You are on page 1of 108

Service Experience with Grade 91 Components

1018151

14166640
14166640
Service Experience with Grade 91 Components
1018151

Technical Update, August 2009

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
14166640800.313.3774 ▪ 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.

ORGANIZATION(S) THAT PREPARED THIS DOCUMENT

Electric Power Research Institute (EPRI)

Structural Integrity Associates, Inc.

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 © 2009 Electric Power Research Institute, Inc. All rights reserved.

14166640
CITATIONS
This document was prepared by
Electric Power Research Institute (EPRI)
1300 W. T. Harris Blvd.
Charlotte, NC 28262
Principal Investigator
K. Coleman
Structural Integrity Associates, Inc.
2904 South Sheridan Way, Suite 303
Oakville, ON, L6J 7L7
Principal Investigators
M. Clark
J. D. Parker
This document describes research sponsored by EPRI.
This publication is a corporate document that should be cited in the literature in the following
manner:
Service Experience with Grade 91 Components. EPRI, Palo Alto, CA: 2009. 1018151.

14166640 iii
14166640
PRODUCT DESCRIPTION
This report documents recent utility experience with Grade 91 alloy components operating at
high temperatures. It presents case studies that illustrate the performance trends that are
emerging as experience is gained with this relatively new alloy.

Results and Findings


Through the analysis of these longer-term failures, it is becoming clear that Grade 91 alloys are
inherently susceptible to the Type IV failure mode associated with weldments. The identified
trends fall into three general categories: 1) improper or poor control of initial heat treatment
variables that result in inferior properties or damage, 2) design or operational issues that increase
stresses and failures, and 3) longer-term performance issues of the alloy.
This report provides possible solutions or strategies to mitigate these problems in terms of proper
control of design and processing heat treatments, as well as suggestions regarding the material
properties that should be measured or monitored.

Challenges and Objectives


The objective of this study was to investigate the service experience and material performance of
Grade 91 alloy components.

Application, Value, and Use


This report should be valuable to utility engineers or managers who are charged with operating
plants that contain components fabricated with Grade 91 alloy. It provides examples of failures
that might be prevented if the utility is aware of the potential risks and adopts strategies of
monitoring and assessment.

EPRI Perspective
Overall, Grade 91 alloys have performed well. Most of the failures are related to fabrication or
heat treatment inconsistencies. When properly handled, these materials offer an excellent
alternative to existing CrMo steels. Precautions must be exercised, however, to realize long-term
component life.

Approach
The research team reviewed existing literature and industry experience with Grade 91 alloys.
This report presents case studies that illustrate the causes of damage and failures and provides
suggested solutions.

Keywords
Grade 91
Failure analysis
Heat treatment
Service performance

14166640 v
14166640
ABSTRACT
The Grade 91 alloy was developed to address a recognized need for a high-temperature alloy
with enhanced properties, including high creep rupture strength and satisfactory fracture
toughness. The alloy is similar to the 9Cr1Mo (T9) alloy, with controlled additions of vanadium,
niobium, and nitrogen. The alloy is intended to be used in the normalized and tempered
condition. Normalization at 1904°F (1040°C) for 1 hour, followed by air cooling, results in a
fully hardened martensitic structure, even in relatively thick sections, because of the high
hardenability of the alloy. Tempering of this structure (typically at 1345–1400°F [730–760°C])
results in a fine-grained, tempered martensite structure. Proper heat treatment results in M23C6
carbides at prior austenite and martensite lath boundaries, in addition to niobium and vanadium
intergranular carbonitrides. This combination of precipitates and martensitic structure provides a
marked increase in the high-temperature strength and creep resistance.
Mechanically, the modified 9Cr alloy provides improved high-temperature strength compared to
traditional alloys such as P11 or P22. This higher strength makes Grade 91 attractive for utility
applications because of the potential for higher operating temperatures and lower wall
thicknesses. The alloy has been used since the 1990s as a retrofit material in conventional
subcritical fossil plants and in high-temperature locations (up to 1100°F [593°C]) in new
combined-cycle and supercritical plants.
This report summarizes recent experiences and problems that utilities and suppliers have
encountered with the alloy. Detailed case studies of representative failures and problem areas are
used to illustrate the emerging trends. The identified trends fall into three general categories:
1) improper or poor control of initial heat treatment variables that result in inferior properties or
damage, 2) design or operational issues that increase stresses and failures, and 3) longer-term
performance issues of the alloy. Through the analysis of these longer-term failures, it is
becoming clear that Grade 91 is inherently susceptible to the Type IV failure mode associated
with weldments. This report provides suggestions for possible solutions to these issues in terms
of control of heat treatment and fabrication variables, improved designs, and monitoring
programs. Techniques for longer-term life assessments of Grade 91 components will be
addressed in separate Electric Power Research Institute (EPRI) programs.

14166640 vii
14166640
CONTENTS
1 INTRODUCTION AND BACKGROUND ................................................................................1-1
1.1 History and Characteristics of Grade 91 ..................................................................1-1
1.2 Typical Applications of Grade 91..............................................................................1-2
1.3 Scope of Current Report ..........................................................................................1-2
2 INDUSTRY EXPERIENCE—DAMAGE OR FAILURE IN TUBING........................................2-1
2.1 Reported Failures.....................................................................................................2-1
2.2 Case Studies–Fabrication-Related Damage ............................................................2-3
2.2.1 Case Study 1......................................................................................................2-3
2.2.2 Case Study 2......................................................................................................2-6
2.3 Case Studies—Oxidation-Related Failures..............................................................2-9
2.3.1 Case Study 3......................................................................................................2-9
2.3.2 Case Study 4....................................................................................................2-13
2.3.3 Case Study 5....................................................................................................2-17
2.4 Solutions.................................................................................................................2-21
3 INDUSTRY EXPERIENCE—BASE METAL DAMAGE OR FAILURE IN HEAVY SECTION
COMPONENTS, SHORT TERM ...............................................................................................3-1
3.1 Reported Failures.....................................................................................................3-1
3.2 Case Studies—Fabrication-Related Failures ...........................................................3-2
3.2.1 Case Study 6......................................................................................................3-2
3.2.2 Case Study 7......................................................................................................3-8
3.2.3 Case Study 8....................................................................................................3-13
3.3 Case Studies—Design-Related Failures................................................................3-15
3.3.1 Case Study 9....................................................................................................3-15
3.3.2 Case Study 10..................................................................................................3-18
3.4 Case Studies—Operation-Related Failures ...........................................................3-21
3.4.1 Case Study 11..................................................................................................3-21
3.5 Solutions.................................................................................................................3-24
4 INDUSTRY EXPERIENCE—BASE METAL DAMAGE OR FAILURE IN HEAVY SECTION
COMPONENTS, MEDIUM TERM .............................................................................................4-1
4.1 Reported Failures.....................................................................................................4-1
4.2 Case Studies—Fabrication-Related Failures ...........................................................4-2
4.2.1 Case Study 12....................................................................................................4-2
4.3 Case Studies—Material Chemistry–Related Failures ..............................................4-6
4.3.1 Case Study 13....................................................................................................4-6
4.4 Case Studies—Design-Related Failures................................................................4-13
4.4.1 Case Study 14..................................................................................................4-13
4.4.2 Case Study 15..................................................................................................4-18

14166640 ix
4.5 Solutions.................................................................................................................4-22
5 INDUSTRY EXPERIENCE—WELD-RELATED DAMAGE OR FAILURE .............................5-1
5.1 Reported Failures.....................................................................................................5-1
5.2 Case Studies—Fabrication-Related Failures ...........................................................5-2
5.2.1 Case Study 16....................................................................................................5-2
5.3 Case Studies—Design-Related Failures..................................................................5-6
5.3.1 Case Study 17....................................................................................................5-6
5.3.2 Case Study 18..................................................................................................5-13
5.4 Solutions.................................................................................................................5-16
6 DAMAGE ASSESSMENT TECHNIQUES..............................................................................6-1
6.1 Objective 1—Initial Material Integrity........................................................................6-2
6.2 Objective 2—Performance in Service ......................................................................6-2
7 CONCLUSIONS AND RECOMMENDATIONS ......................................................................7-1
8 REFERENCES .......................................................................................................................8-1

14166640 x
1
INTRODUCTION AND BACKGROUND
1.1 History and Characteristics of Grade 91
In 1974, the United States Department of Energy initiated a study to address deficiencies in types
304 and 316 stainless steels for use in the liquid-metal fast-breeder reactor program. An early
recommendation of that study was that there was a need to develop improved ferritic steels.
To that end, Combustion Engineering (now Alstom) developed a super 9Cr alloy, modified by
the addition of vanadium and niobium, which possessed enhanced mechanical properties,
including high creep-rupture strength coupled with satisfactory fracture toughness [1]. Oak
Ridge National Laboratory performed further assessments and additional laboratory tests, which
led to the Grade 91 alloy being approved by the ASME Boiler and Pressure Vessel Code
Sections I and VIII in 1983–1984 [2].
The Grade 91 alloy is similar to the 9Cr-1Mo (T9) alloy, with controlled additions of vanadium,
niobium, and nitrogen. The alloy is intended to be used in the normalized and tempered
condition. Normalization at 1904°F (1040°C) for one hour, followed by air cooling, results in a
fully hardened martensitic structure, even in relatively thick sections, because of the high
hardenability of the alloy. Tempering of this structure (typically at 1345–1400°F [730–760°C])
results in a fine-grained, tempered martensite structure. Proper heat treatment results in M23C6
carbides at prior austenite and martensite lath boundaries, in addition to niobium and vanadium
intragranular carbonitrides. This combination of precipitates and martensitic structure provides a
marked increase in the high-temperature strength and creep resistance. Table 1-1 lists the
chemical composition of the modified 9Cr material as originally specified by ASME.

Table 1-1
Chemical Composition of T91 (ASME SA-213/SA-213M Grade T91-1992)

Cb
C Si Mn P S Ni Cr Mo V Al N
(Nb)
0.08– 0.20– 0.30– 8.00– 0.85– 0.18– 0.030– 0.06–
≤0.020 ≤0.010 ≤0.40 ≤0.04
0.12 0.50 0.60 9.50 1.05 0.25 0.070 0.10

Mechanically, the modified 9Cr alloy provides improved high-temperature strength compared to
traditional alloys such as P11 or P22. This higher strength makes the Grade 91 alloy attractive
for utility applications because of the potential for higher operating temperatures and lower wall
thicknesses.

14166640 1-1
1.2 Typical Applications of Grade 91
The improved high-temperature strength of Grade 91 has made it attractive as a retrofit material
in conventional subcritical power plants and as a candidate material for advanced supercritical
plants. Plants with steam temperatures of up to 1100°F (593°C) are potential candidates for
Grade 91 components. The superior strength of Grade 91 permits thinner wall thicknesses in
some components.
As a retrofit material, sections of the boiler in which conventional ferritic steels have a high rate
of creep failures can potentially be replaced with T91 at less cost than austenitic alloys.
Waterwall, reheater, and superheater tubes have been replaced with T91 in some plants.
Superheater and reheat headers are other areas in which retrofits in P91 have been made. New
combined-cycle plants have also made extensive use of T91/P91 in main steam and reheat piping
as well as in boiler and steam generator components.

1.3 Scope of Current Report


The Electric Power Research Institute (EPRI) report Performance Review of P/T91 Steels
(1004516), published in 2002, documented some early failures and experiences [3]. The intent of
this report is to present some more recent experiences and representative examples in the form of
detailed case studies.

14166640 1-2
2
INDUSTRY EXPERIENCE—DAMAGE OR FAILURE IN
TUBING
2.1 Reported Failures
This section provides an updated table of reported failures and damage in T91 tubing. Table 2-1
lists documented failures, including those presented in the EPRI report Performance Review of
P/T91 Steels (1004516) [3].
Two general trends in the type of failures appear to be emerging: 1) short-term failures due to
tube fabrication or heat treatment issues that produce poor properties and 2) medium-term
failures that also might be related to less-than-optimum properties or to oxidation issues
producing tube blockage or poor heat transfer. The case studies provide examples of both types.

Table 2-1
Industry Experience with Tubing Failures

Case
Plant Source Type Description
Study
TVA Kingston EPRI report Superheater
No damage in 1996 (116,000 hours) —
U5 1004516 tubing
AEP Tanners EPRI report Superheater
No reported damage in 2002 —
Creek U3 1004516 tubing
Superheater Failure at 38,000 hours, creep failure in large-
Duke Power EPRI report
tubing, terminal radius bends, improper tempering resulted in —
Allen U3 1004516
tube degraded creep properties
Failure at 14,800 hours, short-term overheating
Hawaiian Secondary
EPRI report due to steam side oxides, or blockage or
Electric, Kahe superheater —
1004516 furnace modifications, improper heat treatment
U5 tubing
(intercritical region)
Superheater heat
EPRI report Early failures at header weld, dissimilar metal
Plant B recover steam —
1004516 weld with P22 header, no details on cause
generator tubing
United Kingdom
EPRI report No failures, heavy oxides and thin walls,
plant (not Reheater tubing —
1004516 thermal degradation
identified)
Japan (not EPRI report Superheater Failure at 63,000 hours, creep due to inner

identified) 1004516 tubes oxide scale
Presentation,
EPRI Boiler Failure at 15,114 hours, 13 in. (330.2 mm) from
Hong Kong Superheater
Tube Failure Grade 22 to Grade 91 transition, short-term —
Lamma U8 tubes
Conference, overheating from gas flow imbalance
Calgary, 2006

14166640 2-1
Table 2-1 (continued)
Industry Experience with Tubing Failures

Case
Plant Source Type Description
Study
Superheater
RWE, West EPRI report Failure at 20,000 hours, Type IV at weld,
transition —
Burton 1004516 possible overtempering
bottle
Japan (not EPRI report Superheater Failure at 18,000 hours and 40,000 hours,

identified) 1004516 and reheater low hardness due to tempering
J. Arnold
Unidentified presentation to Stress corrosion cracking at tube bends
Reheater tubes 1
subcritical plant EPRI workshop, during fabrication
2003
J. Parker
Thin-lipped fishmouth failure related to
presentation to
Unidentified Reheater tube oxide scale issues, exfoliation and blockage, —
EPRI workshop,
or excess scale producing overheating
2008
Consumers
MAT Project Superheater Failure at 300 hours, short-term overheating
Energy, —
0700278, July 2007 tube due to slag debris blockage
Campbell U3
J. Alice, J. Henry Failure at 38,000 hours; fishmouth failure
Superheater
Unidentified presentation, creep; tempering temperature might be too —
terminal tube
April 2008 high, leading to reduced creep life
Secondary
Draft EPRI report, Failure at 61,000 hours, long-term
Glen Lyn U6 superheater 4
2008 overheating creep due to oxide scale
tubing
P.W. Woo Failure at 15,000 hours, short-term
Hong Kong Superheater
presentation, overheating, gas flow problems due to oxide 3
Lamma Unit 8 tubing
October 2007 blockage
26,000 hours, multiple fishmouth failures at
KEPRI Secondary
bends and next to welds, hardness at
Yung Heong #1 presentation, superheater 2
153 Vickers hardness (HV), suspect
January 2008 tubing
postweld heat treatment
6400 hours, Grade 92, bends and welds,
KEPRI, G Jung
severe swelling, degraded ferritic
Tae An presentation, Reheat tubing —
microstructure, creep voids, hardness at
January 2008
117 HV, suspect manufacturing damage
Failures in replacements tubes for T22. Type
ESB ETD Seminar, June IV failure after 16,000 hours in tube-to-
Reheater tubes 5
Moneypoint GS 20–21, 2007 header weld, significant degradation in gas
path tubes after 20,000–30.000 hours

14166640 2-2
2.2 Case Studies–Fabrication-Related Damage

2.2.1 Case Study 1


This study provides an example of how excessive cold work or untempered martensite can
produce early stress corrosion cracking in tube bends [4].

2.2.2.1 Construction Details and History


The subject tubes were intended as replacements for a reheater in a coal-fired, subcritical utility
boiler. The 2.125 in. (53.975 mm) outside diameter (OD) × 0.0148 in. (0.37592 mm) minimum
wall thickness (MWT) tubes were fabricated in SA-213 T91 with a 90° bend. The bends were
produced by cold bending to a 7.5 in. (190.5 mm) radius (Figure 2-1). Cracks were detected in
the bend region before the tubes entered service.

Figure 2-1
Fabricated 90° Tube Bends

2.2.2.2 Crack Characteristics


Circumferentially oriented cracks were detected with the bend region close to the start of the
bent portion on the inside radius (Figure 2-2). On a macro scale, the cracks were straight,
singular, and centered on the bend intrados. Metallographic sections of a through-wall crack
showed a slightly branched but relatively straight crack path (Figure 2-3).

14166640 2-3
Figure 2-2
Appearance of Cracks on Inner Radius

Figure 2-3
Metallographic Section of a Through-Wall Portion of Crack

Heavy oxidation of the fracture surfaces was present. Sections of a nonpenetrating portion
(Figure 2-4) revealed the crack features more clearly—more obvious branching and an
intergranular crack path, apparently following prior austenite boundaries in the martensitic
microstructure. These crack features are typical of a stress corrosion cracking mechanism.
Further investigation revealed that the tubing had been swaged and normalized but then left in
the hardened condition before bending and subsequent storage. Tempering was intentionally not
to be done until after welding into assemblies as part of the postweld heat treatment (PWHT)
process.

14166640 2-4
Figure 2-4
Metallographic Details of Non–Through-Wall Portion

2.2.2.3 Conclusions
The following conclusions were reached:
• Cracking in the tube bends was caused by a stress corrosion cracking mechanism.
• The tubing material had been left in a vulnerable, hardened condition because tempering had
been delayed until after welding. In addition, the bending process must have introduced
residual stresses.
• The environmental factor was provided by the normally benign, ambient storage conditions
for 6 weeks.

14166640 2-5
2.2.2 Case Study 2
This case study provides an example of a tube failure that might be related to initial fabrication
heat treatment.

2.2.2.1 Construction Details and History


The failure was in a Korean power station, Yung Heong #1, which is an 800-MW supercritical
unit. Main steam temperature was 1056°F (569°C). Several superheater tube failures occurred
over a period of 2 years, with an increase in the number of events in 2007, after about 26,000
hours (Figure 2-5). The material was SA-213 T91 tubing of 1.68 in. (42.7 mm) OD and 0.29 in.
(7.4 mm) thickness [5].

Figure 2-5
Failure Locations in Superheater T91 Tubing

2.2.2.2 Failure Characteristics


Several short fishmouth failures occurred at the bottom of bends (Figure 2-6a). Some had fairly
limited extents and were accompanied by local swelling that is characteristic of blister-type
failure (Figure 2-6c). Others were longer in axial extent, and some were in the straight sections
(Figure 2-6b). One failure was associated with a weld and featured local swelling with a thin-
edged fracture (Figure 2-6d and Figure 2-7). There was also some evidence of oxide exfoliation
in the area. Hardness in the sections away from the failure site averaged 188–220 Vickers
hardness (HV), but it was 159 HV in the area shown in Figure 2-7.

14166640 2-6
a b

Figure 2-6
Features of Several Tube Failures

14166640 2-7
Figure 2-7
Failure Site Adjacent to Weld, Showing Local Swelling and Thin-Edge Fracture

Microstructures in most of the tube were tempered martensite, but close to the failure site, they
were ferrite and large carbides (Figure 2-8).

Figure 2-8
Comparison of Microstructures Away from and at the Fracture Site

14166640 2-8
Oxide measurements in the failed and non-failed sections showed fracturing and delamination at
the failure site compared to the normal multilayered oxides (Figure 2-9). Oxide thickness was
5.9–10.6 mil (150–270 µm).

Figure 2-9
Comparison of Oxide Scale in Intact Section and at Failure Site

2.2.2.3 Conclusions
The following conclusions were reached:
• The failure mode in the tube was characterized as long-term overheating, possibly linked to
oxide scale exfoliation.
• The microstructure and hardness of the area close to the weld provides evidence that the weld
did not receive proper PWHT. The temperature must have exceeded the Ac1 temperature in
order to produce the ferrite–carbide microstructure with low hardness and poor creep
properties.

2.3 Case Studies—Oxidation-Related Failures

2.3.1 Case Study 3


This case study provides another example of a tube failure related to oxidation—an accumulation
of exfoliated steam side scale that produced tube blockage and overheating.

2.3.1.1 Construction Details and History


The failure was in the Hong Kong Electric Lamma power station, which is a 350-MW coal-fired
unit with a forced circulation drum reheat boiler. Main steam and reheat steam temperatures
were 1056°F (569°C) and 1005°F (541°C), respectively. The failures occurred after 15,114 hours
and 41 start–stop cycles [6].

14166640 2-9
2.3.1.2 Failure Characteristics
The failure sites were located in the tight radius bends in the lower portion of the fourth
superheater (Figure 2-10). One tube had ruptured and another displayed swelling in a similar
location. (Figure 2-11). The rupture site was localized and had the features of short-term
overheating (Figure 2-11).

Figure 2-10
Failure Location in Fourth Superheater T91 Tubing

14166640 2-10
Figure 2-11
Features of Primary Failure Site

The tube midwall hardness was satisfactory (195–214 HV) except at the failure site, where it had
decreased to 150 HV. No evidence was found of extensive overheating or creep life consumption
in the general tube microstructure.
Examination of the inside surfaces of the tube showed evidence of delamination of the steam-
side scale. Accumulations of spalled oxides were found in the bottom of the tube bends in this
unit and in other similar locations in a sister unit (Figure 2-12).

Figure 2-12
Characteristics of Exfoliated Steam-Side Oxide Scale

14166640 2-11
Oxide thicknesses were measured in excess of 11.8 mil (300 µm). The spalled oxide retained the
tube curvature and measured 1.18 × 0.39 in. (30 × 10 mm) maximum. Cross sections of the
oxides showed them to be multilayered, consisting of spinel inner layers and magnetite outer
layers (Figure 2-13). Gaps were apparent between the spinel and magnetite layers, suggesting
that this was the source of the delamination.

Figure 2-13
Features of Multilayered Oxide with a Spinel Inner Layer, a Magnetite Outer Layer, and Evidence of
Separation of the Layers

2.3.1.3 Conclusions
The following conclusions were reached:
• The failure mode in the tube is characteristic of short-term overheating.
• Exfoliation of the multilayered steam-side oxides have produced blockage or restrictions in
the tube and produced high local metal temperatures.
• The thick oxides and the gaps between the layers suggest that high thermal strains in the
oxides could have been produced by thermal shocks to the boiler by forced cooling or
start-up rates.

14166640 2-12
2.3.2 Case Study 4
This study is an example of the way in which oxidation of T91 tubing can possibly lead to
medium-term overheating creep failures [7].

2.3.2.1 Construction Details and History


Glen Lyn Unit 6 is a 2400-psig (16.65-MPa) pulverized coal-fired subcritical boiler with natural
circulation; it went into service in 1957. The unit is a 240-MW radiant boiler with a single reheat
section. The design conditions for main steam are 2075 psig (14.31 MPa), with superheat and
reheat steam temperatures of 1050°F (565.6°C). Before the failure, the unit operated at a
superheater outlet pressure of 2,000 psi (13.79 MPa) and a bulk outlet temperature of 1050°F
(565.6°C). The secondary superheater consists of 66 elements (33 elements to the west and 33 to
the east side of the division wall). Four parallel tube circuits create 24 rows (6 passes × 4 tube
circuits).
Historical review revealed that the secondary superheater section has been replaced twice since
1957. The original 2 in. (50.8 mm) OD × 0.260 in. (6.6 mm) MWT type 321 stainless steel in the
secondary superheater was upgraded to a 2 in. (50.8 mm) OD × 0.284 in. (7.21 mm) MWT 304H
stainless steel in 1969, after 12 years of service. After an additional 26 years of service, the 304H
stainless steel was replaced with 2 in. (50.8 mm) OD × 0.285 in. (7.24 mm) MWT T91 in the
spring of 1995.
In August 2003, a failure occurred in the superheater, approximately 2 ft (0.61 m) below the roof
on the outlet tube leg (Figure 2-14). The failure occurred in element number 26 of the 66
elements (counted from the right side of the unit while looking toward the rear of the unit) in the
second of the four tube rows (that is, the sixth tube from the front of the secondary superheater)
of the final secondary superheater outlet leg leading to the outlet header. The failure occurred
just beneath the top alignment bar. Visual examination of the superheater revealed that the tube
circuits, including the failed circuit, were in excellent alignment and showed no evidence of any
bowing or distortion. The T91 failure occurred after approximately 8 years (61,000 hours) of
service.

14166640 2-13
Figure 2-14
Failure in T91 Tubing, Just Beneath the Top Alignment Bar

2.3.2.2 Crack Characteristics


Laboratory metallurgical investigations revealed that the failed tube exhibited a thick-lipped,
fishmouth rupture with severe intergranular creep cavitation across the entire wall thickness
(Figure 2-15).

Figure 2-15
Metallurgical Section at Failure Site, Showing Creep Cavitation

14166640 2-14
The tube exhibited the thickest steam-side oxide scale (0.029 in. [0.737 mm]), the highest density
of creep cavities (2 per in2 [1260 per mm2]), the thinnest wall thickness (0.29 in. [7.37 mm]), and
the lowest midwall hardness (83.6 Rockwell B) at the location of the rupture. The microstructure
of the tube consisted of tempered martensite. No evidence of intercritical overheating or
improper heat treatment of the tube was observed. The chemistry of the tubing was within the
limits for SA-213 T91 material, although the carbon and chromium contents were near the low
end of the required range. The thick steam-side oxide scale was composed of a complex, duplex
scale with the following layers (Figure 2-16):
• A very thin, irregular inner chromium-rich layer that penetrates into the tube metal at the
oxide scale–metal interface
• A thick multilayered scale that probably is a combination of iron-chromium spinel and
magnetite and small amounts of ferrite and wustite
• A thick chromium-free magnetite layer
• A thin hematite layer at the steam interface
There was significant porosity, delamination, and blistering of the steam-side scale, raising
concern regarding its effective thermal conductivity and its effect on steam cooling of the tubing.
The delamination tended to occur along porosity located at the spinel–magnetite interface. The
blistering was observed only in the outer magnetite–hematite layers, suggesting that these layers
have significant compressive growth strains as the oxide thickens.

14166640 2-15
Figure 2-16
Characteristics of Steam-Side Oxide Scale Near Rupture, Showing Evidence of Delamination

14166640 2-16
2.3.2.3 Conclusions
The following conclusions were reached:
• The failure mode is characteristic of long-term overheating creep.
• The root cause of the failure has not been established, but it is thought to be related to the
insulating effect of the thick oxide layer, leading to an increase in effective metal temperature
(estimated as 1230–1240°F [665.6–671.1°C]). This overheating compromised the excellent
long-term creep properties of the T91 material and produced the premature failure.
2.3.3 Case Study 5
This study is an example of a Type IV failure in a stub tube weld and severe degradation of T91
tubing in the form of softening and excessive oxidation [8].

2.3.3.1 Construction Details and History


Moneypoint GS is operated by the Electricity Supply Board of Ireland. The station has an
installed capacity of 915 MWe, with three 305-MWe units that have steam turbines and coal-
fired drum boilers. The nominal rating for the main steam is 2393 psi (16.5 MPa) at 1000°F
(540°C), with reheat steam at 609 psi (4.2 MPa) and 1000°F (540 °C).
The upper sections of the reheaters were originally tubed in SA-213 T22. Early sampling from
the original T22 tubes found unexpected levels of degradation and oxide growth, indicating
exposure to temperatures that were higher than design temperatures. The root cause investigation
of this degradation found an uneven temperature profile across the boiler at the outlet. Some of
the tubes were experiencing steam temperatures of 1060–1110°F (570–600°C) at the outlet
header. Tube metal temperatures in excess of the steam temperatures were expected in the flue
gas stream. In 1998, 16 of the 47 upper panel sections were replaced with T91 (BS3059PT2
S291). At the same time, the reheat outlet header was replaced with P91 material. The revised
design temperatures for the replacements were 1190°F (645°C) for the tube panels and 1148°F
(620°C) for the header.

2.3.3.2 Crack and Damage Characteristics


After 30,000 hours of operation on the replacement parts, some cracks were discovered in the
tube-to-header welds (Figure 2-17). Analysis of the failure showed the damage to be in the soft,
intercritical region of the heat-affected zone (HAZ)—a type IV failure. Excessive system stresses
were identified as important in this premature failure. Appropriate modifications were made to
reduce the stresses, and no further failures have been reported. A significant conclusion was that
the P91 material might be more susceptible to the type IV failure mode than the original P22
material, which had experienced no failures in this location.

14166640 2-17
Figure 2-17
Failed Tube-to-Header Weld

The tubing in the gas path was assessed after 20,000–30,000 hours, and significant thermal
degradation in the form of a reduction of average hardness (from 220 HV to 180 HV) and
indications of internal oxide scale thicknesses of 11 mil (280 µm) were found. In situ
metallography on some tubes also detected changes in the tube microstructure. Figure 2-18
shows the microstructure in a tube in the dead space (unaffected by temperature), which consists
of the expected tempered martensite and fine carbides.

Figure 2-18
Microstructure of Tubing in Dead Space After 24,000 Hours (x500)

14166640 2-18
Figure 2-19 shows the microstructure in the heated portion of the tube, showing significant
growth of the carbides. A tube sample taken after 34,000 hours showed further coarsening of the
carbides and, significantly, the appearance of general creep cavitation (Figure 2-20).

Figure 2-19
Microstructure of Tubing in Gas Path After 24,000 hours (x 500)

Figure 2-20
Microstructure of Tubing After 34,000 Hours (x200)

14166640 2-19
Examination of the oxides on the removed tubes identified that excessive oxidation and spalling
might be an issue with the T91 replacement tubing (Figures 2-21 and 2-22). Oxide measurements
seemed to be related to service hours as well as location; they varied from 13.8 to 27.6 mil
(350 to 700μm).

Figure 2-21
Internal Oxide Spalling (Oxide Thickness, 650 μm)

The visual appearance and degree of oxide spalling were more related to tube position than to its
thickness. Higher levels of spalling were observed in tubes that were known to run hot. The
oxide structure was layered with three distinct structures (Figure 2-22). It appears that the
outside, porous layer is separating from the more dense inner layers, possibly due to differences
in thermal expansion behavior. The spalling can have detrimental effects on downstream
components such as valves, drains, and turbines.

14166640 2-20
Figure 2-22
Internal Oxide on Tube After 24,000 Hours (x500)

2.3.3.3 Conclusions
The following conclusions were reached:
• After 34,000 hours of operation, the replacement reheater tubes displayed degradation of
microstructure, reduction in hardness, and creep cavitation. In addition, heavy internal
oxidation and oxide spalling were observed.
• The degradation of the T91was considered severe enough that selective replacement of the
tubes in the critical areas was recommended.

2.4 Solutions
Careful attention must be paid to tube fabrication procedures, particularly heat treatment issues.
Tempering after normalizing should not be delayed, because the as-hardened material is
vulnerable to stress corrosion in normally benign ambient environments. If the tubes must be
stored for lengthy periods in this condition, they should be in temperature- and humidity-
controlled facilities.
Careful control of tempering temperatures during manufacturing and PWHT is also required,
because temperatures exceeding Ac1 can result in ferritic microstructures and inferior creep
properties. If tempering irregularities are suspected, hardness testing and replication can be used
as inspection tools.
The solution for the overheating failures is not so obvious, because some failures appear to be
related to the inherent characteristics of the oxide layers. Oxidation rates of the Grade 91 alloy in
steam appear to be higher than originally anticipated. This higher rate produces thick, insulating

14166640 2-21
oxides and higher rates of wall thinning. The steam-side oxides on T91 are multilayered and
might be prone to interlayer porosity or cracking defects. Exfoliation of these defective oxides
can also lead to tube blockage and overheating. Tube bends can be monitored by radiographic
inspection to detect potential blockages before failure occurs. The long-term solution might be
replacement tubes made from more oxidation-resistant type 304 or 316 alloys.

14166640 2-22
3
INDUSTRY EXPERIENCE—BASE METAL DAMAGE OR
FAILURE IN HEAVY SECTION COMPONENTS, SHORT
TERM
3.1 Reported Failures
Several failures have been reported in Grade 91 heavy section components early in the plant life
(less than 40,000 hours) or even before entering service. Several of these early failures appear to
have been caused by fabrication issues such as incorrect or omitted heat treatments. Table 3-1
provides a list of reported short-term heavy section failures and damage.

Table 3-1
Industry Experience with Short-Term, Heavy Section Base Metal Failure

Plant Source Type Description Case


Study
United Kingdom EPRI report Stub-to-header Failure at 16,000 hours, Type IV at —
power plant 1004516 weld stub-to-header weld due to high local
strains from attachment
RWE, West EPRI report Superheater Failure at 20,000 hours, Type IV at —
Burton 1004516 transition bottle weld, possible overtempering
RWE Allen, Brett, Header endplate Failure at 36,000 hours, Type IV 10
Symposium, Milan, cracking
Italy, 1999
Unidentified J. Parker presentation Hot reheat bypass Low cycle thermal fatigue from —
to EPRI workshop, line failure attemperator sprays
2008
Southern M. Sims presentation Hot reheat piping Failure in June 2005 at 5400 hours, 7
Company to EPRI workshop, fabrication flaws after forming before
April 2008 normalizing
Progress Energy Progress Energy Hot bent piping Stress corrosion cracking due to 8
presentation, EPMC, outside storage before heat treatment
June 2006
Salt River SRP presentation, Hot reheat piping Type IV, 18,000 hours 9
Project, Kyrene Clearwater meeting Weldolet fitting
U7
Salt River SRP presentation, High-pressure Thermal fatigue cracks, 18,000 hours 11
Project, Kyrene Clearwater meeting steam bypass
U7 valves

14166640 3-1
Table 3-1 (continued)
Industry Experience with Short-Term, Heavy Section Base Metal Failure

Plant Source Type Description Case


Study
Dominion Thielsch report, Hot reheat piping Axial crack before hydro test, heat 6
Fairless, January 2004 treatment
combined-cycle
plant
Unidentified, J. Hald, TU Denmark, Interconnect pipe New construction, hard regions due —
new plant presentation to IIW to heat treatment after bending
Comm. IX-C
Intermediate meeting,
Arnhem, NL, May
2008

3.2 Case Studies—Fabrication-Related Failures

3.2.1 Case Study 6


This study provides an example of the way in which heat treatment issues or site erection issues
can cause cracking. A 13-in. (330.2-mm) axial crack was discovered in a 24-in. (609.6-mm) hot
reheat line before hydro testing before the plant was commissioned [9, 10].

3.2.1.1 Construction Details and History


This failure occurred in a 550-MW combined-cycle plant in the hot reheat piping system. The
24-in. (609.6-mm) diameter × 0.688-in. (17.475-mm) thickness pipe was specified as ASTM
A335 P91 and designed to ASME B31.1 requirements for a temperature of 1050°F (565°C) and
a pressure of 650 psi (4.48 MPa). Before fabrication, the pipe section had been hydro tested and
inspected with ultrasonic testing, and it was found to be leak free. The section was field welded
to a bend, using appropriate filler metal and a combination of tungsten inert gas and shielded
metal arc welding procedures. Records showed that the weld had been manufactured according
to the procedure, with the required 450°F (232°C) preheat, followed by postweld bake-out at
650°F (343°C) and PWHT for 2 hours at 1380°F (750°C). Radiography and inspection after the
welding and PWHT had not revealed the existence of the crack, although subsequent review of
the radiograph indicated that the tip of the crack might have appeared on the radiographic image
but misinterpreted. During plant commissioning, the section of piping was being filled for hydro
testing when a leak was detected and traced to the linear defect adjacent to a girth weld
(Figure 3-1).

14166640 3-2
Figure 3-1
Location of Linear Defect in Hot Reheat Pipe

3.2.1.2 Crack Characteristics


Removal of the pipe section and close-up examination showed that the defect was a 13-in.
(330.2-mm) crack located in a horizontal section of piping at the 5 o’clock position (Figures 3-2
and 3-3). One end of the crack was about 2 in. (50.8 mm) from the girth weld (Figures 3-1 and
3-2). Fractographic examination indicated that the crack had initiated on the outside surface, had
a brittle appearance, and was oxide filled, indicating exposure to high temperatures after it had
been formed (Figure 3-4).

Figure 3-2
Sample Excised for Examination, Showing Location of Section A-A Relative to Edge of Weld

14166640 3-3
Figure 3-3
Close-Up of Crack on Outside Diameter Surface

Figure 3-4
Fracture Surfaces in Center of Crack

14166640 3-4
The metallurgical investigation established that the pipe material was correct and generally met
the requirements of the specification in terms of hardness, tensile properties, and microstructure,
except that the crack was located in the center of a hard zone of material. The hardness in a
section 180° opposite the crack location was 199–207 Brinell hardness (BHN), whereas in
section A-A (Figure 3-5) the hardness was 347–377 BHN. This section is only about 3 in.
(76.2 mm) from the edge of the girth weld. The microstructure in the crack area was also
consistent with a high-temperature excursion, displaying a larger grain size and untempered
martensite (Figures 3-5 and 3-6). In contrast, the remainder of the pipe, away from the crack
area, had a much finer grain size and a tempered martensite structure (Figure 3-7).

Figure 3-5
Metallographic Section of Crack at A-A

14166640 3-5
Figure 3-6
Microstructure and Crack Features at A-A

14166640 3-6
Figure 3-7
Typical Microstructure in Pipe (Section Opposite Side to Crack)

14166640 3-7
3.2.1.3 Conclusions
The following conclusions were reached:
• The local area containing the axial through-wall crack had been heated above the upper
critical temperature, leading to large grain size and martensite production. Material hardness
of greater than 347 BHN showed that this martensite had not been tempered.
• The precise timing and cause of this overheating event was not established by the
investigation, but it was thought to have occurred after the original normalizing and
tempering treatments for the pipe spool and before the welding and PWHT of the adjacent
girth weld. However, it might have also have occurred either during or after the PWHT
because the PWHT should have tempered at least the hard zone that was closest to the girth
weld. Incorrect placement or control of the PWHT pads could be the source of the
overheating.
• Regardless of the exact cause, this crack event illustrates the care needed during any heat
treatment of Grade 91 components to ensure that the critical temperatures are not exceeded
and production of hard brittle martensite is prevented.
3.2.2 Case Study 7
This case study provides an example of an undetected flaw that developed during fabrication.
Axial cracking occurred in a 30-in. (762-mm) diameter hot reheat pipe in a combined-cycle plant
after 5400 hours of operation [11].

3.2.2.1 Construction Details and History


This failure is from a 2-on-1 combined-cycle unit with two 172-MW combustion turbines, two
heat recovery steam generators, and one 193-MW steam turbine. The design temperature is
1075°F (580°C), and the design pressure is 550 psig (3.79 MPa). Commercial operation began in
June 2003, and the failure was first detected in June 2005, after 5479 hours of operation and 194
starts. The failure site was in the 30-in. (762 mm) OD × 0.751-in. (19.075-mm) MWT hot reheat
piping, which was made from A-335 Grade P91 seamless pipe (Figure 3-8).

14166640 3-8
Figure 3-8
Location of Cracks in the Hot Reheat System

3.2.2.2 Crack Characteristics


The initial leak was discovered in 2005 and was traced to a 22-in. (558.8-mm) long axial crack in
a straight pipe spool (Figure 3-9). This crack was not investigated, but it was repaired by
grinding and local weld repair.

14166640 3-9
Figure 3-9
Appearance of Initial Crack

The entire pipe spool was removed later, at which time a metallurgical examination was
conducted. A second non–through-wall crack was discovered 7 in. (177.8 mm) upstream from
the original leak. This straight axial crack had initiated from the inside diameter (ID) surface and
progressed to 85% of the wall (Figure 3-10). The crack path was in a mostly integranular mode,
except for a small fatigue area at the crack tip (Figure 3-11).

14166640 3-10
Figure 3-10
Sections of Non–Through-Wall Crack

Figure 3-11
Fractography at Crack Tip

14166640 3-11
The microstructure at the crack location and elsewhere was satisfactory, indicative of a
normalized and tempered martensite (Figure 3-12). Hardness and tensile properties were also
normal and within the specification requirements. A review of manufacturing procedures,
including the shop and field welding, detected no abnormalities that could have caused or
influenced the cracking.

Figure 3-12
Typical Microstructures Near Crack

3.2.2.3 Conclusions
The following conclusions were reached:
• The results of the investigation suggested that the flaws were present during manufacture,
before shop fabrication or installation. The profile of the crack surface suggests that it
occurred when the grain size was large—that is, before the normalizing and tempering
operation.
• The intergranular fracture mode suggests that stress corrosion cracking or a hydrogen-
induced cracking mechanism was responsible. Presumably, this might have occurred from
atmospheric exposure after pipe forming but before heat treating. Operational in-service
fatigue cycling had propagated the flaws by a small amount.
• The subsequent heat treatment and welding processing had produced a satisfactory
microstructure and mechanical properties in the entire pipe, including the cracked areas.
• Shop weld and field welds were not close enough to have imposed significant residual
stresses or microstructural changes that could have influenced the cracking.

14166640 3-12
• Future procurements of P91 will specify 100% ultrasonic examination to detect such
pre-existing flaws.
• This case study provides another example of the ways in which flaws in Grade 91 can
develop during fabrication and also how such flaws can by caused by stress corrosion or
hydrogen in untempered martensite microstructures.
3.2.3 Case Study 8
This study provides another example of the ways in which stress corrosion cracks can develop in
thick section components during fabrication if tempering or PWHT is not performed in a timely
fashion [12].

3.2.3.1 Construction Details and History


This example is from Progress Energy field experience with Grade 91 component assemblies.
The bending process for thick section bends involves heating a band of the pipe with an
induction coil as the pipe moves through the required arc (Figure 3-13). With the temperatures
(1470–2010°F [800–1100°C]) induced under the coil, the bend region is heated into the
normalizing range and should be in the hardened condition after cooling. Regions just outside the
bend can be heated only into the transition region and might not be fully martensitic. The entire
pipe spool should be completely re-normalized and tempered after bending to produce uniform
microstructure and mechanical properties.

Figure 3-13
Induction Hot Bending of Pipe

14166640 3-13
3.2.3.2 Crack Characterisitics
Initially, hot-bent pipe was being stored outside before the final normalizing and tempering heat
treatment (Figure 3-14). This was also true of welded assemblies before final PWHT. It was
discovered that moisture from condensation, dew, rain, and lubricants can cause stress corrosion
cracking of hot-bent and as-welded Grade 91 steels.

Figure 3-14
Outside Storage of In-Progress Pipe Spools

3.2.3.3 Conclusions
The following conclusions were reached:
• In the hot-bent and as-welded condition, before stress relief, the pipe is susceptible to stress
corrosion cracking in moist environments.
• Hot-bent pipe and welded assemblies should be stored in low-moisture facilities
(Figure 3-15) until the final heat treatment.

Figure 3-15
Temporary Heated Storage Facility for Component Assemblies

14166640 3-14
3.3 Case Studies—Design-Related Failures

3.3.1 Case Study 9


This example illustrates that poor design can lead to problems in Grade 91, specifically with the
use of Weldolet type fittings at branch connections [13, 14].

3.3.1.1 Construction Details and History


Kyrene Unit 7 is a 250-MW combined-cycle plant that began operation in October 2002. It
consists of one combustion turbine, one heat recovery steam generator, and one steam turbine.
The main steam system operating conditions are approximately 1050°F (565°C) and 1800 psig
(12.4MPa), and those of the hot reheat system are approximately 1050°F (565°) and 400 psig
(2.75MPa). The first Grade 91 high-energy piping inspection took place in May 2006, when the
unit had about 18,000 operating hours and 800 starts.

3.3.1.2 Crack Characteristics


The area inspected was an 8- × 14-in. (203.2- × 355.6-mm) Weldolet fitting that connected an
18-in. (457.2 mm) pipe (leading to the east cold reheat valve) to a 14-in. (355.6-mm) pipe
(Figure 3-16). A 9-in. (228.6-mm) crack was found along the toe of the weld on the 18-in.
(457.2-mm) pipe side at about the 12 o’clock position (Figure 3-17).

Figure 3-16
Weldolet Fitting in Hot Reheat Main Pipe

14166640 3-15
Figure 3-17
Crack at Edge of Weld

Replicas showed the cracking to be associated with extensive creep cavitation in the coarse-
grained and fine-grained HAZs in the 18-in. (457.2-mm) pipe (Figure 3-18). Shear wave
ultrasonics determined the depth of the cracking to be about halfway through the wall.

Figure 3-18
Replica Metallograph, Showing Extensive Creep Cavitation in the Coarse-Grained Heat-Affected
Zone

14166640 3-16
Boat samples were removed for subsequent metallurgical evaluation, which confirmed the
presence of creep cavitation associated with the crack along its depth (Figure 3-19). The cracking
was attributed to Type IV creep damage in the inherently soft, fine-grained region of the HAZ
(Figure 3-20). The fact that the cracking appeared after only 18,000 hours also suggests that this
weld might be highly stressed in addition to the stress concentration induced by the geometry.
The entire saddle weld was ground out and replaced; however, the Weldolet fitting will
eventually be replaced with a forged tee piece.

Figure 3-19
Metallographic Section from Boat Sample from Weldolet Fitting, Also Showing Creep Cavitation

Figure 3-20
Hardness Profile Across a Typical Weld

14166640 3-17
3.3.1.3 Conclusions
The following conclusions were reached:
• Weldolet type fittings in Grade 91 will be prone to early Type IV creep cracking because of
the high stress concentrations produced at the welds.
• Weldolet type fittings should be inspected for early creep damage or avoided in new
construction.
3.3.2 Case Study 10
This study describes header endplate creep cracking that occurred early in life due to geometric
stress concentrations close to a weld [15, 16].

3.3.2.1 Construction Details and History


Due to ligament cracking, secondary superheater outlet headers had been replaced with Grade 91
components in the early 1990s. The failed header had been replaced in 1991, and the cracks
appeared after 36,500 hours at 1055°F (568°C), after 469 hot starts and 72 cold starts. The design
of the endplate was a recessed flat plate, with the root of the endplate weld close to the internal
corner (Figure 3-21). This design obviously produces an increased stress on the weld geometry;
however, the design was acceptable to BS1113 code.

Figure 3-21
Geometry of the Header Endplate

14166640 3-18
One of the headers failed catastrophically in 1996, with the weld failing completely, resulting in
ejection of the endplate and depressurization of the boiler (Figure 3-22). After the failure,
ultrasonic inspection discovered other damaged endplate welds.

Figure 3-22
Failed Header Endplate

14166640 3-19
3.3.2.2 Crack Characteristics
The fracture path was along the edge of the weld, on the endplate side of the weld, following the
Type IV region in the parent material (Figure 3-23). On a microscopic scale, creep cavitation
was found associated with the entire depth of the crack. The densest cavitation was found just
beneath the bore surface, indicating that the cracking had initiated there and had progressed first
to the inside surface and then to the outer surface. No cavitation was found in the weld or in the
HAZ on the header side of weld. These features identify the mechanism as Type IV creep
cracking.

Figure 3-23
Metallographic Section of Failed Weld

Microhardness surveys of the failed header and other similar headers found that the forged parent
material had entered service with hardness lower than expected and lower than that of the header
material. This low hardness was also reflected in lower hardness in the intercritical HAZ, where
the creep damage had initiated.
Detailed stress analysis of the endplate geometry showed that stresses could be 2–3 times the
allowable stress in the failure location. Although this is a less-than-optimal design, it has been
used successfully in other applications with low-alloy steel (2.25Cr-1Mo and 1.25Cr-0.5Mo).
Failures have occurred in these steels, but only after much longer operating times than in the
present case. It was concluded that the endplate geometry was a factor, but it could not be
responsible in isolation for the failure. The low hardness of material was also considered to be an
important factor in this failure.

14166640 3-20
3.3.2.3 Conclusions
The accelerated rate of damage accumulation in these Grade 91 components was attributed to the
following:
• Material that had been subjected to preservice heat treatment, which resulted in hardness and
strength values below the expected ranges. The low strength was also reflected in the
strength of the HAZ. This material effect was probably influenced by a low nitrogen–
aluminum ratio in the failed components
• The location of the weld and the selected component geometry, although it complied with
applicable codes, resulted in high local stress concentrations.
The components were repaired and then replaced at the next convenient outage. Inspection
methods were improved to permit greater reliability in damage detection for subsurface cracking.

3.4 Case Studies—Operation-Related Failures

3.4.1 Case Study 11


This study provides an example of the ways in which operation issues, such as thermal cycles
induced by attemperation in high-temperature systems, can lead to early damage [13, 14, 17].

3.4.1.1
This example is from Kyrent Unit 7, the same 250-MW combined-cycle plant that was described
in case study 8. The main steam system operating conditions are approximately 1050°F (565°C)
and 1800 psig (12.4 MPa), and those of the hot reheat system are approximately 1050°F (565°C)
and 400 psig (2.75 MPa).
The unit began operation in October 2002, and the first Grade 91 high energy piping inspection
took place in May 2006, at which time the unit had about 18,000 operating hours and 800 starts.
A second inspection was performed in November 2007, at which time the unit had 24,000
operating hours and 1185 starts. The high-pressure steam bypass control valve contained a
dissimilar metal weld (P91 to P22) on the outlet side, just downstream of the attemperator spray
nozzles (Figure 3-24).

14166640 3-21
Figure 3-24
High-Pressure Bypass Valves, Showing Location of the Dissimilar Metal Weld Relative to the
Attemperator Nozzles

3.4.1.2 Crack Characteristics


The circumferential weld adjacent to the attemperator spray nozzles was the primary focus of the
inspection, and it was first examined internally using a videoprobe. During the first inspection,
crack-like indications were found along the downstream weld toe (P22 side) and at the
counterbore corner on the P91 side of the weld. No indications were found on the OD by wet
fluorescence magnetic particle testing. In the second inspection, it was discovered that the
internal cracks at the circumferential weld had grown in length and depth (Figure 3-25).

14166640 3-22
Figure 3-25
Crack Indications on the Inside Surface at the Weld Toe and Counterbore

Shear wave ultrasonic inspection indicated that the cracks at the downstream weld toe were
halfway through the wall. A section was removed for metallurgical evaluation (Figure 3-26).
Crack features were revealed as multiple, slightly branched, and transgranular (Figure 3-27).
These features are indicative of a thermal fatigue mechanism.

Figure 3-26
Metallographic Section of the Dissimilar Metal Weld, Showing Thermal Fatigue Cracks

14166640 3-23
Figure 3-27
Microstructure and Crack Features in the P22 Side

3.4.1.3 Conclusions
The following conclusions were reached:
• The bore surface cracks in the turbine steam bypass valves were caused by a thermal fatigue
mechanism from the operation of the attemperator sprays.
• The attemperator sections of such high-pressure bypass valves are vulnerable to this
mechanism and should be designed with a sacrificial liner.

3.5 Solutions
Similar to the fabrication of tubing, the processing of thick section Grade 91 components
requires careful attention. The normalizing step in the heat treatment should be followed as soon
as possible by the tempering operation in order to prevent stress corrosion cracking. If storage is
necessary, it should be in a controlled environment. PWHT should also be performed as soon
after welding as possible. Pipe bends should undergo a full normalizing and tempering treatment
to produce the correct microstructure in all regions of the bend. High-sensitivity ultrasonic
inspection of fabricated assemblies should be considered to ensure detection of processing flaws
before final assembly.
Weldolet type fittings should be avoided in design and should be replaced with forgings in
existing plants. Careful attention should be paid to areas that place welds close to areas of high
stress concentrations, such as flat ends on headers. Inspection programs should focus on such
high stress concentration areas. Areas close to attemperator nozzles that are prone to thermal
fatigue should also be inspected closely. If problems are encountered, consideration should be
given to replacement with designs that incorporate liners.

14166640 3-24
4
INDUSTRY EXPERIENCE—BASE METAL DAMAGE OR
FAILURE IN HEAVY SECTION COMPONENTS,
MEDIUM TERM
4.1 Reported Failures
Table 4-1 lists the known examples of reported medium-term (40,000–80,000 hours) heavy
section failures and damage. Several medium-term failures or cracks have now been reported.
Several utilities are also discovering components with hardness and strength values that are less
than the recommended values. These results have raised concerns about the long-term
performance of the components.

Table 4-1
Industry Experience with Medium-Term, Heavy Section Base Metal Failure

Plant Source Type Description Case


Study
American Electric EPRI report Reducers in turbine After 9 years of operation, Type IV —
Power, Sporn and 1004516 piping failures in reducer-to-pipe welds
Tanners Creek due to thermal growth, hanger
adjustments
RWE, Aberthaw S. Brett, various Header stub tube Failure at 58,000 hours, Type IV 13
plant papers and cracking and
presentations branch welds
Combined-cycle Case studies JP, Low hardness in Low hardness due to initial heat —
plant (unidentified) Black and Veatch, main steam and hot treatment, postweld heat treatment,
May 2005 reheat bends, or both
spools
Progress Energy Case studies JP, SI Main steam tee Low hardness and poor —
report microstructure, no cracks
Progress Energy Progress Energy Soft induction Possibly the method of portable —
presentation, bends hardness testing
September 2008
Reliant Energy Reliant Energy Low hardness Low hardness ferrite —
Seward Plant report 0250R-07- regions in hot microstructure
085 reheat elbow
TECO Bayside SI Report E-07- Main steam and No damage —
Plant 1 381R hot reheat piping

14166640 4-1
Table 4-2 (continued)
Industry Experience with Medium-Term, Heavy Section Base Metal Failure

Plant Source Type Description Case


Study
Salt River Project, SRP presentation, Main steam piping Soft elbows, incorrect material 12
Kyrene U7 Clearwater meeting heat treatment
FPL combined- Rapkin Weld on lateral Type IV failure at 35,000 hours 14
cycle plant presentation, fittings
January 2009
Japanese 700-MW ASME committee Seam-welded Type IV cracking at 65,000 hours 15
combined-cycle presentation, June elbows
plant 2006

4.2 Case Studies—Fabrication-Related Failures

4.2.1 Case Study 12


This example describes the detection of low hardness, which was produced during the initial or
PWHT, in plant components. The case study also provides some insight on the potential for
reduced creep life related to the lower hardness [13, 14].

4.2.1.1 Construction Details and History


An initial inspection of the high-energy piping at the Salt River Project’s Kyrene Unit 7 took
place in May 2006, when Unit 7 had about 18,000 operating hours and 800 starts. At that time,
four small (6-in. [152.4-mm] OD × 0.984-in. [24.994-mm] thickness) elbows at the high-
pressure steam outlet were found to be soft (140–150 BHN as opposed to the desired minimum
190 BHN) (Figure 4-1).

14166640 4-2
Figure 4-1
Location of Soft Elbows

Replication of the microstructure also detected a ferrite–carbide structure rather than martensite
(Figure 4-2). During the next inspection (November 2007), the elbows were replaced and
subjected to metallurgical evaluation.

Figure 4-2
Microstructure (from Field Replica)

4.2.1.2 Damage Assessment


The elbows contained no cracks or other damage; they were removed only because of the low
hardness readings and the potential for early creep damage. The evaluation—consisting of
hardness testing, metallographic examination, chemical analysis, tensile testing, creep
indentation, and short-term rupture testing—was aimed at assessing this potential. Hardness was
found to be quite uniform around the cross-section of the elbow, averaging 80 Rockwell B

14166640 4-3
(150 BHN). This result provided confirmation of the field hardness techniques. Metallography
also confirmed the microstructure as a ferrite–carbide mixture instead of the desired martensite
(Figure 4-3).

Figure 4-3
Microstructure

The analysis of chemical composition confirmed that the elbows met the requirements of the
specification (Table 4-3). However, the aluminum content was at the maximum limit of the
specification, resulting in a nitrogen–aluminum ratio of about 1.725. Currently, it is thought that
a nitrogen–aluminum ratio of greater than 2.5 is necessary to control the amount of aluminum
nitride formed in the microstructure and the undesirable effect on creep properties (see Section
4.3.1, Case Study 13, for details).

Table 4-3
Results of Chemical Analysis of Elbow (wt %)

Element C Cr Mo V Nb N Al Ni
Elbow 0.11 8.37 0.96 0.21 0.07 0.069 0.04 0.01
0.08– 8.00– 0.85– 0.18– 0.06– 0.030– 0.04 0.40
A234-91
0.12 9.50 1.05 0.25 0.10 0.070 maximum maximum

The ultimate tensile strength measured 78.5–80.0 ksi (541–552 MPa) and the yield strength
measured 36.9–38.6 ksi (254–262 MPa). These values are less than the ASTM minimums of
85 ksi (586 MPa) ultimate tensile strength and 60 ksi (414 MPa) yield strength.
Material was removed from one elbow, and specimens were machined for uniaxial creep rupture
testing to determine whether the material was also weak in creep strength compared to normal
Grade 91. Assuming average Grade 91 properties, test conditions were designed to produce
failure in 300 hours and in 1000 hours. However the 300-hour test of the weak material failed in
only 18.7 hours, a 94% reduction in creep rupture life. Similarly, the 1000-hour test failed in
55.3 hours, also a 94% reduction in life. These results are consistent with previous work that
shows that improper heat treatment, which results in lower-than-desired hardness, also reduces
the creep rupture life (Figure 4-4).

14166640 4-4
Figure 4-4
Effect of Hardness on Creep Rupture Life

4.2.1.3 Conclusions
The following conclusions were reached:
• The low field hardness in the elbows was confirmed by the metallurgical evaluation.
• The low hardness values were reflected in a tensile strength that was less than specified
minimums and, most important, a significant reduction in creep rupture life.
• Improper heat treatment can result in incorrect microstructures and degraded material
properties.
• An inspection program incorporating hardness testing and microstructural replication is
necessary to identify such soft material.

14166640 4-5
4.3 Case Studies—Material Chemistry–Related Failures

4.3.1 Case Study 13


This case study describes an experience with cracking in replacement headers and the
relationship between crack susceptibility and material chemistry, particularly the nitrogen and
aluminum contents [18–22].

4.3.1.1 Background
Some early failures that occurred in Grade 91 components in RWE power plants in the United
Kingdom prompted an analysis and testing program that established a relationship between high-
temperature crack susceptibility and a low nitrogen–aluminum ratio in the material chemistry
[18].
It was shown that if the aluminum content in the steel is too high relative to the nitrogen content
(a low nitrogen–aluminum ratio), coarse aluminum nitride precipitates are produced at the
expense of the fine vanadium nitride precipitates, which are partially responsible for the creep
resistance of Grade 91 steels [19]. The result is that low nitrogen–aluminum ratio material has
both poor parent and Type IV creep strength. Figure 4-5 compares the creep life of a low
nitrogen–aluminum ratio material to the Grade 91 mean data. The parent material properties (for
the longer-term tests) are less than the Grade 91 mean - 20% line. The Type IV region rupture
lives depart even further from the mean - 20% line.

Figure 4-5
Creep Rupture Data for a Low Nitrogen–Aluminum Ratio P91 Material (Bar 257)

14166640 4-6
The bar 257 material used to generate these data were from a heat of steel that had been used to
fabricate several in-service components. Two retrofitted headers in the RWE npower Aberthaw
power station units U8 and U9 were identified as having been fabricated with this suspect cast of
P91 material. Figure 4-6 compares the heat analysis chemistry and hardness of the materials used
in the headers with data from the earlier failures [18]. It was concluded that the two headers
might be equally susceptible to cracking because both the hardness and the nitrogen–aluminum
ratios are comparable.

Figure 4-6
Nitrogen–Aluminum Ratio Versus Hardness in Suspect Casts of P91

4.3.1.2 Construction Details and History


The headers in question are shown schematically in Figure 4-7 [20]. These headers were
designed to BS1113:89 with a design pressure of 2550 psig (17.58 MPa) and a design
temperature of 1075° (580°C). Each header was constructed from six cylindrical barrel sections
(17.716 in [450 mm] OD × 1.97 in. [50 mm] thickness) made to ASTM A335 P91 in U8 and
ASTM A182 F91 in U9.

14166640 4-7
Figure 4-7
Schematic Arrangement of the Headers

The barrel sections were separated by four ASTM A182 F91 forged T-pieces and a central
circumferential butt weld. The ends of the header were closed by forged, domed ends, and the
header was fitted with four ASTM A182 F91 safety valve branches (7.48 in. [190 mm] OD ×
2.24 in. [57 mm] thickness), one ASTM A182 F91 main steam atmospheric pass-out branch
(8.27 in. [210 mm] OD × 2.13 in. [54 mm] thickness), and two smaller pressure tapping
branches. A total of 408 ASTM A213 T91 stubs (2.13 in. [54 mm] OD × 0.315 in. [8 mm]
thickness) were distributed along the header body, grouped mainly in 68 elements of 6 stubs
(A–F) each. Most of the stubs were attached to the barrel sections, but a smaller number were
attached on the forged T-pieces. On the barrels, the six stubs in each element were arranged at
50° intervals around the circumference, between 55° and 305° from top dead center position.
(Figure 4-8) [20]. A number of attachment welds were also present—main hanger supports and
antirotation lugs.

14166640 4-8
Figure 4-8
General View of the Header, Showing the Stub Arrangement

The suspect headers were installed in Aberthaw power station in 1991 (U9) and 1992 (U8). Due
to the concerns about the material, an inspection program was initiated in 2003. Extensive
cracking was discovered in the U8 header in 2004, after 58,000 hours of operation. The Type IV
cracking was mostly in the stub tube and attachment welds on the header side of the weld. In the
equally susceptible U9 header, cracking was not seen until 2007, after 78,000 hours, and then
only in the four large branch welds.

14166640 4-9
4.3.1.3 Crack Characteristics
In the U8 header, cracking was found on all large branches, on significant numbers of small
branches (stubs), and on the majority of large attachment welds. The Type IV cracking was
always found on the header side of the weld (Figure 4-9). Virtually all stub cracking was found
on the barrel-side flank positions, centered on the 0° and 180° positions, implying that the
dominant stress was the hoop stress on the header.

Figure 4-9
Crack Characteristics in Stub Tube Welds

No cracking or evidence of creep damage was found on any of the circumferential butt welds—
the barrel to T-piece, the domed end welds in the body of the header, or the T-piece outlet welds.
There is therefore no evidence of significant axial loading, either from inadequate support of the
header body or from system stresses arising from the steam pipe work leading out from the
header.
In U8, the stub tube cracking was more prevalent in 3 of the 6 barrel sections [20]. The incidence
of cracking appeared to be related to the nitrogen–aluminum ratio (Figure 4-10). Barrels 5 and 6
had a nitrogen–aluminum ratio of less than 1.5, and barrel 4 had a nitrogen–aluminum ratio of
2.8. The rest of the barrels had much lower aluminum content (and a higher nitrogen–aluminum
ratio). The contention that the nitrogen–aluminum ratio is an important factor in Type IV
cracking is thus supported by these data from the U8 header.

14166640 4-10
Figure 4-10
Distribution of Cracking Across the Six Barrel Sections in U8

The identical U9 header had the same or possibly greater susceptibility (nitrogen–aluminum ratio
less than 2.0) than the U8 header (see Figure 4-6). If chemistry were the only factor in the header
cracking, U9 would have been predicted to have the same severity of cracking as U8.
In U8, it was observed that the stub tube cracking was concentrated at the positions of highest
temperature along the header, indicating that temperature was an obviously important factor in
the cracking (Figure 4-11). In U9, the temperature distribution along the header was more
uniform, and the average temperature was slightly lower (Figure 4-12). Plant thermocouple data
also showed that the average steam temperature was slightly higher in U8 than in U9 (1050°F
[563°C] compared to 1060°F [570°C]). Thus, the delay in cracking in U9 compared to U8 can be
attributed to a slightly lower average outlet steam temperature in U9 compared to U8. The
importance of the nitrogen–aluminum chemistry factor is not negated by the performance
differences between the two headers.

14166640 4-11
Figure 4-11
Distributions of Cracked Stubs and Temperatures Along the U8 Header

Figure 4-12
Temperature Profiles in the Two Headers (Determined from Oxide Dating)

14166640 4-12
4.3.1.4 Conclusions
The following conclusions were reached:
• The prediction that material with a low nitrogen–aluminum ratio is susceptible to early Type
IV damage was substantiated by the results from the two suspect headers in Aberthaw U8
and U9, wherein extensive Type IV cracking was discovered at half the design life.
• The crack characteristics, along with related analysis, showed that the cracking was driven by
the pressure hoop stress and occurred at the highest temperature locations in the headers.
Differences in severity of cracking between U8 and U9 can be attributed to slight
temperature differences in the two headers.
• The cracking severity within the individual sections of the header correlated with differences
in the nitrogen–aluminum ratios.
• Grade 91 components of the same kinds of chemistry (particularly low nitrogen–aluminum
ratios) and operating conditions are vulnerable to early Type IV cracking.

4.4 Case Studies—Design-Related Failures

4.4.1 Case Study 14


This case study provides an example of ways in which a fitting design can produce stress
concentrations that can lead to premature failure. The fittings in question are termed integrally
reinforced, welded on lateral fittings [23].

4.4.1.1 Construction Details and History


Lateral fittings are used to connect two systems into a common header or to split flow in a header
into two separate flows. In the main steam and hot reheat systems of a power plant, they can be
used where two boilers or heat recovery steam generators are combined into a single line going
to the steam turbine. Figure 4-13 shows a typical fitting and the associated design drawing. In
contrast to 90° Weldolet fittings, lateral fittings connect to the pipe or header at less than 90°,
typically 45° or 60°. Design rules usually result in a thickness mismatch, as shown in Figure
4-13, and the consequent production of large, full-penetration welds.

14166640 4-13
Figure 4-13
Typical Design of Lateral Fitting

This example is from a 1470-MW combined-cycle plant designed for 2645 psig (18.24 MPa) and
1065°F (574°C) and operating at 2200 psig (15.17 MPa) and 1055°F (569°C). After 35,000
operating hours, a leak developed in one fitting in a main steam line (Figure 4-14). Subsequent
inspections found cracks in a total of five lateral fittings.

14166640 4-14
Figure 4-14
Leak Location in the Main Steam Line

4.4.1.2 Crack Characteristics


Cracks were found in two locations in the fitting. These are shown in an 18- × 12-in.
(457.2- × 304.8-mm) lateral fitting in Figure 4-15.

Figure 4-15
Crack Locations in Lateral Fitting

14166640 4-15
The axial cracks were found on both sides of the fittings. Details are shown in Figures 4-16 and
4-17. The north side crack had penetrated to almost a 2-in. (50.8-mm) depth.

Figure 4-16
Features of Transverse and Axial Cracks

14166640 4-16
Figure 4-17
Features of Axial Cracks

Metallographic sectioning showed the crack to be a typical Type IV crack in the piping side
HAZ (Figure 4-18). Hardness testing showed normal values for the weld metal (273–280 BHN),
the HAZ (217–267 BHN), and the base metal (211–218 BHN).

Figure 4-18
Metallographic Section of Crack

14166640 4-17
4.4.1.3 Conclusions
The following conclusions were reached:
• The axial cracks along the sides of the fittings were OD-initiated Type IV creep cracks
• The cracks in the crotch position had propagated by creep mechanisms but were initiated by
subsurface weld-related flaws (ID mismatch, slag, and start and stop of welds).
• Finite element analysis showed that high stresses are generated at both of these locations and
that they are of sufficient magnitude to produce premature creep failures.
• Inspection of these areas should be performed by liquid penetrant, replica, and ultrasonic
techniques to detect cracks and early creep damage.
• Replacement of damaged welded fittings with one-piece forged fittings is recommended.
4.4.2 Case Study 15
This study provides an example of the way in which a design that incorporates a seam weld into
a component, namely a 90° elbow in a reheat line, can produce failure [24].

4.4.2.1 Construction Details and History


The failed component was a 23-in. (587-mm) diameter hot reheat elbow in a combined-cycle
plant. The elbow had been fabricated from SA-387 Grade 91 plate, hot-formed into an elbow,
with the edges of the plate joined by a single seam weld at the intrados of the elbow. The plant
had been in service for 8 years (65,000 hours) at a design temperature of 1105°F (596°C) and
725 psig (5 MPa) pressure when the steam leakage occurred. The elbow was apparently located
close to the reheat header (Figure 4-19).

14166640 4-18
Figure 4-19
Schematic Showing Location of Leaking Elbow

4.4.2.2 Crack Characteristics


The cracks were located in the intrados seam weld, approximately in the middle or apex of the
bend angle (Figure 4-20). Visually, the grouping of parallel cracks appeared to be located in the
middle of the weld metal of the seam weld.

14166640 4-19
Figure 4-20
Location of Leaking Cracks

Metallographic sectioning, however, showed that the cracking had originated in the type IV
region in the submerged metal arc welding root passes and had progressed into the weld metal of
the submerged arc welding passes that completed the seam weld (Figure 4-21). There was some
evidence that the submerged arc welds had been repaired, which had produced microcracking in
the weld metal (Figure 4-22).

14166640 4-20
Figure 4-21
Sketch of Crack Path from Sectioning

Figure 4-22
Metallographic Section at Cusp Region

4.4.2.3 Conclusions
The following conclusions were reached:
• The reheat elbow failed after 65,000 hours due to Type IV cracking that initiated in the cusp
region of a seam weld at the intrados of the elbow.
• The crack path to the outside surface might have been influenced by microcracking in the
submerged arc welding portion of the weld metal.
• Stress concentrations due to the seam weld location at the intrados of the elbow and the
relatively large groove angle of the weld profile were factors in the failure.

14166640 4-21
4.5 Solutions
For a new plant, careful attention must be given to the heat treatment steps to ensure that low-
hardness or low-strength components are not put into service. Material test certificates should be
scrutinized so that aluminum and nitrogen contents meet or exceed the code requirements.
Aluminum content requirements have been lowered from a 0.040% maximum in the original
ASTM standard to a 0.02% maximum in the current standard. In the current VGB PowerTech
standard, the nitrogen content range has been tightened from 0.030–0.070% to 0.040–0.060% in
response to the concerns raised by the field experience [25]. The analysis methods for these
elements should also be identified on material certificates.
For existing plants, inspection programs should incorporate hardness testing to identify hard or
soft regions in components. Programs should include material sampling for chemical analysis to
identify components most at risk due to low nitrogen–aluminum ratios. Inspection programs
should also focus on certain types of lateral fittings. Appropriate inspection programs (including
focused beam ultrasonic testing) should be developed if any seam-welded components are
present in the system.

14166640 4-22
5
INDUSTRY EXPERIENCE—WELD-RELATED DAMAGE
OR FAILURE
5.1 Reported Failures
Some Type IV creep failures associated with weld HAZs are described in Sections 2–4. Several
reported failures are directly attributable to the welds themselves or to weld-related heat
treatment (Table 5-1). These weld failures can be related to either fabrication or design.

Table 5-1
Industry Experience with Weld-Related Failures

Case
Plant Source Type Description
Study
Weld failure at safe end transition weld to
Midwest U.S. EPRI report Stainless steel P22, weld cracking due to delay in postweld

plant, 2001 1004516 outlet header heat treatment, fatigue failure, thermal
stresses and bending
Dissimilar metal weld after 6 years, located
Superheater
EPRI report at toe of weld on P91 side, stress corrosion
Alstom, 2001 header to P22 —
1004516 cracking due to storage before postweld
safe ends
heat treatment
Main steam
Failure at 1000 hours in P91 heat-affected
EPRI report (P91) transition
Italian plant zone, carbide migration from P91 to the 625 —
1004516 piece (Inconel
filler, leaving carbide-depleted zone
625)
Dissimilar metal weld between P91 fitting
EPRI report Main steam, wye and P22 transition piece, carbon-depleted
Plant A, 1998 —
1004516 block zones during postweld heat treatmentdue to
chromium differences
Reliant Energy Main steam
Stress corrosion cracking failure due to high
Reliant Energy report 0250R- small-bore pipe- 16
hardness, weld not stress relieved
07-092 flow transmitter
Main steam pipe
J. F. Henry,
to stop valve in Failure at <5000 hours at dissimilar metal
J. D. Fishburn
Alstom heat recovery weld between CrMoV valve and P91 17
presentation,
steam generator piping, no cold spring during erection
June 2006
plant
Iberdrola Main steam
Presentation, Dissimilar metal weld failure at 4000 hours,
combined- piping to high- 18
Aug 2008 inadequate weld profile, high stresses
cycle plant pressure casing

14166640 5-1
5.2 Case Studies—Fabrication-Related Failures

5.2.1 Case Study 16


This case study provides an example of how a delay or omission of PWHT can result in damage
due to stress corrosion cracking [26].

5.2.1.1 Construction Details and History


This example was from Reliant Energy’s Choctaw Unit 101, a combined-cycle gas-fired plant
that was commissioned in 2003. The failure occurred in 2007 (operating hours were not
reported) in a small-bore main steam transmitter pipe at a socket weld (Figure 5-1). The piping
was 1 in. (25.4 mm) diameter × 0.25 in. (6.35 mm) wall thickness, with material specified as
P91. Operating conditions were 2010 psi (13.86 MPa) and 650°F (343°C).

Figure 5-1
Failure Site in Flow Transmitter Pipe

14166640 5-2
5.2.1.2 Crack Characteristics
The main fracture was on the socket side of the weld and appeared to have propagated through
the entire weld. A secondary crack was present on the pipe side of the weld (Figure 5-2).

Figure 5-2
Fracture Site—End View of Main Fracture and Side View of Secondary Crack

14166640 5-3
A metallographic section through the weld (Figure 5-3) showed that both fractures had initiated
at the outside surface and propagated toward the bore.

Figure 5-3
Section Through Weld, Showing Main Fracture (Top) and Secondary Crack (Bottom)

14166640 5-4
The crack path at the outer surface was intergranular (Figure 5-4), but it changed to transgranular
closer to the inside wall.

Figure 5-4
Portion of Main Crack at Outside Diameter Surface

Scanning electron microscope examination confirmed the intergranular nature of the crack
initiation portion (Figure 5-5).

Figure 5-5
Fracture Surface, Showing Intergranular Features

14166640 5-5
Metallographic examination of the weld metal showed it to be martensitic, with a high hardness
of 45–47 Rockwell C (Figure 5-6). A submitted sample of the pipe insulation was analyzed as
having a pH of 9 and a high sodium content.

Figure 5-6
Microstructure in the Weld Metal (45–47 Rockwell C)

5.2.1.3 Conclusions
The following conclusions were reached:
• The small-bore piping had failed due to a stress corrosion cracking mechanism.
• Sodium leached from wet insulation was suspected as the cracking medium.
• The weld had not been postweld stress-relieved and was in a fully martensitic condition,
rendering it susceptible to stress corrosion cracking.

5.3 Case Studies—Design-Related Failures

5.3.1 Case Study 17


P91 can be prone to type IV cracking in the HAZ zone of weldments relatively early in life,
particularly when the strain localization is a consequence of piping stresses and local geometry.
This case study illustrates the way in which strain localization can also produce weld metal
cracking that can appear similar to Type IV cracking. In this case, the failure was attributed to
additional stresses arising from the lack of a cold pull [27].

5.3.1.1 Construction Details and History


This failure occurred in a main steam line, feeding a turbine from a heat recovery steam
generator. The main steam piping was 18 in. (457 mm) OD, Schedule 140, 1.562 in. (40 mm)
wall thickness, manufactured in SA-335 Grade 91 material. At the leak location, the P91 pipe

14166640 5-6
was joined to a stop–control valve made from 1.25Cr-1Mo-0.25V material with a dissimilar
metal weld with filler metal of 2.25Cr-1Mo. The design outlet steam temperature was 1050°F
(565°C), and the design steam pressure was 1800 psi (12.4 MPa). The plant had less than 5000
hours of service when, during an in-service walkdown survey of the system, the leak was
observed and a slightly enlarged and soft section of insulation was discovered. This insulation
was carefully probed, and the presence of steam became apparent (Figure 5-7).

Figure 5-7
Failure Location at Turbine Valve

5.3.1.2 Crack Characteristics


Examination of the leak site revealed a relatively tight circumferential crack that appeared to be
located between the Grade 91 pipe and the weld (Figure 5-8). From the top of the weld (the
12 o’clock position as viewed from the upstream side of the weld), the circumferential crack
extended in the clockwise direction from the 2:30 position to the 7 o’clock position—
approximately 135° around the circumference of the weld, or a distance of 24.5 in. (620 mm).
Visual examination of the ID surface confirmed that the crack observed on the OD surface of the
pipe did, in fact, extend completely through the pipe wall and was approximately 7 in. (180 mm)
long on the ID surface. The crack appeared to have initiated at or near the OD surface.

14166640 5-7
Figure 5-8
Crack on Outside of Pipe

A metallographic section close to the center of the crack (Figure 5-9) showed that the crack
closely followed the weld fusion line on the piping (P91) side of the weld.

Figure 5-9
Section Through Weld, Close to Center of Crack Indication

14166640 5-8
Detailed examination showed that the cracking had propagated through a narrow zone of
material in the 2.25Cr-1Mo weld deposit, immediately adjacent to the weld fusion boundary that
had been partially decarburized (Figure 5-10).

Figure 5-10
Microstructural Details of Crack Path

14166640 5-9
Within this carbon-depleted zone, the intergranular cracking was associated with the formation
of cavities along individual grain boundaries, which eventually linked together to form a
continuous crack (Figure 5-11).

Figure 5-11
Evidence of Creep Cavitation and Intergranular Cracking

14166640 5-10
These features are characteristic of a creep mechanism. Hardness testing of the weld and parent
metals showed that the weld had received proper PWHT, but likely at the low end of the required
range (Table 5-2). Microhardness testing of the carbon-depleted zone showed it to be
surprisingly high in hardness, indicating that it must been subjected to a significant strain
hardening from in-service stresses (Table 5-3). This was thought to be due to a local strain
concentration effect, resulting from the geometry of the joint and the mismatch in hot strength of
the P91 piping metal and P22 weld metal.

Table 5-2
Results of Hardness Testing of the Parent, Heat-Affected Zone, and Weld Metals

Vickers Hardness Values (HV)


Quarter Point P91 HAZ Weld Metal
12 o’clock 262, 254, 245 306, 317, 306 238, 245, 225, 237, 227
3 o’clock 247, 240, 274 274, 306, 311 258, 240, 238, 237, 238
6 o’clock 262, 233, 237 304, 317, 285 297, 302, 247, 260, 238
9 o’clock 264, 230, 251 319, 322, 309 311, 281, 266, 235, 232

Table 5-3
Results of Microhardness Testing of the Heat-Affected Zone, Carbon-Depleted Zone, and Weld
Metals

Vickers Hardness Values (HV)


Quarter Point HAZ Carbon-Depleted Zone Weld Metal
12 o’clock 296, 301, 301, 307 324, 324, 336 216, 219, 230
6 o’clock 290, 307, 301, 312, 31 336, 356 223, 237, 230

5.3.1.3 Stress Analysis


Detailed stress analysis was carried out to attempt to explain this early failure. The results
showed that the primary stresses and the secondary stresses due to thermal expansion of the
steam line piping were within the limits prescribed by ASME B31.1, Power Piping; therefore,
they should not have been a factor in the premature failure of the joint.
However, the analysis showed that the effect of axial stresses of even moderate magnitude
(<10 ksi [<70 MPa]) acting on the dissimilar metal joint, particularly the decarburized zone,
could be significant in the life of the joint and that the axial stresses associated with thermal
expansion might also be a factor. It was determined that the piping system had not been erected
with a cold pull, and this would have increased the axial stresses.
The different stress relaxation responses of the P22 and P91 materials might also have been a
factor. The P22 material, being relatively weak in creep, can undergo a significant amount of
primary creep at typical design stresses to dissipate stress concentrations in areas of high stress
concentration, such as in this transition joint. The P91 material, having higher creep strength, is
not as accommodating. The weld joint was redesigned to reduce the stress concentration effects
(Figures 5-12 and 5-13).

14166640 5-11
Figure 5-12
Original Design of Transition Weld, Illustrating the Geometrical Stress Concentrating Effect

Figure 5-13
Redesigned Joint, Incorporating a P91 Transition Piece to Reduce the Stresses

14166640 5-12
5.3.1.4 Conclusions
The following conclusions were reached:
• The piping support system and geometry were designed and implemented in a way that was
fully consistent with the rules of the design code—primary and secondary stresses were
below the required limits.
• Thermal transients played no significant role in the failure.
• Lack of a cold pull contributed to raising the axial stress level in the hot condition during the
initial period of operation.
• The axial stress across a decarburized zone was a significant factor.
• The effect of a relatively low creep relaxation rate in the creep-strength-enhanced ferritic
alloys, when used in mixed alloy systems, combined with the adverse but currently
considered acceptable geometric and material transition, is the probable primary cause of the
high axial stresses and, therefore, the failure.
5.3.2 Case Study 18
This study provides another example of the way in which design-related local stress
concentrations—in this case, augmented by a fabrication error—can lead to failure [28].

5.3.2.1 Construction Details and History


This example is from the main steam system in a combined-cycle plant. Failure location was at a
dissimilar metal weld between the piping (SA-335 P91) and the inlet nozzle of the high-pressure
turbine casing (SA-356 Grade 9, 1.25Cr-1Mo-0.25V). The weld filler metal was E9018-B9. The
failure occurred after 4000 hours of operation.

5.3.2.2 Cracking Characteristics


A significant crack was detected on the nozzle side (Grade 9) of the weld metal and followed the
weld toe and extended around three-fourths of the nozzle circumference (Figure 5-14).

14166640 5-13
Figure 5-14
Crack Location in High-Pressure Inlet Nozzle

Fracture examination showed that the cracks appeared to mirror the weld bead profile
(Figure 5-15).

Figure 5-15
View of Fracture Surface on Nozzle Side, Showing Weld Bead Profile

14166640 5-14
This observation was confirmed by a metallographic section of the weld metal (Figure 5-16),
which also showed that the crack was located in the outer part of the HAZ on the nozzle side and
that there was accompanying damage in the form of cavities and microcracks in the HAZ. This
damage appeared to be in a carbon-depleted zone due to the differences in chromium content
between the nozzle and the Grade 91 weld bead.

Figure 5-16
Metallographic Section Through Crack and Weld Beads

14166640 5-15
The investigation also determined that the welding had not been done correctly and that the weld
profile was inadequate (Figure 5-17).

Crack
Localización Missing
deLocation
la grieta
Weld Beads
Cordones sin realizar

Cordones existentes
Existing
Weld Beads

Figure 5-17
Schematic of Weld Profile

5.3.2.3 Conclusions
The following conclusions were reached:
• The crack was located in the weaker (P9) nozzle material and appeared to be associated with
a carbon-depleted zone in the dissimilar metal weld.
• Inadequate weld profile contributed to the stresses at the failure location.
• Some minor discrepancies in the welding procedure were discovered, but it was not clear
how they might have contributed to the failure.

5.4 Solutions
These examples provide further evidence that welds in Grade 91 material must be properly
processed, particularly with regard to PWHT parameters. It also appears that dissimilar metal
welds with P91 might be vulnerable to strain localization damage due to the mismatch in creep
properties and any additional geometrical stress concentrations. Carbon-depleted zones can
develop in P91 dissimilar metal welds and lead to premature failures. Inspection programs
should pay careful attention to dissimilar metal welds.

14166640 5-16
6
DAMAGE ASSESSMENT TECHNIQUES
Within power generating plants, there is an increasing use of Grade 91 both in tubing and in
components of larger sections, such as headers and piping. It is apparent that a small but
significant amount of the Grade 91 material installed as critical pressure part components in
existing plants has been processed in such a way that the properties are uncertain. Thus, evidence
suggests that the elevated-temperature performance of this material likely will not meet the
expectations of the designers of those components.
Recent inspections at both standard fossil-fuel-fired and heat recovery type steam generators
have uncovered evidence of soft material with an undesirable microstructure condition in several
components. This is clear evidence that proper controls were not maintained during some phase
of the manufacturing or erection of the components. The deficient material has involved both
weldments and base metal independent of any weld. In most of the cases, the available
documentation has not been adequate to allow a determination of how the deficient condition
developed. Therefore, it has not been possible to accurately determine the extent of the
deficiencies without resorting to costly component testing and plant inspections.
A range of short- and medium-term cracking problems have also been identified. In many cases,
these have been associated with bends or welds. The short-term issues tend to be associated with
stress corrosion cracking, whereas medium-term issues are mostly in the form of localized creep
within decarburized weld zones or Type IV cracking in the HAZ. Of particular concern is that,
although creep cracking under high bending loads occurs at the outside surface of the
component, creep damage appears to be initiating midwall in other cases. In most applications,
the occurrence of cracking subsurface significantly increases the challenges associated with
detection and characterization of damage.
For plant operators, there is a need to verify initial quality and to assess damage during service,
so that component failures can be avoided. Several current programs are studying factors that
affect the performance of the creep-strength-enhanced ferritic steels. In particular, ongoing
research managed by EPRI is evaluating critical information on all aspects of the use of this class
of steels. In addition to reviews of existing information, new work is in progress. The assessment
tools developed, the performance data, and the analysis of key information will ensure that the
practical solutions established will provide the optimum overall approach to lifing Grade 91
components. The key aspects of the EPRI program are organized under the following two major
objectives:
• Objective 1: Initial material integrity—purchase instructions, materials and processing
specifications, inspection protocols, and post-sampling protocols
• Objective 2: Performance in service—elevated-temperature behavior and life assessment

14166640 6-1
6.1 Objective 1—Initial Material Integrity
This objective focuses on front-end issues—those issues that pertain to how the material is
ordered, how it is processed, how quality control is maintained during processing, and how the
material is inspected in the shop and in the field to determine its condition before or very soon
after initial operation. The intent of this objective is to ensure that deficient material is never
installed.
An initial aspect, which is vital to an effective life cycle management approach, is to develop the
necessary specifications and procedural documents that will enable users to control the quality of
the components supplied. Effective quality control is necessary at every stage of implementation,
from the manufacturer’s original purchase of the material to the manufacturing of individual
components to construction of multicomponent systems to life management after the plant is
operational.
The work performed under this objective includes an assessment of the most effective methods
for finding and characterizing deficient material in components. This entails identifying best
practices for an inspection strategy, including the following:
• Which components to inspect and at what level of detail
• Which inspection methods to use (such as hardness testing and metallographic replication
and the appropriate surface preparation for each)
• Post-inspection activity (such as sample removal, engineering analysis, weld repair, and heat
treatment).

6.2 Objective 2—Performance in Service


The broad issue of life management during service is critical. However, simply applying
previous approaches is not appropriate, given the acknowledged metallurgical complexity of
these steels and recognizing that creep-strength-enhanced ferritic steels are inherently unstable
and will degrade during normal operation at elevated temperatures. It is clear that both an
understanding of how the physical evidence of damage changes during operation and the
validation of methods to document this damage are required.
A basis for all assessment methods is a comprehensive database, linking microstructure to
properties. Therefore, the research team is seeking to characterize, as fully as possible, the
elevated-temperature mechanical behavior of not only material that has been properly processed
but also—and of equal importance—deficient material in all of the variations of deficiency that
might exist.
A critical step in this characterization has involved establishing a master record, or atlas, of
microstructures. This atlas correlates a composition and a specific condition of microstructure
with a range of hardness and creep behavior. It has been established through extraction, review,
and compilation of information that currently exists, followed by critical analysis to identify
knowledge gaps and areas for additional laboratory work. One of the values of such an atlas will
be the ability to compare information obtained during field inspections to the data contained in
the atlas as a basis for a qualitative assessment of material serviceability. In characterizing the
material behavior, separate evaluations have been performed for the base metal and weld metal

14166640 6-2
because it is well understood that weldment properties can vary significantly from those of the
base material and that, in many cases, the weldment properties govern the life of the component.
This phase of the study will also include 1) validation of new testing and assessment techniques
that have been applied to these steels, including small-specimen creep indentation testing and the
use of elevated-temperature strain gauges for direct measurement of on-load strain accumulation
and 2) review of the relative merits of various life prediction methodologies, including the more
common parametric methods, the Omega approach, and the Monkman-Grant relationship. This
review provides the basis to define specific material properties and how those properties change
with composition, fabrication, and operating variables. When the study is complete, the intent is
to publish a handbook that will provide the necessary information to successfully control all
major aspects of the use of these steels.

14166640 6-3
14166640
7
CONCLUSIONS AND RECOMMENDATIONS
The family of martensitic creep-strength-enhanced ferritic steels was developed to address a
recognized need for high-temperature alloys with enhanced properties, including high creep
rupture strength along with satisfactory fracture toughness. Typically, steels from this family
enter service in the normalized and tempered condition. Proper heat treatment results in M23C6
carbides at prior austenite and martensite lath boundaries in addition to columbium (niobium)
and vanadium intergranular carbonitrides. This combination of precipitates and martensitic
structure provides a marked increase in the high-temperature strength and creep resistance.
However, some components require multiple heat treatments because of further fabrication
operations such as bending or welding. It is now well established that the level of control
associated with producing the expected tempered martensitic structure is quite high. Moreover,
the consequences associated with inadequate control can be quite severe; for example, exceeding
a defined maximum temperature could result in a large reduction in strength. Even more
challenging is the fact that the allowable variation of the alloy composition within the
specification introduces further difficulties associated with a lack of detailed control of heat
treatment. This means that the effects introduced by the same level of temperature excursion can
vary depending on the specifics of the actual component composition.
Properly heat-treated, the modified 9Cr alloys such as Grades 91 and 92 provide improved high-
temperature strength compared to traditional low-alloy steels such as P11 or P22. This higher
strength makes creep-strength-enhanced ferritic steels attractive for utility applications because
the higher operating temperatures and lower wall thickness improve efficiency and the capability
for faster cycling. In particular, Grade 91 steel has been widely used since the 1990s as a retrofit
material in conventional subcritical fossil plants and in high-temperature locations (up to or even
greater than 1100°F [593°C]) in new combined-cycle and supercritical plants.
This report summarizes recent experiences and problems that utilities and suppliers have
encountered with the alloy. Detailed case studies of representative failures and problem areas are
used to illustrate the general trends that are emerging. The issues fall into the following three
categories:
• Improper or poor control of initial heat treatment variables that result in inferior properties or
damage or both
• Design or operational issues that produce increased stresses
• Longer-term performance issues of the alloy
Analysis of these longer-term failures has made clear that Grade 91 steel is inherently susceptible
to the Type IV failure mode associated with weldment HAZs. Suggestions are made for possible
solutions to all these issues in terms of control of heat treatment and fabrication variables,
improved designs, and monitoring programs. Techniques for longer-term life assessment of
Grade 91 components are being studied and will be addressed in later reports.

14166640 7-1
14166640
8
REFERENCES
1. R. Viswanathan and W. Baker, “Materials for Ultrasupercritical Coal Power Plants—
Boiler Materials: Part 1, Journal of Materials Engineering and Performance, Vol. 10,
No. 1, pp. 81–93 (February 2001).
2. J. R. DiStefano, “Summary of Modified 9Cr-1Mo Steel Development Program, 1975–
1985,” Oak Ridge National Laboratory: Oak Ridge, TN: October, 1986. ORNL-6303.
3. Performance Review of P/T91 Steels. EPRI Palo Alto, CA, and European Technology
Development Ltd.: 2002. 1004516.
4. J. Arnold, “Stress Corrosion Cracking of Creep-Strength Enhanced Ferritic Steels,”
presented at the EPRI Conference (November 2003).
5. KEPRI, “Yung Heong #1 FSH Tube Failure,” presentation (January 2008).
6. P. W. Woo, presentation (October 17, 2007).
7. Remaining Life Assessment of SA213 T91 Steel Superheater and Reheater Tubes
Subjected to Long term Overheat Creep Damage—Glen Lyn Unit 6 Secondary
Superheater Case Study. EPRI: Palo Alto, CA: 2008. Unpublished draft report.
8. S. Scully and P. Bernard, “The Life Management of Reheater Tube Panels in a Coal-
Fired Boiler with ‘Upgraded’ T91 Tube Material,” presented at the ETD Seminar on
Modified 9% Chrome Steels (June 20–21, 2007).
9. Dominion Fairless, “P91 Pipeline Failure,” presentation (2006).
10. Metallurgical Evaluation of Pipe Segment Subject to Cracking As a Result of Localized
Overheating, Hot Reheat Piping System Block No. 1, Dominion Fairless Energy, LLC,
Fairless Hills, Pennsylvania, January 20, 2004. Thielsch Engineering: Cranston, RI:
2004. Report 10923.
11. M. Sims, “P91 Pipe Failure Experience and Inspection/monitoring program,” presented at
the Workshop on P91/P92, Clearwater Beach, FL (April 7–9, 2008).
12. Progress Energy, “Grade 91 Experience in PGN–EPMC,” presentation (June 2006).
13. J. A. Alice and J. Henry, “SRP’s Experience with P91 Piping at Kyrene Generating
Station Unit 7,” presented at the Workshop on P91/P92, Clearwater Beach, FL
(April 7–9, 2008).
14. Inspection of Selected High-Energy Piping Welds at Kyrene Unit 7. Salt River Project,
Tempe, AZ: May 30, 2006. Report MEV-3880.
15. D. J. Allen and S. J. Brett, “Premature Failure of a P91 Header Endcap Weld: Minimising
the Risks of Additional Failures,” Proceedings of an International Symposium on Case
Histories on Integrity and Failures in Industry, Milan, Italy, pp. 133–143 (September 28–
Oct 1, 1999).
16. S. J. Brett, D. J. Allen and J. Pacey, “Failure of a Modified 9Cr Header Endplate,”
Proceedings of an International Symposium on Case Histories on Integrity and Failures
in Industry, Milan, Italy, pp. 873–884 (September 28–Oct 1, 1999).

14166640 8-1
17. Inspection of Selected High Energy Piping Welds and Fittings at Kyrene Unit 7. Salt
River Project: Tempe, AZ. November, 15, 2007. Report MEV-3931.
18. S. J. Brett, D. J. Allen, and L. W. Buchanan, “The Type IV Creep Strength of Grade 91
Materials,” presented at the Third International Conference on the Integrity of High
Temperature Welds, IOM3, London (April 24–26, 2007).
19. S. J. Brett, J. S. Bates, and R. C. Thomson, “Aluminium Nitride Precipitation in Low
Strength Grade 91 Power Plant Steels,” presented at the EPRI Fourth International
Conference on Advances in Materials Technology for Fossil Power Plants, South
Carolina (October 25–28, 2004).
20. S. J. Brett, “Early Type IV Cracking on Retrofit Grade 91 Steel Headers,” presented at
the IIW International Conference on Safety and Reliability of Welded Components in
Energy and Processing Industry (2008).
21. S. J. Brett, D. L. Oates, and C. Johnston, “In-Service Type IV Cracking on a Modified
9Cr (Grade 91) Header,” presented at the ECCC Conference: Creep & Fracture in High
Temperature Components–Design & Life Assessment Issues, IMechE, London
(September 12–14, 2005).
22. S. J. Brett, “Case Study—Retrofit Modified 9Cr (Grade 91) Steel Headers,” presented at
the Workshop on P91/P92, Clearwater Beach, FL (April 7–9, 2008).
23. K. Rapkin, “FPL Experience, Fabricated Grade 91 Laterals,” presentation (January 28,
2009).
24. “Property of Modified 9Cr 1Mo Weld Joint” IHI Presentation (June 2006).
25. Material Specification for Components under Pressure in Fossil-Fired Power Plants,
X10CrMoVNb9-1 (P/T91). VGB PowerTech e.V., Klinkestraße, Essen: 2007.
VGB-R109.
26. Failure Analysis of MS Flow Transmitter Line Choctaw Unit 101. Reliant Energy:
Houston, TX: August 2007. Report 0250R-07-092.
27. J. F. Henry and J. D. Fishburn, “Investigation of a Leak in a Main Steamline Piping Joint:
Causes and Implications,” presentation (June 2006).
28. Iberdrola Power Generation Technology Dept., “Failure in the Main Steam Pipe Welding
in a CC Power Plant,” presentation (August 2008).

14166640 8-2
14166640
Export Control Restrictions The Electric Power Research Institute (EPRI)
Access to and use of EPRI Intellectual Property is The Electric Power Research Institute (EPRI), with
granted with the specific understanding and major locations in Palo Alto, California; Charlotte,
requirement that responsibility for ensuring full North Carolina; and Knoxville, Tennessee, was
compliance with all applicable U.S. and foreign export established in 1973 as an independent, nonprofit
laws and regulations is being undertaken by you and center for public interest energy and environmental
your company. This includes an obligation to ensure research. EPRI brings together members, participants,
that any individual receiving access hereunder who is the Institute’s scientists and engineers, and other
not a U.S. citizen or permanent U.S. resident is leading experts to work collaboratively on solutions to
permitted access under applicable U.S. and foreign the challenges of electric power. These solutions span
export laws and regulations. In the event you are nearly every area of electricity generation, delivery,
uncertain whether you or your company may lawfully and use, including health, safety, and environment.
obtain access to this EPRI Intellectual Property, you EPRI’s members represent over 90% of the electricity
acknowledge that it is your obligation to consult with generated in the United States. International
your company’s legal counsel to determine whether participation represents nearly 15% of EPRI’s total
this access is lawful. Although EPRI may make research, development, and demonstration program.
available on a case-by-case basis an informal
assessment of the applicable U.S. export classification Together . . . Shaping the Future of Electricity
for specific EPRI Intellectual Property, you and your
company acknowledge that this assessment is solely
for informational purposes and not for reliance
purposes. You and your company acknowledge that it
is still the obligation of you and your company to make
your own assessment of the applicable U.S. export
classification and ensure compliance accordingly. You
and your company understand and acknowledge your
obligations to make a 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.

© 2009 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.

1018151

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 • PO Box 10412, Palo Alto, California 94303-0813 • USA
14166640800.313.3774 • 650.855.2121 • askepri@epri.com • www.epri.com

You might also like