to Longer-Term Service Component
Failure Prediction.”
24. Abaqus Users Manual, Dassault Simulia Systems Corp., Providence, RI.
25. Grade 22 Low Alloy Steel Handbook. EPRI, Palo Alto, CA: 2005. 1012840.

10371024 9-2
A
CASE STUDIES: HOT REHEAT AND MAIN STEAM
PIPING
A.1 History
The unit began operating in 1977. The earliest available inspection report is from 1991. This
refers to a range of problems as follows:
 A bulge below the top wye is thought to be due to overheating during erection for alignment
for the final weld.
 Hangers 1 and 2 are supporting nearly twice the design load during the hot condition.
 Hangers 3, 5, and 7 are “undersupporting” the pipe by 11 ksi (75,842,330.2 Pa) based on
comparison with design data.
 Other hangers are bottomed or topped out.
 Cracks and defects in the lower wye and in the wye circumferential weld HAZs.
 Cracks in hanger attachments and lugs.
 Hangers 7, 12, and 14 were topped out in the cold condition.

This report summarizes actions from earlier reports and focuses on the low point at the bottom of
the riser that would prevent natural draining. Replacements after the 1991 reports were the upper
wye and 10 ft (3 m) of pipe below the wye and repairs, replacements, and adjustments to eight
hangers.
A review in 1992 noted that there was a low point at the bottom of the riser that would prevent
natural draining. Replacements after the 1991 reports were the upper wye and 10 ft (3 m) of pipe
below the wye and repairs, replacements, and adjustments to eight hangers.
A 1994 report covered reinspection of main steam and hot reheat (HR) piping components. At
this time (1993), Unit 1 had seen 120,000 operating hours operating time. Some material was
removed at HR lower wye before welding repairs. Replicas showed signs of creep damage in a
girth weld. The main steam inspection was restricted to a repaired thermowell.
In 1997 [4], the HR wye was inspected. Girth weld cracking in HR-12 region. This wye has a
history of weld cracking. The inlet weld was replaced twice due to extensive creep damage.
Hanger adjustments were made in the early 1990s to reduce high stress concentrations on the
wye attachment weld.
In 2008 [5], an NDE and metallurgical evaluation was performed. NDE and metallurgical
evaluations were performed on 10 girth weld and eight seam welds. No evidence of creep
damage was found.

10371024 A-1
In 2009 [6], a Unit 1 pipe hanger walkdown report refers to broken springs on the final steam
hanger outlet header spring hangers. Mentions bad design of HRH 5, 7 ,8 ,9 ,11, 14, 15, 16
hangers. There were discrepancies with design loads and displacements and detailed walkdown
information on hanger loads and displacements. Recommendations for individual hanger repairs
and adjustments were made.

A.2 Conclusions from Reports


There is a clear history of continual hanger problems with a number of adjustments and changes.
Notable piping problems are the replaced bulged section below the top wye and repeated girt
weld repairs on the lower wye. No plausible operational or design cause for this bulge can be
determined.
The HR line contrasts with the Unit 1 main steam line, which does not appear to have had a
history of hanger problems. The changes in main steam hanger loads are presumed to be
associated with either observations of piping deflection or with a review of the design. The
calculated deflections in this report point to similar problems.

A.3 HR Piping
Design Data

Table A-1
Design data

Pressure (psi) Temp (°F/°C)


825 1015/546.1

3850 1010/543.3

Material SA335 P22 Seamless


Elbow Specified Minimum
Outer Diameter Radius Wall Nominal Wall
Boiler - top wye 30.25 150 1.595 1.9
Main single lead 39.75 196 2.25 2.375
Turbine leads 30.25 150 1.595 1.8
Turbine
connections 28.25 28 1.5 1.6

All values are in inches (1 in. = 25.4 mm); 1 psi = 6894.8 Pa.

10371024 A-2
Figure A-1
HR line with constant load hangers. Based on the proximity to vertical legs, the likely critical
hangers that reduce piping life when locked are the first in the lower horizontal section and the
hangers are closest to the outlet.

A.3.1 Analysis Cases


Analysis of load histories.
1. Step 1: Self-weight, hanger loads
2. Step 2: Thermal + pressure ramp (completion of analysis for limit and creep analyses)
3. Step 3: Thermal + pressure cycles (cyclic analysis)
4. Check for stress, displacements
5. Shakedown (no yielding on final unloading)
6. Limit analysis for reference stress calculation
7. Creep stress analysis (unsuccessful for pipe elements)
8. Shakedown analysis to identify critical hangers that will reduce piping life if they lock

10371024 A-3
A.3.2 Analysis 1: Elastic-Plastic Thermal-Mechanical Loading
Figure A-2 shows a plot of stresses at with thermal and mechanical loading. The two wyes are
regions of high stress that will be evaluated. The piping did not yield during the cycle startup-
shutdown cycle, indicating that cyclic loading is not expected to reduce life under design
conditions.

Figure A-2
HR line: elastic-plastic analysis. Stress maximal at wyes and western outlet elbow

10371024 A-4
A.3.3 Analysis 2: Isochronous Elastic-Plastic Analysis
Figures A-2 and A-3 show deformation and strain predictions for 1  105 and 5  105 hours. It
predicts that the boiler-top wye sections will lift, and the horizontal section will drop in the
northeast corner.

Figure A-3
Isochronous elastic-plastic analysis. Strain (0.7%) and vertical displacement (−19 in. [482.6 mm] at
horizontal bend): 1e5 hours

10371024 A-5
Figure A-4
Isochronous elastic-plastic analysis. Strain and vertical displacement: 5e5 hours The horizontal
bend has not dropped significantly after 100,000 hours.

10371024 A-6
A.3.4 Analysis 3: Creep and Damage Analysis

Figure A-5
Creep analysis showing mean creep life calculated from von Mises stress limited by top wye and
outlets to the turbine at 1.4e6 hours. Max creep strain = 1.2%.

10371024 A-7
A.3.5 Analysis 4: Limit Analysis for Reference Stress Creep Damage Calculation
Table A-2 shows a summary of the creep reference stress life calculation.
The limit analysis gives plastic strain values at each increment. Table A-1 shows the maximum
value that occurs at the east termination outlet. The table shows the reference stress calculations
and corrections for welds, MPS, and strain. The damage contours in Figure A-4 are based in von
Mises stress and can be compared with the calculations in Table A-1. For plastic strain between
0.003% and 1.5%, the predicted mean life is between 1.4  106 and 1.6  106 hours. The creep
stress analysis gives a value of 1.4  106 hours.
This shows that the reference stress calculation is consistent with the results of the full creep
stress analysis. When applied to rupture with weld and MPS corrections, the predicted life is
between 4.3  104 hours (minimum data) and 1.9  105 hours (mean data).

10371024 A-8
Table A-2
Interpretation of limit analysis for creep life prediction

Reference
von Mises Rupture (MPS)
Stress
Plastic/ Strain MPS Weld Minimum Mean Minimum Mean
Step Time Deformation Rupture
Creep Strain Factor Factor Factor Life Life Life Life
0.688 0.00E+00 1.00 1.15 0.90 7.27 9.29 1.8E+05 7.8E+05 4.0E+04 1.7E+05
0.702 1.57E-06 1.00 1.15 0.90 7.13 9.11 2.1E+05 8.8E+05 4.5E+04 2.0E+05
0.716 8.53E-06 1.00 1.15 0.90 6.99 8.93 2.3E+05 9.9E+05 5.1E+04 2.2E+05

0.737 1.80E-05 1.00 1.15 0.90 6.79 8.67 2.8E+05 1.2E+06 6.1E+04 2.6E+05
0.745 2.28E-05 1.00 1.15 0.90 6.71 8.58 3.0E+05 1.3E+06 6.5E+04 2.8E+05
0.756 3.13E-05 1.00 1.15 0.90 6.61 8.45 3.3E+05 1.4E+06 7.2E+04 3.1E+05
0.774 1.52E-02 1.11 1.15 0.90 6.46 9.16 2.0E+05 1.6E+06 4.3E+04 1.9E+05
0.779 2.36E-02 1.16 1.15 0.90 6.42 9.52 1.6E+05 1.6E+06 3.4E+04 1.5E+05
0.783 3.70E-02 1.22 1.15 0.90 6.38 9.98 1.2E+05 1.7E+06 2.6E+04 1.1E+05
0.784 4.13E-02 1.24 1.15 0.90 6.38 10.10 1.1E+05 1.7E+06 2.4E+04 1.0E+05
0.799 3.50E-01 1.39 1.15 0.90 6.26 11.14 5.9E+04 1.9E+06 1.3E+04 5.5E+04

Note: Strain is in inches (1 in. = 25.4 mm). Deformation and rupture stress are in ksi (1 ksi = 6,894,757.3 Pa); life is in hours. Factors are unitless.

10371024 A-9
A.3.6 Analysis 5: Shakedown Analysis to Identify Critical Hangers
The effect on life of any of the hangers in Figure A-5 becoming locked is identified with a cyclic
analysis. It is trial-and-error procedure. Figure A-6 shows yield and minimum rupture stress for
grade 22 material.

Figure A-6
Use of yield, 1  105 and 5  105 hour rupture data to define “yield” stresses for shakedown analysis

Figure A-7
HR line with constant load hangers, identifying the critical hanger when locked. The analysis
checked the first three hangers on the horizontal section.

10371024 A-10
Figure A-8
Inlets to upper wye are the predicted regions for creep-fatigue damage when the critical hanger is
locked

A.4 Main Steam Piping

A.4.1 Design Data


Table A-3
Design data

Pressure (psi) Temperature (°F/°C)


Design drawing 4025 1020/548.9
Inspection specification 3850 1010/543.3

Material SA335 P22 Seamless


Caesar Design
Outer Diameter (in.) Specified Minimum Wall
Wall Thickness
Boiler pipe 22 5.3125 5.4375
Boiler - top wye 20 4.25 4.4375
Main single lead 26.5 5.432 5.5625
Turbine leads 16.5 3.5 3.5625

Note: 1 in. = 25.4 mm; 1 ksi = 6,894,757.3 Pa.

10371024 A-11
A.4.2 Analysis Cases
This section demonstrates the use of shell elements for predictions of constant load strain,
deflection, and creep damage.
The analyses in this section are as follows:
 Isochronous elastic-plastic analyses
 Creep stress and damage analyses
 Limit load and reference stress analyses

The available reports did not mention problems with the main steam hangers. Inspection reports
mention isolated defects rather than evidence of structural and material degradation such as creep
and creep fatigue. Table A-4 shows design and current hanger loads. The changes are mainly in
the distribution in the upper piping, possibly due to observed deformations similar to those in
Figure A-11.
Life assessments are based on operating conditions: pressure = 3850 psig (26,544,980 Pa) and
temperature = 1010°F (543.3°C).

A.4.3 Comparison of Reinforced and Unreinforced Joints Between Turbine Leads


and Piping
Details of joints between turbine leads and the 26.5-in. (673.1-mm) pipe were not available.
Initial analysis of joints without reinforcement indicated early damage at the joints (Figure A-9).
This is unlikely based on the absence of any mention of cracking at the joints. The analysis then
was performed with creep and damage in the joints suppressed by eliminating temperature, as
shown in Figure A-10.

10371024 A-12
Table A-4
Main steam preloads and constant loads (in kips)

Original Current
Left Center Right Left Center Right
Inlet horizontal 13.46 13.46 14.3 12.1
Horizontal +2  14.6 +2  14.9 +2  28.9 +2  29.6
Horizontal 29.6 16.7 17.5 0
Horizontal 23.8 23.2 23.8 23.2
Wye net load 54 54
Vertical 60.7 55.8
Vertical 60.6 60.6
Horizontal 44.7 44.2
Horizontal 43.5 42.7
Horizontal 36 39
Horizontal 44 44
Wye net load 39 39
Turbine lead 11.4 11.4 10.9 11.4

Note: 1 kip = 453.6 kg-force.

Figure A-9
Predicted minimum life of unreinforced turbine lead joints = 25,000 hours

10371024 A-13
Figure A-10
Temperature plot showing how the turbine leads unreinforced joints are eliminated from creep
calculations

A.4.4 Elastic-Plastic Isochronous Analysis for Deflection and Strain


Figure A-11 shows predicted deflections at 105 hours based on an isochronous elastic-plastic
analysis. It suggests that the hangers near the horizontal bend and on the left-hand top pipe may
have been slightly oversized. Figure A-12 shows strain predictions, identifying the inner surface
of the last turbine lead as a likely high damage point.

A.4.5 Creep and Damage Analysis


Figure A-13 shows the high creep damage region in the turbine lead as expected. Predicted
minimum creep life is 105 hours.

A.4.6 Limit Analysis and Reference Stress Calculations of Life


Comparison with reference stress life prediction.
A limit analysis was performed with an arbitrary yield stress = 5 ksi (34473786.5 Pa). The piping
reached its limit at pressure = 0.764 of the operating pressure = 2.94 ksi (20270586.4 Pa). Table
A-5 shows the results of reference stress and rupture life calculations. The region is the same
region of the turbine lead, as identified in Figure A-13. The calculated minimum life is in good
agreement with Figure A-13.

10371024 A-14
A.4.7 Calculation of Probability of Failure
Using the methods of Section 7, the probability of failure can be calculated under conditions of
changing stress. Figure A-14 shows the comparison of risk in the turbine lead with and without
thermal stress. Elimination or reduction of thermal stress is possible with cold pull during
erection. This turbine leads are vulnerable to thermal stress because they are weak compared
with the main steam pipe and will absorb all the creep strain required to relax piping thermal
stress. The case with thermal stress has higher stress and risk during the first 20 years of life.
After that, the differences in risk decrease. This makes clear the need to consider a weak part in
the piping system, which then absorbs concentrated strain during stress redistribution.

Figure A-11
Predicted 100,000-hour vertical deflections for original hanger loads. Plots have different
magnifications. Top leads + 1 in. (25.4 mm), bottom horizontal bend +10 in. (254 mm)

10371024 A-15
Figure A-12
Isochronous elastic-plastic analysis. Predicted 100,000-hour inelastic strains (maximum 0.7%)

10371024 A-16
Figure A-13
Damage plots identifying the critical turbine lead with minimum life = 1.0  105 hours

10371024 A-17
Table A-5
Reference stress life calculations

Reference
Deformation Rupture
Stress
Step Plastic/Creep Strain MPS Weld Minimum
Time Deformation Rupture Minimum Life Mean Life Mean Life
Strain Factor Factor Factor Life
0.719
0.00E+00 1.00 1.15 1.00 6.96 8.00 2.9E+05 1.3E+06 8.3E+04 3.6E+05
0.738
4.47E-05 1.00 1.15 1.00 6.78 7.80 3.4E+05 1.5E+06 9.7E+04 4.2E+05
0.756
6.41E-04 1.00 1.15 1.00 6.61 7.61 4.0E+05 1.7E+06 1.1E+05 4.8E+05
0.761
1.35E-03 1.00 1.15 1.00 6.57 7.57 4.1E+05 1.8E+06 1.2E+05 5.0E+05
0.762
1.66E-03 1.00 1.15 1.00 6.56 7.56 4.1E+05 1.8E+06 1.2E+05 5.0E+05
0.763
2.11E-03 1.00 1.15 1.00 6.55 7.56 4.1E+05 1.8E+06 1.2E+05 5.1E+05
0.764
2.39E-03 1.00 1.15 1.00 6.55 7.55 4.1E+05 1.8E+06 1.2E+05 5.1E+05
0.764
2.74E-03 1.00 1.15 1.00 6.54 7.55 4.1E+05 1.8E+06 1.2E+05 5.1E+05
0.765
3.15E-03 1.00 1.15 1.00 6.54 7.55 4.1E+05 1.8E+06 1.2E+05 5.1E+05
0.766
3.67E-03 1.01 1.15 1.00 6.53 7.55 4.1E+05 1.8E+06 1.2E+05 5.1E+05
0.766
4.33E-03 1.01 1.15 1.00 6.53 7.56 4.1E+05 1.8E+06 1.2E+05 5.1E+05
0.767
5.16E-03 1.01 1.15 1.00 6.52 7.56 4.1E+05 1.8E+06 1.2E+05 5.0E+05
0.767
6.23E-03 1.01 1.15 1.00 6.52 7.57 4.1E+05 1.9E+06 1.2E+05 5.0E+05
0.768
7.63E-03 1.01 1.15 1.00 6.51 7.58 4.1E+05 1.9E+06 1.2E+05 5.0E+05
0.769
9.53E-03 1.01 1.15 1.00 6.51 7.59 4.0E+05 1.9E+06 1.1E+05 4.9E+05
0.769
1.22E-02 1.02 1.15 1.00 6.50 7.62 3.9E+05 1.9E+06 1.1E+05 4.8E+05

10371024 A-18
0.016 14

0.014 12
grade 22 7.5 ksi design stress at 1010 F
0.012
Probability per year

10 turbine ( lead no thermal stress)

0.01 turbine lead (with thermal stress)

Stress
8 stress (no thermal stress)
0.008 stress (with thermal stress)
6
0.006

4
0.004

0.002 2

0 0
0 5 10 15 20 25 30 35 40 45

Years

Figure A-14
Comparative risks: turbine leads with and without thermal stress and grade 22 material design
stress (7.5 ksi [51710679.7 Pa]) at 1010°F (543.3°C)

A.4.8 Conclusions
The main steam deflections and strains indicate a better designed HR line. It is not surprising that
no significant problems with hangers were reported.
The turbine lead on the end of the line absorbs the most strain and damage due its relative
weakness. The inner crotch indicated in the figures would be the area of focus for inspection and
any sample testing. If the joints to the other turbine leads are not reinforced, then they are the
most vulnerable regions.

10371024 A-19
10371024
B
ISOCHRONOUS AND CREEP DATA: GRADE 22 AND
GRADE 91
This appendix presents ASME III NH isochronous creep data in the form of hardening yield
stress.
The ASME III NH data are in graphical form. The following tables were calculated by reading
the isochronous curves and subtracting elastic strain to give total inelastic strain.
Table B-1
Grade 22 at 1000°F (537.8°C)
10,000 Hours 100,000 Hours 500,000 Hours
Stress Plastic Strain Stress Plastic Strain Stress Plastic Strain
4.00 0.0000 3.50 0.0000 2.4000 0.0000
8.82 0.0030 6.76 0.0031 5.2651 0.0031
10.16 0.0063 7.57 0.0064 6.0329 0.0064
10.86 0.0096 8.05 0.0097 6.5814 0.0097
11.34 0.0129 8.36 0.0130 6.7642 0.0131
11.69 0.0162 8.60 0.0163 6.9470 0.0164
12.02 0.0195 8.78 0.0196 7.1298 0.0197
12.24 0.0228 8.95 0.0230 7.3126 0.0230
12.46 0.0262 9.15 0.0263 7.4954 0.0264
12.68 0.0295 9.41 0.0296 7.6782 0.0297

Note: Stress is in ksi (1 ksi = 6,894,757.3 Pa); plastic strain is in inches (1 in. = 25.4 mm).

Table B-2
Grade 22 at 1050°F (565.6°C)
10,000 Hours 100,000 Hours 500,000 Hours
Stress Plastic Strain Stress Plastic Strain Stress Plastic Strain
2.00 0.0000 2.00 0.0000 1.00 0
6.47 0.0031 4.80 0.0031 3.5 0.0032
7.53 0.0064 5.62 0.0064 4.1 0.0065
8.05 0.0097 5.99 0.0098 4.6 0.0098
8.46 0.0130 6.25 0.0131 4.7 0.0131
8.68 0.0163 6.44 0.0164 4.8 0.0165
8.92 0.0196 6.66 0.0197 5.0 0.0198
9.16 0.0230 6.88 0.0230 5.1 0.0231
9.33 0.0263 7.01 0.0264 5.3 0.0264
9.55 0.0296 7.27 0.0297 5.4 0.0298
Note: Stress is in ksi (1 ksi = 6,894,757.3 Pa); plastic strain is in inches (1 in. = 25.4 mm).

10371024 B-1
Table B-3
Grade 91 at 1000°F/537.8°C

1000 Hours 10,000 Hours 100,000 Hours


Stress Plastic Strain Stress Plastic Strain Stress Plastic Strain
17.00 0.0000 13.00 0.0000 9.00 0.0000
22.35 0.0010 18.81 0.0012 15.62 0.0013
27.01 0.0028 23.01 0.0030 19.27 0.0032
29.48 0.0047 25.14 0.0049 21.00 0.0051
31.01 0.0066 26.35 0.0068 22.15 0.0070
31.93 0.0086 27.09 0.0088 22.75 0.0090
33.45 0.0125 28.48 0.0128 23.79 0.0130
34.46 0.0165 29.54 0.0167 24.59 0.0169
35.15 0.0205 30.29 0.0207 25.25 0.0209

Note: Stress is in ksi (1 ksi = 6,894,757.3 Pa); plastic strain is in inches (1 in. = 25.4 mm).

Table B-4
Grade 91 at 1050°F (565.6°C)

1000 Hours 10,000 Hours 100,000 Hours


Stress Plastic Strain Stress Plastic Strain Stress Plastic Strain
10.00 0.0000 7.00 0.0000 5.00 0.0000
15.80 0.0013 12.67 0.0014 9.96 0.0015
19.56 0.0031 16.21 0.0033 12.93 0.0034
21.55 0.0050 17.96 0.0052 14.42 0.0053
22.99 0.0069 19.06 0.0071 15.40 0.0073
23.95 0.0089 19.85 0.0091 16.02 0.0093
25.45 0.0128 21.12 0.0130 17.00 0.0132
26.48 0.0168 22.05 0.0170 17.91 0.0172
27.30 0.0208 22.65 0.0210 18.51 0.0212

Note: Stress is in ksi (1 ksi = 6,894,757.3 Pa); plastic strain is in inches (1 in. = 25.4 mm).

10371024 B-2
Table B-5
Grade 91 at 1100°F (/593.3°C)

1000 Hours 10,000 Hours 100,000 Hours


Stress Plastic Strain Stress Plastic Strain Stress Plastic Strain
8.00 0.0000 6.00 0.0000 3.00 0.0000
14.32 0.0014 11.42 0.0015 8.74 0.0016
17.39 0.0033 13.96 0.0034 10.70 0.0036
18.68 0.0052 14.92 0.0054 11.49 0.0055
19.47 0.0072 15.66 0.0074 12.07 0.0075
20.04 0.0092 16.31 0.0093 12.52 0.0095
21.10 0.0131 17.17 0.0133 13.29 0.0135
21.55 0.0171 17.48 0.0173 13.65 0.0174
22.08 0.0211 17.94 0.0213 14.13 0.0214

Note: Stress is in ksi (1 ksi = 6,894,757.3 Pa); plastic strain is in inches (1 in. = 25.4 mm).

Data for Abaqus creep stress analysis model.


n
d c  Q   
 exp   sinh   
dt  RT   0 

Data source: Reference 1 data for mean secondary creep rate.


Units: ksi (Pa), F° (C°)

Grade 22
Temperature range: 900–1100°F (482.2–593.3C°)
Stress range: 2–20 ksi

A 7.933E+20
0 5.671
Q 4.674E+05
n 5.033
R 4.619

Note: 1 ksi = 6,894,757.3 Pa

10371024 B-3
Grade 91
Temperature range: 900–1300°F (482.2–704.4°C)
Stress range: 3–30 ksi

A 2.156E+34
0 11.532
Q 6.983E+05
n 6.456
R 4.619

Note: 1 ksi = 6,894,757.3 Pa

Data source: Reference 2

10371024 B-4
10371024
Export Control Restrictions The Electric Power Research Institute Inc.,
Access to and use of EPRI Intellectual Property is granted (EPRI, www.epri.com) conducts research and
with the specific understanding and requirement that development relating to the generation, delivery
responsibility for ensuring full compliance with all applicable and use of electricity for the benefit of the public.
U.S. and foreign export laws and regulations is being An independent, nonprofit organization, EPRI
undertaken by you and your company. This includes an
brings together its scientists and engineers as well
obligation to ensure that any individual receiving access
hereunder who is not a U.S. citizen or permanent U.S. as experts from academia and industry to help
resident is permitted access under applicable U.S. and address challenges in electricity, including
foreign export laws and regulations. In the event you are reliability, efficiency, health, safety and the
uncertain whether you or your company may lawfully obtain environment. EPRI also provides technology, policy
access to this EPRI Intellectual Property, you acknowledge and economic analyses to drive long-range
that it is your obligation to consult with your company’s legal
research and development planning, and supports
counsel to determine whether this access is lawful.
Although EPRI may make available on a case-by-case research in emerging technologies. EPRI’s
basis an informal assessment of the applicable U.S. export members represent more than 90 percent of the
classification for specific EPRI Intellectual Property, you and electricity generated and delivered in the United
your company acknowledge that this assessment is solely States, and international participation extends to 40
for informational purposes and not for reliance purposes. countries. EPRI’s principal offices and laboratories
You and your company acknowledge that it is still the
are located in Palo Alto, Calif.; Charlotte, N.C.;
obligation of you and your company to make your own
assessment of the applicable U.S. export classification and Knoxville, Tenn.; and Lenox, Mass.
ensure compliance accordingly. You and your company
understand and acknowledge your obligations to make a Together…Shaping the Future of Electricity
prompt report to EPRI and the appropriate authorities
regarding any access to or use of EPRI Intellectual Property
hereunder that may be in violation of applicable U.S. or
foreign export laws or regulations.

© 2010 Electric Power Research Institute (EPRI), Inc. All rights reserved.
Electric Power Research Institute, EPRI, and TOGETHERSHAPING THE
FUTURE OF ELECTRICITY are registered service marks of the Electric
Power Research Institute, Inc.
1019632

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813 • USA
10371024800.313.3774 • 650.855.2121 • askepri@epri.com • www.epri.com

You might also like