You are on page 1of 814

Proceedings of the 2000 International Conference on Fatigue of Reactor Components (MRP-46)

PWR Materials Reliability Program (PWRMRP)


1006070 Also referenced as OECD/NEA/CSNI/R (2000) 24

Proceedings, June 2001

Cosponsors Organisation for Economic Co-operation and Development (OECD) Nuclear Energy Agency/Committee on the Safety of Nuclear Installations (NEA/CSNI) U.S. Nuclear Regulatory Commission

EPRI Project Manager S. Rosinski J. Carey

EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA 800.313.3774 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 EPRI

ORDERING INFORMATION
Requests for copies of this report should be directed to EPRI Customer Fulfillment, 1355 Willow Way, Suite 278, Concord, CA 94520, (800) 313-3774, press 2. Electric Power Research Institute and EPRI are registered service marks of the Electric Power Research Institute, Inc. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power Research Institute, Inc. Copyright 2001 Electric Power Research Institute, Inc. All rights reserved.

CITATIONS
This proceedings was prepared by EPRI 1300 W.T. Harris Blvd. Charlotte, NC 28262 Principal Investigator S. Rosinski This proceedings contains information presented at the First International Conference on Fatigue of Reactor Components, sponsored by EPRI, OECD/NEA/CSNI, and the U.S. NRC. The report is a corporate document that should be cited in the literature in the following manner: Proceedings of the 2000 International Conference on Fatigue of Reactor Components (MRP46): PWR Materials Reliability Program (PWRMRP), EPRI, Palo Alto, CA, and Organisation for Economic Co-operation and Development (OECD/NEA/CSNI/R[2000] 24), and the U.S. NRC: 2001. 1006070.

iii

REPORT SUMMARY

This report contains information presented at the First International Conference on Fatigue of Reactor Components held July 31August 2, 2000, in Napa, California. The conference sponsored by EPRI, Organisation for Economic Co-operation and Development Nuclear Energy Agency/Committee on the Safety of Nuclear Installations (OECD NEA/CSNI), and the U.S. Nuclear Regulatory Commission (U.S. NRC)provided a forum for the technical discussion of fatigue issues that affect the integrity and operation of light water reactor components. Approximately 90 fatigue experts, representing 12 countries, participated in the conference. Strong representation was shown by nuclear operators, vendors, regulatory agencies, research and development organizations, and other experts. Background Fatigue is a primary degradation mechanism that affects nuclear power plant components worldwide. The effective management of fatigue is important to the continued safe operation of plant components during present operation and as plants consider long-term operation. The EPRI Materials Reliability Program (MRP) identified the need to bring together international experts to discuss significant fatigue issues affecting nuclear plant operations in order to share common experiences and identify outstanding technical issues. Objectives N To provide a forum for the technical discussion of fatigue issues that affect the integrity and operation of light water reactor components
N N

To share common experiences regarding fatigue of reactor components to ensure continued safe operation To identify common areas of interest in order to foster future international research and collaboration activities

Approach The conference was organized in a series of technical presentations and group discussion sessions focused on the following fatigue-related topics:
N N N N N

Thermal fatigue Environmental fatigue Nondestructive evaluation/testing Fatigue monitoring/evaluation Codes and standards v

N N

Vibration/high cycle fatigue The conference was structured to benefit utility and plant managers as well as system, materials, structural integrity, licensing, and maintenance/repair engineers.

Results Approximately 90 fatigue experts, representing 12 countries, participated in the conference. Strong representation was shown by nuclear operators, vendors, regulatory agencies, research and development organizations, and other experts. At the conclusion of the conference, major discussion points were summarized. While not intended to represent a comprehensive list of conference conclusions endorsed by all participants, these points were the subject of considerable discussion during the conference and may be used to foster future international research and collaboration activities. Based on the degree of technical exchange that occurred and the breadth of information provided during the conference, the participants recommended that another conference on this topic be held in 1824 months. EPRI Perspective Fatigue management is an important aspect of the continued safe operation of plant components. Periodic discussion of fatigue-related issues in an international forum allows the sharing of common experiences and fosters international collaboration in the resolution of fatigue issues. It is anticipated that future conferences in this series will continue to be a major forum for the discussion of plant component fatigue issues. Keywords Thermal fatigue Environmental fatigue Fatigue evaluation

vi

ACKNOWLEDGMENTS
Appreciation is extended to the conference organizers for their contributions: EPRI S. Rosinski J. Carey OECD E. Mathet U.S. NRC E. Hackett And to the following individuals for preparing this proceedings volume: C. Laundon L. Perry

vii

CONFERENCE SUMMARY
The primary objective of the conference was to provide a forum for the technical discussion of fatigue issues that affect the integrity and operation of light water reactor components. Approximately 90 fatigue experts, representing 12 countries, participated in the conference. Strong representation was made by nuclear operators, vendors, regulatory agencies, research and development organizations, and other experts. Technical discussions were focused in the following areas: Thermal Fatigue Environmental Fatigue Nondestructive Evaluation/Testing Fatigue Monitoring/Evaluation Codes and Standards Vibration/High Cycle Fatigue Following the technical presentations, a general discussion was held to summarize major points identified by various speakers during the conference. While not intended to represent a comprehensive list of conference conclusions endorsed by all participants, these points received considerable discussion during the conference and may be used to foster future international research/collaboration activities. The major discussion points identified by the participants are provided below.
N N

Collaboration and cooperation on an international scale are critical to the success of resolving fatigue issues, including sharing of data, test programs, and theories. It is recognized that conservatism exists in the ASME Code fatigue design procedures. Plant-specific analyses using actual plant operating parameters (transient occurrence and severity) may significantly reduce the conservatism in overall fatigue usage factor determination. Significant advancements have been made in the international community regarding the effects of thermal fatigue and reactor water environment. Additional research and international collaboration are recommended in these areas in order to resolve technical issues and utilize the results of these efforts in various operating plant criteria. An understanding should be developed between ASME Code analysis and laboratory testing regarding reactor water effects on fatigue life.

ix

The characterization of thermal hydraulic phenomena is complex and an important aspect in the quantification of thermal fatigue. Additional work regarding the proper characterization of thermal hydraulic phenomena is recommended. A background document regarding the development of implicit fatigue design criteria in B31.1 should be developed. The following fatigue design Code changes/improvements were discussed: Modification of low cycle fatigue analysis procedures Addition of thermal fatigue analysis procedures for Class 2 piping Addition of warnings in Class 1, 2, and 3 design codes for dead legs/stratification and mixing tees with corresponding thresholds Changes in the existing Code fatigue design S-N curves based on additional data and the further evaluation of these data Determination of updated reduction factors Differentiation between thermal and mechanical loads Consideration of surface striping/craze cracking

N N

N N

The development of expert tools is recommended to provide a larger understanding of fatigue degradation mechanisms. Weld overlay repairs were presented as an effective method for repairing leaking standard socket welds and providing sufficient fatigue resistance to operate to the next outage and beyond. In addition, welding process and geometry enhancements were reported to improve the fatigue life of socket welds. Instrumentation and monitoring can confirm the existence of high cycle thermal loads, except for high frequency fluctuations. Field experience indicates that the relatively limited number of locations that experience thermal fatigue are due primarily to the following: Stratification Dead legs and vortex conditions without a leak Reversing zones with large T

N N

N Risk-informed considerations, including reactor operating experience, should be applied to


the management of fatigue technical issues. Based on the degree of technical exchange that occurred and the breadth of information provided during the conference, the participants recommended that another conference on this topic be held in 18-24 months.

ATTENDEE LIST
Frank Ammirato EPRI 1300 W.T. Harris Blvd. Charlotte, NC 28262 704-547-6129 / 704-547-6168 (fax) fammirat@epri.com Robert Bain Stone & Webster Engineering 245 Summer St. Boston, MA 02210 617-589-2109 / 617-589-1315 (fax) robert.bain@stoneweb.com

Gabriel Barslivo Swedish Nuclear Power Inspectorate SE-106 58 Stockholm Sweden 46-869-88660 / 46-866-19086 (fax) gabriel.barslivo@ski.se

Alan Bilanin Continuum Dynamics, Inc. 34 Lexington Ave Ewing, NJ 08618-2302 609-538-0444 x108 / 609-538-0464 (fax) alan@continuum_dynamics.com

Urs R. Blumer CCI AG PO Box 8404 Winterthur Switzerland 41-52-262-5936 / 41-52-262-0039 (fax) urs.blumer@ccivalve.ch Kevin Bratcher Framatome Technologies Inc. 3315 Old Forest Rd. Lynchburg, VA 24506 804-832-2789 / 804-832-2831 (fax) kbratcher@framatech.com

Scott Bond Ameren UE PO Box 620 Fulton, MO 65251 573-676-8519 / 573-676-4334 (fax) smbond@cal.ameren.com

John Carey EPRI 3412 Hillview Ave. Palo Alto, CA 94304-1344 650-855-2105 / 650-855-7945 (fax) jcarey@epri.com

Bob Carter EPRI 1300 W.T. Harris Blvd. Charlotte, NC 28262 704-547-6019 / 704-547-6035 (fax) bcarter@epri.com

Kenneth Chang American Electric Power 500 Circle Drive Buchanan, MI 49107 616-697-5525/616-697-5570 (fax) kcchang@aep.com

xi

Omesh Chopra Argonne National Laboratory 9700 South Cass Ave., Bldg. 212 Argonne, IL 60439 630-252-5117/630-252-3604 (fax) okc@anl.gov

Roy Corieri Niagara Mohawk PO Box 63 Lycoming, NY 13093 315-349-4051 / 315-349-1581 (fax) corierir@nimo.com

Kurt Cozens Nuclear Energy Institute 1776 I Street, NW - Suite 400 Washington, DC 20006 202-739-8085 / 202-785-1898 (fax) koc@nei.org

J. Michael Davis Duke Energy Corp. 526 South Church Street - M/C EC090 Charlotte, NC 28202 704-382-1784 / 704-382-3993 (fax) jmdavis0@duke-energy.com

Art Deardorff Structural Integrity Associates 3315 Almaden Expwy - Suite 24 San Jose, CA 95118 408-978-8200 / 408-978-8964 (fax) adeardor@structint.com

Shashi Dhar Niagara Mohawk PO Box 63 Lycoming, NY 13093 315-349-4732 / 315-349-1836 (fax) dhars@nimo.com

Hoang Dinh Duke Energy Corp. 13225 Hagers Ferry Road Huntersville, NC 28078 704-875-5675/704-875-4364 (fax) hvdinh@duke-energy.com

Frederic Dulcere EDF 6, Quai Watier 78401 Chatou France 33-130-877806 / 33-130-877323 (fax) frederic.dulcere@edf.fr Claude Faidy EDF 12-14, Avenue Dutrievoz 69628 Villeurbanne Cedex France 33-472-82-7279 / 33-472-82-7699 (fax) claude.faidy@edf.fr Yogen Garud APTECH Engineering Services 1282 Reamwood Ave. Sunnyvale, CA 94089 408-745-7000 / 408-734-0445 (fax) ygarud@aptecheng.com

Robin Dyle Southern Nuclear Operating Co. 40 Inverness Center Parkway Birmingham, AL 35242 205-992-5885 / 205-992-5793 (fax) rldyle@southernco.com

Glenn Gardner Northeast Utilities PO Box 128 Waterford, CT 06385 860-440-0373/860-440-2025 (fax) gardnga@gwsmtp.nu.com

xii

Colombe Gomane IPSN BP 6 92260 Fontenay aux Roses France 46-54-8887 / 47-46-1014 (fax) colombe.gomane@ipsn.fr Steve Gosselin Pacific Northwest National Laboratory 902 Battelle Blvd, MS-K526 Richland, WA 99352 509-375-4463/509-375-6497 (fax) stephen.gosselin@pnl.gov

Jeff Goodwin Southern Company Services PO Box 2625 Birmingham, AL 35202 205-992-6962/205-992-0324 (fax) jwgoodwi@southernco.com

Prasoon Goyal Toledo Edison 5501 North State Route 2 Oak Harbor, OH 43449 419-321-7351 / 419-249-2342 (fax) pkgoyal@firstenergycorp.com

Mark Gray Westinghouse Electric Co. PO Box 355 Monroeville, PA 15146 412-374-6481 / 412-374-6277 (fax) grayma@westinghouse.com

Michael Guthrie Carolina Power and Light Co. 3581 West Entrance Rd Hartsville, SC 29550 843-857-1049 / 843-857-1368 (fax) michael.guthrie@cplc.com

Makoto Higuchi IHI Co., Ltd. 1 Shin-nakahara, Isogo-ku Yokohama, 2358501 Japan 81-45-759-2194 / 81-45-759-2125 (fax) makoto_higuchi@ihi.co.jp Chris Hoffmann Westinghouse - CE Nuclear Power 2000 Day Hill Rd. - Dept 9483-1903 Windsor, CT 06095 860-285-4929 / 860-285-4232 (fax) hoffmannc@asme.org

Paul Hirschberg Structural Integrity Associates 3315 Almaden Expwy - Suite 24 San Jose, CA 95118 408-978-8200 / 408-978-8964 (fax) phirschb@structint.com

John Hoffman Vermont Yankee Nuclear Power Corp 546 Governor Hunt Rd Vernon, VT 05354-9766 802-451-3095 / 802-451-3030 (fax) john.hoffman@vynpc.com

Jon Hornbuckle Southern Nuclear Operating Co. 40 Inverness Center Parkway Birmingham, AL 35201 205-992-7974 / 205-992-6108 (fax) jehornbu@southernco.com

Ping Hsu Southern Nuclear Operating Co. 42 Inverness Parkway Birmingham, AL 35242 205-992-0989 / 205-992-0234 (fax) phsu@southernco.com

xiii

Karl Jacobs New York Power Authority 123 Main Street White Plains, NY 10601 914-739-5276 / 914-287-3710 (fax) jacobs.k@nypa.gov

Douglas Kalinousky US NRC M/S T10 E10 Washington, DC 20555 301-415-6788 / 301-415-5160 (fax) dnk@nrc.gov

Dietmar Kalkhof Paul Scherrer Institut CH-5232 Villigen PSI Switzerland 41-563-102620 / 41-563-102199 (fax) dietmar.kalkhof@psi.ch

Ariadni Kapsalopoulou New Jersey Dept. of Environmental Protection 33 Arctic Parkway - PO Box 415 Trenton, NJ 08625 609-984-7539 / 609-984-7513 (fax) akapsalopoulou@dep.state.nj.us

Nikolai Karpunin SEC NRS 14/23, Avtozavodskaya St. Moscow, 109280 Russia 795-275-5548 / 795-275-5548 (fax) opsvt@ntc.asvt.ru Suresh Khatri PG&E Co. PO Box 56 Avila, CA 805-545-4169 / 805-545-6515 (fax) sgk1@pge.com

Greg Kessel Dominion Generation (Virgina Power) 5000 Dominion Blvd. 1NE Glen Allen, VA 23060 804-273-3604/804-273-3614 (fax) gregory_kessel@dom.com

Dae Whan Kim Korea Atomic Energy Research Institute PO Box 105 Yusung, Taejon 305-600 Korea 82-42-868-2046 / 82-42-868-8346 (fax) dwkim1@kaeri.re.kr Wolfgang Korner Swiss Federal Nuclear Safety Inspectorate CH-5232 Villigen-HSK Switzerland 41-56-3103960/41-56-3103854 (fax) koerner@hsk.psi.ch

In Sup Kim Korea Advanced Inst. of Science & Technology 373-1 Kusong-dong Yusong-ku, Tajeon 305-701 Korea 82-42-869-3815 / 82-42-869-3810 (fax) iskim@sorak.kaist.ac.kr Jan Lagerstrom Ringhals AB 430 22 Varobacka Sweden 46-640-667624 / 46-340-660186 (fax) jalg@ringhals.vattenfall.se

Jean-Alain Le Duff Framatome EE/S - BALF0721 A Tour Framatome 92084 Paris France 33-479-63078 / 33-479-60501 (fax) jaleduff@framatome.fr

xiv

Sang Gyu Lee Korea Advanced Inst. of Science & Technology 373-1 Kusong-dong Yusong-ku, Tajeon Korea 82-42-869-3855 / 82-42-869-3895 (fax) s_youngll@kaist.ac.kr Bob Lisowyj Omaha Public Power District PO Box 399 Fort Calhoun, NE 68023 402-533-6491 / 402-533-7390 (fax) blisowyj@oppd.com

Young Ho Lee Korea Advanced Inst. of Science & Technology 373-1 Kusong-dong Yusong-ku, Taejon Korea 82-42-868-3855 / 82-42-868-3895 (fax) butcher@kaist.ac.kr Eduardo Maneschy Eletrobras Temonuclear S.A. - Eletronuclear Rua da Candelaria 65 - 6 Andar Rio de Janeiro, RJ 20091-020 Brasil 55-21-588-7681 / 55-21-588-7260 (fax) emanesc@eletronuclear.gov.br Eric Mathet OECD Nuclear Energy Agency Le Seine St-Germain - 12, boulevard des lles 92130 Issy-les-Moulineaux France 33-145-24-1057 / 33-145-24-1110 (fax) eric.mathet@oecd.org Frank Michel GRS Schwertnergasse 1 50667 Kln Germany 49-221-2068-753 / 49-221-2068-888 (fax) mif@grs.de Itaru Muroya Mitsubishi Heavy Industries 2-1-1 Shinhama Arai-cho Takasago Hyogo 676-8686 Japan 81-794-45-6800 / 81-794-45-6080 (fax) muroya@wq.trdc.mhi.co.jp Jeong Na Energy Services Group 8979 Pocahontas Trail Williamsburg, VA 23185 757-864-8471 / 757-864-4914 (fax) j.na@larc.nasa.gov

Ted Marston EPRI 3412 Hillview Ave. Palo Alto, CA 94304-1344 650-855-2997 / 650-855-2213 (fax) tmarston@epri.com

Klaus Metzner PreussenElektra Kernkraft 5 Tresckowstr Hannover D 30457 Germany 49-511-439-4009/49-511-439-4377 (fax) klaus-juergen.metzner@kernkraft.preussenelektra.de Todd Mielke Wisconsin Electric Power Co. 6610 Nuclear Rd. Two Rivers, WI 54241 920-755-6376 / 920-755-6032 (fax) todd.mielke@wepco.com

Joseph Muscara U.S. Nuclear Regulatory Commission Washington, D.C. 20555 301-415-5844 / 301-415-5074 (fax) jxm8@nrc.gov

xv

Takao Nakamura The Kansai Electric Power Co., Inc. 3-3-22, Nakanoshima, kita-ku Osaka 530-8270 Japan 81-6-6441-8221/81-6-6441-4277 (fax) takao_kepco@aol.com Robert Nickell Applied Science & Technology 16630 Sagewood Lane Poway, CA 858-485-9024 / 858-485-9024 (fax) rnickell@home.com

Gilles Navarro EDF 6, Quai Watier 78401 Chatou France 33-130-88571 / 33-130-877323 (fax) gilles.navarro@edf.fr Paul Norris Southern Nuclear Operating Co. Bin B063, PO Box 1295 Birmingham, AL 35242 205-992-5718 / 205-992-5793 (fax) ppnorris@southernco.com

Hitoshi Ohata Japan Atomic Power Company 1-6-1, Ohtemachi, Chiyoda-ku Tokyo 100-0004 Japan 81-3-3201-7087 / 81-3-3215-3930 (fax) hitoshi-ohata@japc.co.jp Susan Otto-Rodgers EPRI 1300 W.T. Harris Blvd. Charlotte, NC 28262 704-547-6072 / 704-547-6168 (fax) sjotto@epri.com

Frank Otremba University of Stuttgart Pfaffenwaldring 32 D-70569 Stuttgart Germany 49-711-685-2601 / 49-711-685-3053 (fax) frank.otremba@mpa.uni-stuttgart.de Michael Robinson Duke Energy Corp. 526 South Church Street, EC090 Charlotte, NC 28202 704-373-3522/704-382-3993 (fax) mrrobins@duke-energy.com

Eberhard Roos University of Stuttgart Pfaffenwaldring 32 D-70569 Stuttgart Germany 49-711-685-2604 / 49-711-685-3144 (fax) eberhard.roos@mpa.uni-stuttgart.de Guy Roussel AVN Avenue Du Roi 157 B-1190 Brussels Belgium 32-25-368-333 / 32-25-368-585 (fax) gr@avn.be

Stan Rosinski EPRI 1300 W.T. Harris Blvd. Charlotte, NC 28262 704-547-6123 / 704-547-6035 (fax) strosins@epri.com

Suresh Sahgal NOK, Nuclear Power Plant Beznau CH-5312 Doetingen Switzerland 41-56-2667-84/41-56-2667702 (fax) sah@nok.ch

xvi

Takeshi Sakai Japan Atomic Power Company 1-Banchi, Myojin-cho, Tsuruga-shi Fukui-ken Japan 81-770-26-1111 / 81-770-26-8081 (fax) takeshi-sakai@japc.co.jp Fred Simonen Pacific Northwest National Laboratory PO Box 999 Richland, WA 99352 509-375-2087- / 509-375-6497 (fax) fredric.simonen@pnl.gov

Andrew Siemaszko FirstEnergy Corp. 5501 North State Route 2 - M/C DB-1056 Oak Harbor, OH 43449 419-321-7341 / 419-249-2340 (fax) ajsiemaszko@firstenergycorp.com

Ram Singal Niagara Mohawk PO Box 63 (M/C ESB2) Lycoming, NY 13093 315-349-4480 / 315-349-1579 (fax) singalr@nimo.com

Larry Smith Constellation Nuclear Services, Inc. 1650 Calvert Cliffs Parkway Lusby, MD 20657 410-793-3416 / 410-793-3431 (fax) larry.d.smith@bge.com

Jussi Solin VTT Manufacturing Technology PO Box 1704 02044 VTT Finland 358-9-4566875 / 358-9-4567002 (fax) jussi.solin@vtt.fi Jean-Michel Stephan EdF Route De Sens Ecuelles F-77818 Moret-Sur-Loing Cedex France 33-1 60 73 60 85 / 33-1 60 73 65 59 (fax) jean-michel.stephan@edf.fr Bruce Swoyer PP&L, Inc. 2 North Nine Street (GENA62) Allentown, PA 18101-1179 610-774-7643 / 610-774-7830 (fax) bmswoyer@papl.com

Les Spain Virginia Power Co. 5000 Dominion Blvd Glen Allen, VA 23060 804-273-2602 / 804-273-3877 (fax) les_spain@dom.com

Yu-Shing Sun PECO Energy CB-63B-3 965 Chesterbrook Blvd. Wayne, PA 19087 610-640-6137 / 610-640-6582 (fax) ysun@peco-energy.com

Mike Testa FirstEnergy Corp. PO Box 4 Shippingport, PA 15077 724-682-5552 / 724-682-5536 (fax) testam@firstenergycorp.com

Raymond To Entergy Operations SR 333 Russellville, AR 72802 501-858-4371 / 501-858-4955 (fax) rto@entergy.com

xvii

Kazuya Tsutsumi Mitsubishi Heavy Industries 2-1-1 Shinhama Arai-cho Takasago Hyogo 676-8686 Japan 81-794-45-6723 / 81-794-45-6080 (fax) tsutsumi@wn.trdc.mhi.co.jp Bernard Van Sant Omaha Public Power District PO Box 399 Fort Calhoun, NE 68023 402-533-7385 / 402-533-6597 (fax) bvansant@oppd.com

Louis Van Der Wiel Ministry of Enviornment PO Box 90801 The Hague Netherlands 31-70-333-5234 lvdwiel@minszw.nl Edward Wais Wais & Associates 2475 Spalding Dr. Atlanta, GA 30350 770-396-8797 / 770-396-4194 (fax) waisassoc@aol.com

Mark Walker Dominion Generation 5000 Dominion Blvd Glen Allen, VA 23060 804-273-2226 / 804-273-3448 (fax) mark_walker@dom.com

Stan Walker EPRI 1300 W.T. Harris Blvd. Charlotte, NC 28262 704-547-6081 / 704-547-6168 (fax) swalker@epri.com

Dan Yuan Southern California Edison 14000 Mesa Rd., Bldg. G48B San Clemente, CA 92674 949-368-2478/949-368-2451 (fax) yuandd@songs.sce.com

xviii

CONTENTS

OPENING SESSION 1 EMBRACING THE FUTUREINTERNATIONAL CONFERENCE ON FATIGUE OF REACTOR COMPONENTS.................................................................................................... 1-1 2 US UTILITY PROGRAM FOR FATIGUE MANAGEMENT .................................................. 2-1 3 ACTIVITIES OF THE OECD-NEA IN THE AREA OF THERMAL FATIGUE IN LWR PIPING.................................................................................................................................... 3-1 4 FATIGUE OF REACTOR COMPONENTS: NRC ACTIVITIES ............................................ 4-1 5 A REGULATOR'S VIEW ON THE FATIGUE ISSUE ........................................................... 5-1 THERMAL FATIGUE I 6 THERMAL FATIGUE IN FRENCH RHR SYSTEM............................................................... 6-1 7 RESULTS OF THERMAL STRATIFICATION MEASUREMENTS AT NUCLEAR POWER PLANT BEZNAU...................................................................................................... 7-1 8 LEAKAGE FROM CVCS PIPE OF REGENERATIVE HEAT EXCHANGER INDUCED BY HIGH CYCLE THERMAL FATIGUE AT TSURUGA NUCLEAR POWER STATION UNIT 2 .................................................................................................................................... 8-1 9 OPERATING EXPERIENCE REGARDING THERMAL FATIGUE OF UNISOLABLE PIPING CONNECTED TO PWR REACTOR COOLANT SYSTEMS ...................................... 9-1 10 AUXILIARY FEEDWATER LINE STRATIFICATION AND COUFAST SIMULATION ..... 10-1 11 FATIGUE EVALUATION IN PIPING CAUSED BY THERMAL STRATIFICATION ......... 11-1

xix

THERMAL FATIGUE II 12 CONSIDERATION OF THERMAL FATIGUE AND CYCLE MONITORING OF A B31.1 PLANT ....................................................................................................................... 12-1 13 MAIN RESULTS OF EDFS EXPERIENCE ON IN SITU MEASUREMENTS RELATED TO THERMAL FATIGUE ON PWR REACTOR COOLANT PIPING ................... 13-1 14 EVALUATION OF OCONEE-2 HIGH PRESSURE INJECTION/NORMAL MAKEUP (HPI/NMU) LINE WELD FAILURE........................................................................................ 14-1 15 CURRENT ACTIVITIES ON GUIDELINES OF HIGH-CYCLE THERMAL FATIGUE IN JAPAN ............................................................................................................................. 15-1 FATIGUE MONITORING/EVALUATION 16 STATUS AND UPDATE OF THE EPRI FATIGUEPRO FATIGUE MONITORING PROGRAM ........................................................................................................................... 16-1 17 REMARKS TO THE DIFFERENT FACTORS INFLUENCING FATIGUE ANALYSIS AND FATIGUE DESIGN CURVES ....................................................................................... 17-1 18 THE RUSSIAN REGULATORY APPROACHES IN THE FATIGUE EVALUATION OF NPP CONSTRUCTION COMPONENTS......................................................................... 18-1 19 THE EVALUATION SYSTEM OF THERMAL STRATIFICATION STRESS USING OUTER SURFACE TEMPERATURE ................................................................................... 19-1 NDE/NDT 20 MICROSTRUCTURAL CHANGES OF PRESSURE VESSEL STEEL DURING FATIGUE IN HIGH TEMPERATURE WATER ENVIRONMENT ........................................... 20-1 21 MICROSTRUCTURAL INVESTIGATIONS AND MONITORING OF DEGRADATION OF LCF DAMAGE IN AUSTENITIC STEEL X6CRNITI 18-10 .............................................. 21-1 22 NDE TECHNOLOGY FOR DETECTION OF THERMAL FATIGUE DAMAGE IN PIPING.................................................................................................................................. 22-1 23 NDE METHOD FOR DETECTION OF FATIGUE DAMAGE ............................................ 23-1

xx

ENVIRONMENTAL FATIGUE I 24 ENVIRONMENTAL EFFECTS ON FATIGUE CRACK INITIATION IN PIPING AND PRESSURE VESSEL STEELS ............................................................................................ 24-1 25 MECHANISM OF FATIGUE CRACK INITIATION IN LIGHT WATER REACTOR COOLANT ENVIRONMENTS............................................................................................... 25-1 26 ENVIRONMENTAL FATIGUE EVALUATION ON JAPANESE NUCLEAR POWER PLANTS................................................................................................................................ 26-1 27 EVALUATION OF ENVIRONMENTAL EFFECTS ON FATIGUE LIFE OF PIPING......... 27-1 28 FATIGUE LIFE REDUCTION IN PWR WATER ENVIRONMENT FOR STAINLESS STEELS................................................................................................................................ 28-1 ENVIRONMENTAL FATIGUE II 29 AN APPROACH FOR EVALUATING THE EFFECTS OF REACTOR WATER ENVIRONMENTS ON FATIGUE LIFE.................................................................................. 29-1 30 DESIGN BASIS ENVIRONMENTAL FATIGUE EVALUATION AT OCONEE ................. 30-1 31 AN UPDATED METHOD TO EVALUATE REACTOR WATER EFFECTS ON FATIGUE LIFE FOR CARBON AND LOW ALLOY STEELS ............................................... 31-1 32 ENVIRONMENTAL FATIGUE, CRACK GROWTH RATES IN TITANIUM STABILIZED STAINLESS STEEL........................................................................................ 32-1 CODES AND STANDARDS 33 STRESS INTENSIFICATION FACTORS ......................................................................... 33-1 34 FATIGUE EVALUATIONS USING ASME SECTION XI NON-MANDATORY APPENDIX L ........................................................................................................................ 34-1 35 AN UPDATE ON THE CONSIDERATION OF REACTOR WATER EFFECTS IN CODE FATIGUE INITIATION EVALUATIONS FOR PRESSURE VESSELS AND PIPING.................................................................................................................................. 35-1 36 CODES AND THERMAL FATIGUE: STATUS AND ON-GOING DEVELOPMENT......... 36-1

xxi

VIBRATION/HIGH CYCLE FATIGUE 37 VIBRATION FATIGUE TESTING OF SOCKET WELDS, PHASE II ................................ 37-1

xxii

OPENING SESSION

1
EMBRACING THE FUTURE INTERNATIONAL CONFERENCE ON FATIGUE OF REACTOR COMPONENTS

Theodore U. Marston, Ph.D. Vice President and Chief Nuclear Officer 30 July 2000 Napa, CA

1-1

International Conference on Fatigue of Reactor Components


-Embracing the Future-

Theodore U. Marston, Ph.D. Vice President and Chief Nuclear Officer 30 July 2000 Napa, CA

1-3

Presentation Outline
Welcome and conference objectives US nuclear industry status Deregulation and privatization impacts EPRI nuclear status and future directions Future directions of the US nuclear program Role of fatigue in the current and future nuclear picture Concluding remarks

1-4

Current U.S. Industry Status


103 plants generating ~25% of the U.S. electricity in 1/00
Improved, risk-informed regulatory process Maintain safety, more effective, efficient and realistic Objective measures show positive trends Capacity factor--up; refueling outage duration--down; forced outage rate--down; operating costs--decreasing License renewal is a reality Process taking less than 2 years and less than $15M First renewal granted to Calvert Cliffs end of March 2000 Oconee granted 23rd of May, cost <$5/Kw(e) Public and stakeholder acceptance increasing Safety record Emission-free energy source and 1400 new Mw(e) Energy diversity and security
1-5

Current License Renewal Situation


Approved Calvert Cliffs 1,2 Oconee 1,2,3 Already filed Arkansas Nuclear One Unit 1 Hatch 1,2 2001 2000 Turkey Point 3,4 Catawba 1,2 McGuire 1,2 Peach Bottom 2,3 North Anna 1,2 Surry 1,2 2002 St. Lucie 1,2 Summer Crystal River 3 Fort Calhoun 2003 2003 Arkansas Nuclear One Unit 2 Cooper Farley 1,2 Robinson 2

1-6

Current U.S. Industry Status (cont.)


Extensive restructuring of generators
Successful efforts to reduce cost of doing business

Transfer of assets, through sale and barter Consolidation of nuclear operating companies
Alignment of operating companies

Joint purchasing agreements, some web-based Increased pressure on service providers to share risks and rewards, including EPRI Consolidation of suppliers and A/Es Retail electricity markets transition

1-7

U.S. Nuclear Picture


Active Companies
50

Merged

Exited Nuclear

45

40

35

Companies

30

25

20

15

10

1990
1-8

2000

2010

U.S. Nuclear Plant Transactions


600
Hydroelectric

500
Oil

Geothermal Coal 378 Gas

447

400

Price $/kW

300

200
136 117

100
30 8 44 21 16 NMP 1 NMP 2 Oyster Crk. Vt. Yankee Beaver Valley NYPA

0
Pilgrim TMI 1 Clinton

Transaction
1-9

Electricity Industry Restructuring (2005)


lVarious

Generators

lPower

Marketers

lAll

Customers

lLDCs lNational

Grid

Current 2000 Cal PX Revenues


$250,000,000 100.0% $225,000,000 90.0%

$200,000,000

80.0%

$175,000,000

70.0%

Daily Revenues

$150,000,000

60.0%

$125,000,000

50.0%

$100,000,000

40.0%

$75,000,000

30.0%

$50,000,000

20.0%

$25,000,000

10.0%

$0 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45 47 49 51 53 55 Ranked Revenues Days 57 1 3 5 7 9

0.0%

1-11

Revenues

EPRI Nuclear Market Share


(Equivalent Technology) U.S. market--97% membership Latin America--1/3 membership Asia--some project participation Europe--51% membership/40% participation
EPRI Participation North America, MWe = 119,199 95%

Europe (134,244 MWe)

Asia (51,358 MWe) MWe)


Full Members 75% Funders Non-participants
Funders Non-participants

40% 51%
Full Members Funders

25% 9%

5%

1-12

Why is EPRI Nuclear Successful in a Deregulated Business Environment?


The global nuclear fleet sinks or sails together
Safety must be maintained Good public relations is still mandatory Need to use international benchmarks, concepts and solutions to optimize the utilization of resources

The real competition is not other nuclear plants, but natural gas-fired combined-cycle plants No one country (or company) has sufficient resources to go it alone It seems to work, but EPRI has to work harder and smarter each year
1-13

EPRI Nuclear Power Business Organization Reflects Our Members Business


Environmental Benefits Management of Nuclear Risk and Resources in Commercial Nuclear Power Plants Profitable electricity and ancillary services sales, bi-lateral, short-term, bi-lateral, spot, etc

Cost Management

30%

Risk Communication

15%
Fuel Operations Maintenance Safety

10%
Radiation & Waste Management

30%

10%

Aging Management

Asset Management

Low Level Waste

Spent Fuel Management

BWR Materials

PWR Materials BOP Materials

Life Cycle Management

Decommissioning

Radiation Field Reduction

Capital Projects

1-14

Opportunities for New Nuclear


Various Scenarios in E-EPIC
2500.00
Nuclear Advanced Nuclear: Pres Admin Advanced Nuclear: 2XC Seq+CapEx Advanced Nuclear: Cap Ex

2000.00

1500.00

Capacity
1000.00 500.00 0.00
90 93 96 99 02 05 08 11 14 17 20 23 26 29 32 35 38 41 44 47 20 19 19 19 19 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 50

Year

1-15

EPRIs Future Direction in Nuclear


To focus on high value-added work for the existing plants and long term (>40 years) operation Shared risks and rewards Leveraged resources To expand the membership base internationally to bring even greater value to all members To conduct cooperative R&D involving all stakeholders To reduce our cost of doing business, similar to all of our members To expand the strategic part of our Nuclear R&D program to support the EPRI Energy Technology Roadmap and embrace the future
1-16

Conference Issues and Opportunities


Fundamental understanding of fatigue is essential for the success of current and future nuclear plants from safety and cost perspectives Collaboration and cooperation on a global scale is critical to our success, including sharing of data, test programs and theories Universal pressure to reduce R&D in general when we may be on the verge of a nuclear renaissance Nuclear programs are in different states in different regions of the world Electricity demand will grow, deregulation and privatization will be the norm leading to new generation requirements and real business opportunities
1-17

2
US UTILITY PROGRAM FOR FATIGUE MANAGEMENT
Michael R. Robinson Duke Energy Corporation Stan T. Rosinski John J. Carey EPRI Arthur F. Deardorff Structural Integrity Associates

presented at International Conference on Fatigue of Reactor Components July 31 August 2, 2000 Silverado Country Club Napa, California

2-1

US UTILITY PROGRAMS FOR FATIGUE MANAGEMENT

by

Michael R. Robinson Duke Energy Corporation Stan T. Rosinski John J. Carey EPRI Arthur F. Deardorff Structural Integrity Associates

presented at International Conference on Fatigue of Reactor Components July 31 August 2, 2000 Silverado Country Club Napa, California

Technical Paper SIT-00-007

2-3

US UTILITY PROGRAMS FOR FATIGUE MANAGEMENT


Michael R. Robinson Duke Energy Corp. 526 So. Church St. MC EC090 Charlotte, NC 28292 Stan T. Rosinski EPRI 1300 Harris Boulevard Charlotte, NC 28262 John J. Carey EPRI 3412 Hillview Ave. Palo Alto, CA 94304 Arthur F. Deardorff Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557

ABSTRACT Thermal fatigue is one of the main causes of leakage from reactor coolant system piping. Although nuclear plants were designed to accommodate the effects of cyclic operation, experience has shown that some loading conditions were not known at the time of initial plant design. As a result, fatigue related cracking and leakage is periodically encountered. In US plants, the response to cracking issues that are generic in nature are generally handled by vendor owner groups or through other industry-directed activities. Thermal fatigue in non-isolable branch piping in pressurized water reactors is currently being addressed by the Materials Reliability Project Thermal Fatigue Issue Task Group (MRP TF-ITG). The EPRI-managed project is targeted to release a Thermal Fatigue Management Guideline in late 2001. This project, with related background information, is described. BACKGROUND Metal fatigue is one of the degradation mechanisms that was explicitly addressed in the design of US nuclear power plants. In the formative years of nuclear plant design, it was recognized that there would be significant pressure/temperature cycling in nuclear power plants. As a result, Section III of the ASME Boiler and Pressure Vessel Code [1] was developing, requiring that the effects of cyclic operation on reactor vessels be considered. Requirements for nuclear piping system design were concurrently developed in USAS B31.7 [2], incorporated into ASME Section III in the early 1970s. Because of these requirements, modern reactor coolant pressure boundary components of nuclear plants were designed to meet Code requirements for a specified set of cycles, expected to be bounding at the time of the initial plant design. At most plants, the number of design transients was listed in plant Technical Specifications, with the number of design cycles being a limitation on plant operation. It was also recognized that inspection of nuclear systems for evidence of cracking or leakage would provide additional assurance of pressure boundary integrity. Thus, Section XI of the Boiler and Pressure Vessel Code [3] was prepared, mandating a program of visual, surface and volumetric inspection at key locations. In the original Section XI Code, high stress/high usage factor locations were given the highest priority for inspection. The plant design, operation and inspection requirements have generally been successful, in that leakage or cracking due to fatigue has not resulted from conditions considered in the original design. However, there have been a number of occurrences of fatigue due to conditions not known or expected to occur at the time of the original plant design. In each case, the nuclear industry has responded with research, inspections, analysis, and other actions to understand the reasons for the fatigue cracking and to put in programs place to manage the potential degradation due to fatigue.

2-4

OVERVIEW OF KEY THERMAL FATIGUE ISSUES The following provides an overview of some of the significant thermal fatigue issues that have affected more than a single plant. These show examples of how utilities have performed evaluations to understand and to resolve an issue and to develop an effective approach for managing fatigue. PWR Steam Generator Feedwater Nozzle Cracking: In many pressurized water reactor (PWR) plants, the feedwater nozzles on the steam generators are used for both the main feedwater and for auxiliary feedwater. During hot standby operating conditions, flow variations of the auxiliary feedwater resulted in cycling stresses due to feedwater nozzle stratification, leading to cracking in the nozzle-to-safe-end weld region [4, 5]. Utilities have responded with improved operational techniques, modified feedwater sparger inlets and increased inspections to manage this issue. BWR Reactor Vessel Feedwater Nozzle Cracking: In early boiling water reactor (BWR) plants, the inlet to the vessel internal feedwater sparger was a slip-fit configuration. Due to leakage past the sparger inlet, high cycle fatigue initiated cracks on the nozzle bore and blend radius regions [6]. A special program was initiated by the utilities to understand the cause of the cracking and to develop a modified design and loading specification that could be used for plant modifications. Many utilities installed a monitoring program to detect for leakage. With the improved designs and with more operating experience, the issue is currently managed by nozzle inspection, supported by plant-specific flaw tolerance analysis to justify the inspection interval. BWR Control Rod Drive Return Nozzle Cracking: In early BWR plants, a reactor vessel nozzle (approx. 3-inch diameter) was provided to return control rod drive water to the reactor vessel. Cracking in these nozzles was discovered, due to low flow cycling and turbulent mixing with the reactor downcomer water [6]. Utilities responded by showing that the line was not needed and capping the nozzle to avoid the thermal mixing that lead to the fatigue cracking. B&W Plant Safety Injection/Makeup Nozzle Cracking: Through-wall leakage has occurred in the nozzle/safe-end region at two plants [7, 8]. This cracking occurs as a result of loosening of the nozzle thermal sleeve, probably due to 1) low normal flow fluctuations allowing hot/cold temperature cycling in the nozzle region and 2) flow induced vibration. The issue is currently managed by plant operations that assure adequate flow in the makeup line and an inspection program that assures that there is no loosening of the thermal sleeves in the nozzle. Farley Safety Injection Piping : NRC Bulletin 88-08 [9] notified PWR operators that certain lines might be susceptible to cracking as a result of valve leakage. Inleakage of cold water from the charging system caused a through-wall leak in Safety Injection piping at the Farley plant. Related leakage had also occurred in foreign plants (including a failure due to out-leakage from an RHR suction line). As a result of this leakage, plant-specific assessments were conducted at all plants. EPRI initiated the TASCS (Thermal Stratification, Cycling and Striping) program [10], performing testing and evaluations to better understand the sources of stress cycling that can occur in normally stagnant branch lines attached to reactor coolant piping. Many utilities implemented monitoring programs, some of which are still in place to assure that significant leakage is not occurring. PWR Surge Line Stratification: At several plants, excess movements were observed in pressurizer surge lines, indicating that the lines were significantly stratified. Bulletin 88-11 [11] was issued, prompting utility Owner Groups to perform testing at a number of plants to quantify the loadings and number of stratification cycles occurring. The

2-5

issue was resolved based on Owners Group analyses for each of the specific vendor types of reactors. For this issue, no cracking or leakage was ever observed. Drain Line Leaks: At TMI-1 and Oconee-1, through-wall leakage has been observed in small stagnant drain lines attached to reactor coolant piping. Testing at TMI showed that cyclic turbulent penetration into the small uninsulated line was producing stratification locally at the elbow, although there was evidence of a preexisting defect at the weld between the elbow and horizontal piping where the leakage occurred. At Oconee, the cracking occurred in the elbow and no pre-existing defects were identified. A similar crack occurred in a foreign plant. RECENT INDUSTRY PROGAMS In addition to Owners Group activities that addressed some of the specific issues noted in the previous section, EPRI has been involved in a number of industry- supported programs to address fatigue. In addition, ASME Section XI has been active in developing Appendix L [12] that specifically provides guidelines for evaluating potential fatigue issues. EPRI TASCS Program The Thermal Stratification, Cycling and Striping (TASCS) program was initiated by EPRI soon after the issuance of NRC Bulletin 88-08 [9]. The objective of the program was to develop models, correlations and guidelines to evaluate TASCS phenomena in reactor coolant piping systems. Based on a survey of existing literature, testing, and plant monitoring results that existed at that time, the TASCS program was developed to conduct additional testing and to develop correlations and models to predict the thermal-hydraulic conditions in nominally stagnant lines attached to reactor coolant systems. The testing program produced a database for correlating turbulence penetration into branch piping, which was identified as a dominant mechanism contributing to thermal cycling. Analytical efforts led to models that would predict stratification height and heat transfer effects in the presence of stratified flow. A handbook was developed that provides screening criteria and evaluation methodology for piping systems potentially affected by TASCS [10]. Key subjects covered in the included: - Height of Stratified Flows - Heat Transfer in the Presence of Leakage Flow - Turbulence Penetration Length and Thermal Cycling - Thermal Striping - Free Convection and Stratification Heat Transfer - Test Results, Correlations and Data Analysis - Example Problems. The final report was issued in early 1994. This handbook has been used by many utilities as a basis for evaluating branch piping systems to determine the susceptibility to TASCS. The screening criteria have been incorporated into risk-informed inservice inspection programs [13]. EPRI Fatigue Management Program As an expansion of the TASCS program to other areas of operating nuclear plants, and to cover all types of fatigue mechanisms that might be encountered in plants, EPRI initiated a project to produce a Fatigue Management Handbook in 1992. The major developments of this program were: - A comprehensive handbook provided to help utilities implement fatigue management programs - Screening criteria for thermal and vibrational effects - Methods for managing fatigue degradation and its impact on plant availability - Cost-effective methods to ensure and extend component fatigue life

2-6

The resulting handbook was provided in four volumes, including: - Volume 1: Fatigue concepts, regulatory issues and experience summary - Volume 2: Thermal/vibrational fatigue screening criteria with operating plant fatigue database - Volume 3: Corrective Measures for fatigue problems - Volume 4: Bibliography and copies of key fatigue references As part of the project, the operating plant fatigue database was developed, showing that a large number of fatigue failures were mechanical, primarily resulting from vibration at small socket weld connections. This finding lead to further research to investigate the fatigue aspects of socket-welded fittings[14]. ASME Section XI Activities In the early 90s, the ASME Section XI Working Group on Operating Plant Criteria began efforts to develop criteria that could be used in operating plants to justify continued operation when a fatigue concern was identified. At the time, utilities were struggling with how to handle issues such as 1) predicted fatigue usage factor greater than 1.0, 2) number of cycles greater than considered in design, or 3) transients encountered with severity greater than considered in design. The outcome of this effort was two reports published by EPRI [15, 16] and the recently added Appendix L to ASME Section XI. MRP THERMAL FATIGUE PROGRAM The Materials Reliability Project (MRP) Thermal Fatigue Industry Issue Group (TF-ITG), is currently developing Thermal Fatigue Management Guidelines (TFMG) that can be used by utilities to address potential thermal fatigue in piping systems that are non-isolable from the reactor coolant system piping in PWRs. The MRP thermal fatigue project is composed of 13 separate tasks. Figure shows the interrelationship between the tasks and how the outcome of each tasks relates to assessing the potential for thermal fatigue or performing additional inspections, monitoring, maintenance, or plant modifications. The overall schedule for the project is shown in Figure 2. Tasks 1 and 2 Project Management This is the overall effort by EPRI to manage the project. Under EPRI leadership, the project is to be completed on time and within the budget established by the member utilities. This task also includes meetings with industry and Nuclear Regulatory Commission to assure that the program outputs will be acceptable for use in performing plant evaluations at project completion. Task 3 Industry Operating Experience Existing industry failure and leakage databases are assembled primarily based on leaks or identified flaws. Collection of industry experience and sharing of monitoring data and information of events that do not meet some higher reporting threshold will provide added value and new insights into where thermal fatigue may be a problem and where it is definitely not a problem. A database is being developed that will collect utility observations regarding plant monitoring or events that do not reach a level requiring reporting. This database will be available on the EPRI web site. It will contain historical experience and will include provisions to capture and make available similar information in the future. Task 4 Thermal Fatigue Screening This task is expanding on the EPRI TASCS models and results from other industry programs to develop improved fatigue screening methodology. The screening approach will identify the thermal fatigue phenomena that might be possible, factors necessary for thermal fatigue to occur, and the logic process for making this determination. When screening identifies a location as susceptible to thermal fatigue, a conservative prediction of the thermal loading will be made, determining an allowable operating time prior to the cumulative usage factor exceeding 1.0. The methodology will have a technically defensible basis and will be validated against known instances of thermal fatigue failures.

2-7

Task 5 Thermal Fatigue Monitoring Guidelines This task will provide guidance to utilities in implementing effective monitoring when analytical means and screening show that thermal fatigue may be an issue. All types of montoring will be discussed. State-of-the-art monitoring technology will be identified. Criteria for effective placement of monitoring sensors will be developed. Methods for interpretation of monitoring data will be provided. Most importantly, criteira will be provided for discontinuance of monitoring programs. The outcome will be practical guidance on the use of monitoring to detect potential thermal fatigue phenomena. Task 6 NDE Inspection Guidelines This task is assembling previous guidance on nondestructive examination (NDE) methodologies such as (RT) or ultra sonics (UT) and evaluating potential new technologies. Recommendations will be developed for specific NDE technology and variables, especially for small diameter piping. Recommendations will be made on the appropriate qualification of NDE examiners and procedures. The recommended means of evaluating NDE data and reporting levels are also being provided. Contacts were made internationally to determine any difficulties in applying NDE for thermal fatigue and to identify laboratory investigations to verify NDE performance. Research has been completed on determining capabilities for producing thermal fatigue cracks. Based on this research, mockups were designed and fabricated. A set of mockups with individual cracks due to thermal fatigue, along with mockups containing thermal craze cracking, were fabricated and used in qualification of techniques for detecting thermal fatigue cracking. The outcome of this effort will be guidance on NDE methodologies and recommendations of specific NDE technology and variables to use when inspecting for suspected thermal fatigue damage. As an adjunct to this task, an interactive training module is being developed to train inspectors on the characteristics of thermal fatigue and how to detect it using the recommended methodologies. Task 7 Plant Operations/Maintenance Guidelines This task focuses on how operations and maintenance (O&M) practices can lead to thermal fatigue damage and how potential changes can eliminate or minimize thermal fatigue damage potential. PWR operating experience will be used to identify plant practices and corrective actions implemented at affected plants. Guidance will be developed for use by utility engineers to aid in identifying O&M practices which may lead to fatigue damage and what actions may be taken to mitigate the consequences. A key area for the research will be valve maintenance. This will result in guidance for plant personnel to identify how O&M practices can create or minimize the potential for thermal fatigue cracking. Task 8 Thermal Fatigue Evaluation This task will develop analytical approaches that may be used to evaluate components where screening can not be conservatively used to demonstrate an acceptable fatigue life. The evaluation methodology will include both a simplified disposition evaluation and a more rigorous analysis guide to aid engineers in evaluating thermal fatigue situations. The analysis guide will document the techniques and describe comprehensive methodologies for analytical reconciliation of thermal fatigue phenomena using or based on ASME Section III methodology. Task 9 Plant Modification Guidelines This task will identify and describe plant modifications that might be implemented for avoiding thermal fatigue. Implementation of cost effective plant changes could potentially eliminate the need for future monitoring and augmented piping inspections where there is significant probability of future cracking.

2-8

Task 10 International Technical Exchange This task will identify and pursue participation in important foreign R&D activities that could contribute to resolution of thermal fatigue issues. This activity is intended to provide awareness of and access to foreign information of value to US utilities for detection, assessment, mitigation, and prevention of thermal fatigue damage. Similarly, this activity can provide information to non-domestic utilities to resolve fatigue issues. The international workshop on thermal fatigue at which this paper is presented is one of the outcomes of this task. Task 11 Thermal Fatigue Management Guidelines This task represents the principal product of the thermal thermal fatigue project. It will assemble the results of the other tasks, document conclusions drawn from that work, and provide recommendations for managing thermal fatigue. An interim version of the guideline will be provided prior to development of the final products from Tasks 4 and 8. The Thermal Fatigue Management Guideline will be a compilation of methods for assessment, screening, monitoring, analysis, and remediation and/or management of thermal fatigue. Task 12: Training This task will develop training applying the Thermal Fatigue Management Guideline. This training will be aimed at utility personnel (operations, maintenance, systems and design engineering) to increase their overall knowledge and awareness of cyclic thermal fatigue and how plant operations and maintenance may be contributors to the phenomena. This will result in more knowledgeable and experienced personnel and more effective management of thermal fatigue issues. Task 13: ASME Activities Monitoring This task provides for monitoring ASME Section III and XI activities related to thermal fatigue and inspection. Inspection guidance developed by the project will be reviewed with appropriate Code groups to determine if such guidance should form the basis for future Code revisions. Other Activities The MRP TF-ITG has just undertaken the responsibility to provide further review of EPRI efforts to conduct fatigue related research in other areas. The specific scope of these activities is under development. LICENSE RENEWAL ACTIVITIES EPRI has conducted several projects in the past to support plants preparing for license renewal. In earlier projects, studies were performed to evaluate the effects of light water reactor environments on reactor coolant components in PWRs and BWRs based on design and plant monitoring data [17-22]. A project has just been initiated to develop a consensus criteria for preparation of license renewal applications. This will develop a common approach for addressing fatigue in license renewal submittals, eliminating some of the uncertainties and providing costs savings for utilities. CONCLUSIONS Fatigue is not a new issue for US utilities with operating nuclear plants. Even though nuclear plants were designed with consideration of the potential effects of fatigue, operating experience periodically identifies an issue that was not considered in the design phase. As a result, various vendor and industry research programs and ASME Code activities have been undertaken in recent years to understand and resolve fatigue issues. With the advent of deregulation of the utility industry, utility personnel must assure that fatiguerelated leakage does not occur, both from the standpoint of plant safety and economics. The current MPR thermal fatigue project is one aimed at assuring that leakage due to thermal fatigue in non-isolable piping systems does not occur. Planning is underway for additional programs to support goals of minimizing

2-9

fatigue failures in the future, and showing that fatigue can be adequately managed during an extended operating period. REFERENCES 1. ASME Boiler and Pressure Vessel Code Section III, Nuclear Vessels, 1963. 2. USA Standard Code for Pressure Piping, Nuclear Power Piping, USAS B31.7-1969. 3. Draft ASME Code for Inservice Inspection of Nuclear Reactor Coolant Systems, ASME, October 1969 (Predesessor of ASME Section XI). 4. U.S. Nuclear Regulatory Commission, Cracking in Feedwater System Piping, IE Bulletin 79-13, Revision 2, October 16, 1979. 5. U.S. Nuclear Regulatory Commission, Investigation and Evaluation of Cracking Incidents in Pressurized Water Reactors, NUREG-0691, September, 1980. 6. U.S. Nuclear Regulatory Commission, BWR Feedwater Nozzle and Control Rod Drive Return Line Nozzle Cracking, NUREG-0619, April, 1980. 7. U.S. Nuclear Regulatory Commission, Cracking in Piping of Makeup Coolant Lines at B&W Plants, IE Information Notice No. 82-09, March 31, 1982. 8. U.S. Nuclear Regulatory Commission, Unisolable Crack in High Pressure Injection Piping, Information Notice 97-46, July 9, 1997. 9. U.S. Nuclear Regulatory Commission, Thermal Stresses in Piping Connected to Reactor Coolant Systems, NRC Bulletin No. 88-08, June 22, 1988, Supplement 1, June 24, 1988, and Supplement 2, August 4, 1988. 10. TR-103581, Thermal Stratification, Cycling and Shaping (TASCS); EPRI, March 1994. 11. U.S. Nuclear Regulatory Commission, Pressurizer Surge Line Thermal Stratification, NRC Bulletin No. 88-11, December 20, 1988. 12. Appendix L to ASME Boiler and Pressure Vessel Code Section XI, Operating Plant Faituge Assessment, 1995 Edition, 1996 Addenda. 13. TR-112657, Risk-Informed Inservice Inspection Evaluation Procedure, Rev. B-A, EPRI, December 1999. 14. TR-107455, Vibration Fatigue of Small Bore Socket-Welded Pipe Joints, EPRI, June 1997. 15. TR-100252, Metal Fatigue in Operating Nuclear Power Plants, EPRI, April 1992. 16. TR-104691, Operating Nuclear Power Plant Fatigue Assessments, EPRI, April 1995. 17. TR-107515, Evaluation of Thermal Fatigue Effects on Systems on Systems Requiring Aging Management Review for License Renewal for the Calvert Cliffs Nuclear Power Plant,, January 1998. 18. TR-105759, An Environmental Factor Approach to Account for Reactor Water Effects in Light Water Reactor Pressure Vessel and Piping Fatigue Evaluations, December 1995. 19. TR-110043, Evaluation of Environmental Fatigue Effects for a Westinghouse Nuclear Power Plant, April 1998. 20. TR-110356, Evaluation of Environmental Thermal Fatigue Effects on Selected Components in a Boiling Water Reactor Plant, April 1998. 21. TR-110356, Environmental Fatigue Evaluations of Representative BWR Components, May 1998. 22. TR-110356, Effect of Environment on Fatigue Usage for Piping and Nozzles at Oconee Units 1,2, and 3, December 1999.

2-10

Events Detected Failures / Flaws Occur Research Valve Testing 3, 10

Screening Tool 4

Screen components

Not Vulnerable

Vulnerable

Knowledge & Experience 12

Maintenance/ Ops Guide 7

Try to Eliminate TF Initiator

Success

Not Successful

Monitoring Guidelines 5

Monitoring Predictive Tool 4 Predict TF Loading (temp, rates, flows, cycles)

Delivered ITG Tasks

License Action

Disposition Tool 8

Try Simplified Tool

Success

Not Successful Analysis Methods Guide 8 Inspection Guidelines 6

Rigorous Analysis

Full Plant Life Confirmed (CUF<1.0)


(1)

Only Partial Plant Life Predicted

Repair/ Replace Guidelines 9

Permanent Monitoring

or

Recurring or Inspections (Flaw Tolerance)

Repair/Replace Mod

OK Continue Operations

(1) Permanent/semi-permanent monitoring may be recommended to confirm predicted significance

Figure 1. Project Flowchart

2-11

Thermal Fatigue ITG Level 1 Schedule


ID Task Name Start 7/28/99 1/3/00 10/1/99 1/3/00 9/1/99 1/3/00 TBD 1/3/00 10/1/99 6/1/00 6/1/00 9/1/01 8/2/99 Finish 11/1/01 11/30/00 8/31/01 10/31/01 9/29/00 10/31/00 TBD 10/31/00 10/1/01 10/31/01 8/31/00 11/1/01 8/31/01 1999 2000 2001 A S O N D J F M A M J J A S O N D J F M A M J JA S O N

1 & 2 Project Management 3 4 5 6 7 8 9 Industry Operating Experience Thermal Fatigue Screening TF Monitoring Guidelines TF Inspection Guidelines Plant O&M Guidelines Thermal Fatigue Evaluation Plant Modification Guidelines

10 International Technology Exchange 11a Interim Fatigue Management Guidelines 11b Thermal Fatigue Management Guidelines 12 Develop & Delivery of Training 13 ASME Section XI WG Activities

Figure 2. Project Schedule

2-12

US UTILITY PROGRAMS FOR FATIGUE MANAGEMENT


Authors: Arthur F. Deardorff* Structural Integrity Associates Stan T. Rosinski and John J. Carey EPRI Michael R. Robinson Duke Energy Corporation International Conference on Fatigue of Reactor Components Napa, California USA July 31 August 2, 2000 *Presenting Author adeardor@structint.com

2-13

OUTLINE

Welcome Background and Historical Perspective Recent Industry Activities Materials Reliability Project (MRP) Fatigue Activity

2-14

BACKGROUND

Nuclear Plant Reactor Coolant Systems Designed for Cyclic Operation


Vessels ASME Section III Piping USAS B31.7/ASME Section III

Key Locations Inspected Per ASME Section XI However, Fatigue-related Failures Occur

Loadings not known at time of initial design Thermal and vibration

Industry Has Responded to Manage the Issues


2-15

REVIEW OF KEY THERMAL FATIGUE ISSUES


PWR Steam Generator Feedwater Nozzle Cracking 1980 Cracking at Nozzle/Safe-end Region Caused by Flow Cycling/Stratification

Flow cycling of auxiliary feedwater Counterbore geometry caused high local stresses Improved operations More frequent inspection Modified designs

Managed by:

2-16

REVIEW OF KEY THERMAL FATIGUE ISSUES


BWR Feedwater Nozzle Cracking 1980 Cracking In Nozzle Bore/Blend Radius Area Caused by High Cycle Fatigue

Bypass of cold leakage flow past sparger inlet Modified designs Leakage monitoring Inspection interval based on flaw tolerance evaluation

Managed by:

2-17

REVIEW OF KEY THERMAL FATIGUE ISSUES


B&W Safety Injection/Makeup Lines 1982 / 1997 Cracking at Safe-End / Piping Weld Caused by Loosening of Thermal Sleeve

Flow induced vibration Flow cycling allowed hot RCS water into region Inspection for thermal sleeve tightness Maintaining adequate flow in nozzle

Managed by:

2-18

REVIEW OF KEY THERMAL FATIGUE ISSUES


Safety Injection Piping 1988 + Others in Non-US Plants NRC Bulletin 88-08 / Cracking in Elbow/Horizontal Pipe Region Caused by Cold Water Inleakage/Turbulence Penetration Managed by:

Monitoring Valve maintenance EPRI TASCS Program

2-19

OTHERS

BWR Control Rod Drive Nozzle Cracking

Flow to nozzle eliminated

PWR Surge Lines

Re-evaluated per NRC Bulletin 88-11

Drain Lines

Current MRP issue

2-20

EPRI PROGRAMS
Thermal Stratification, Cycling and Striping (TASCS) Program

Initiated to address Bulletin 88-08 issues Performed testing/analysis to address stratification and turbulence penetration Final report in 1994 Followed TASCS Program Looking at Remainder of Plant Also addressed vibrational fatigue Led to further programs
socket weld fitting fatigue testing improved stress indices/stress intensification factors
2-21

EPRI Fatigue Management Handbook


ASME SECTION XI ACTIVITIES


Started Efforts about 1990 to Develop Code Approaches for Addressing Fatigue Concerns

Recognized that stress reports produced to show CUF<1.0 Stressed component re-analysis

reduce conservatisms use actual transients

Allowed fatigue monitoring Included flaw tolerance assessment as an alternate to usage factor analysis

Appendix L Approved in 1996 Efforts Continuing to Gain Regulatory Acceptance

2-22

MATERIALS RELIABILITY PROJECT THERMAL FATIGUE PROGRAM


MRP TF-ITG: Thermal Fatigue Issue Task Group Approved by Utilities in Mid-1999 EPRI Managed Program

Providing guidelines for managing thermal fatigue Addressing non-isolable PWR piping

Mike Robinson is Task Group Chairman

2-23

MRP-TF PROJECT GOAL AND SCOPE

Provide EPRI member utilities with a consistent set of guidelines and methodology for addressing piping thermal fatigue issues in 2001 Scope is thermal fatigue issues for those portions of ASME class 1 piping systems that are:

Connected to the reactor coolant pressure boundary Greater than 1 diameter Not isolable from the reactor coolant pressure boundary

Similar to NRC Bulletin 88-08 scope

2-24

STRATEGIC APPROACH

Respond to NRC request for industry leadership and action Provide the industry ability to manage thermal fatigue concerns through current license life and any renewal periods

validate methodology against known failures

Take advantage of earlier thermal fatigue work by EPRI and supplement it with new ideas and enhanced thermal fatigue management capabilities

2-25

MRP-TF TASK 3: INDUSTRY OPERATING EXPERIENCE

Description: Existing industry databases are currently


based on leaks or identified flaws

Data from plant monitoring or other experience also being collected

Outcome: A database that collects utility experience and


observations which do not reach the level of recording through LERs or other means

Data to be available on EPRI web site Process to be established for continued update

2-26

MRP-TF TASK 4: THERMAL FATIGUE SCREENING


Description: This task will review existing program
results for needed enhancements for improved screening and evaluation capability

Thermal fatigue phenomena Factors for thermal fatigue to occur Logic process for screening Load magnitudes/cycles to be conservatively estimated

Outcome: Technically defensible methods for


determining when and where significant thermal fatigue damage may occur in PWR piping systems

2-27

MRP-TF TASK 5: THERMAL FATIGUE MONITORING GUIDELINES


Description: This task will provide guidance for
monitoring

Basis for implementing a monitoring program Identification of state-of-the-art monitoring technology Effective placement of monitoring sensors Interpretation of monitoring data Basis for discontinuing monitoring

Outcome: Practical guidance on the use of monitoring to


detect potential thermal fatigue phenomena

2-28

MRP-TF TASK 6: NDE INSPECTION GUIDELINES


Description: This task will assemble previous guidance
on NDE methodologies (such as RT or UT) and make recommendations for specific NDE technology and variables

Mockups constructed representing thermal fatigue in small piping Various advanced NDE approaches evaluated Special considerations for detecting thermal fatigue cracking and crazing investigated

Outcome: Guidelines on NDE methodologies and


recommendation of specific NDE technology and variables to use when inspecting locations for suspected thermal fatigue damage
2-29

MRP-TF TASK 7: PLANT O&M GUIDELINES

Description: This task focuses on how O&M practices


can lead to thermal fatigue damage and identifies necessary changes to eliminate the damage potential

Valve maintenance Operations generally effects systems normally used, such as charging and auxiliary spray

Outcome: Guidance for use by plant personnel to identify


how O&M practices can create or minimize the potential for causing thermal fatigue events

2-30

MRP-TF TASK 8: THERMAL FATIGUE EVALUATION


Description: This task will develop both a simplified
evaluation methodology and a more rigorous analysis guide to aid engineers in evaluating situations where thermal fatigue may potentially be occurring

Outcome: An analysis guide, documenting the


techniques and describing comprehensive methodologies for analytical reconciliation of thermal fatigue phenomena using or based on ASME Section III methodology, will be developed for evaluating location subject to cyclic thermal loadings

2-31

MRP-TF TASK 9: PLANT MODIFICATION GUIDELINES

Description: This task will identify and describe plant


modifications that may be used for avoiding the potential for thermal fatigue

Outcome: Guidelines for identification of cost effective


plant changes that would eliminate need for future monitoring and piping augmented inspections

2-32

MRP-TF TASK 10: INTERNATIONAL TECHNICAL EXCHANGE


Description: This task will support identification of and
possible participation in important foreign R&D activities which could contribute to resolution of the thermal fatigue issue

Outcome: Awareness of and access to foreign


information that could be of value for detection, assessment, mitigation, and prevention of thermal fatigue damage (This workshop is an outcome of this task)

2-33

MRP-TF TASK 11: THERMAL FATIGUE MANAGEMENT GUIDELINES


Description: This task represents the principal product of
this project, assembling the results of the other tasks, documenting conclusions drawn from that work, and providing recommendations for managing thermal fatigue

Outcome: The TFMG will be a compilation of methods


for assessment, screening, monitoring, analysis, and remediation for and management of thermal fatigue
The ITG will seek NRC staff acceptance of guideline. An interim TFMG will be issued later this year. Final TFMG available late 2001.

2-34

MRP-TF TASK 12: DEVELOP & DELIVER TRAINING


Description: This task develops and delivers the training
for utility engineers and others in applying the TFMGand will be provided so that utility personnel (operations, maintenance, engineering) can increase their overall knowledge and awareness of cyclic thermal fatigue and how plant operations and maintenance may be contributors to the phenomena

Outcome: More knowledgeable and experienced


personnel and more effective management of potential thermal fatigue issues

2-35

PROJECT FLOWCHART

Events Detected Failures / Flaws Occur Research Valve Testing 3, 10

Screening Tool 4

Screen components

Not Vulnerable

Vulnerable

Knowledge & Experience 12

Maintenance/ Ops Guide 7

Try to Eliminate TF Initiator

Success

Not Successful

Monitoring Guidelines

Monitoring Predictive Tool 4 Predict TF Loading (temp, rates, flows, cycles)

Delivered ITG Tasks

License Action

Disposition Tool 8

Try Simplified Tool

Success

Not Successful Analysis Methods Guide 8 Inspection Guidelines 6

Rigorous Analysis

Full Plant Life Confirmed (1) (CUF<1.0)

Only Partial Plant Life Predicted

Repair/ Replace Guidelines 9

Permanent Monitoring

or

Recurring or Inspections (Flaw Tolerance)

Repair/Replace Mod

OK Continue Operations

(1) Permanent/semi-permanent monitoring may be recommended to confirm predicted significance

2-36

PROJECT SCHEDULE
ID Task Name Start 7/28/99 1/3/00 10/1/99 1/3/00 9/1/99 1/3/00 TBD 1/3/00 10/1/99 6/1/00 6/1/00 9/1/01 8/2/99 Finish 11/1/01 11/30/00 8/31/01 10/31/01 9/29/00 10/31/00 TBD 10/31/00 10/1/01 10/31/01 8/31/00 11/1/01 8/31/01 1999 2000 2001 A S O N D J F M A M J J A S O N D J F M A M J J A S O N

1 & 2 Project Management 3 4 5 6 7 8 9


Industry Operating Experience

Thermal Fatigue Screening

TF Monitoring Guidelines

TF Inspection Guidelines

Plant O&M Guidelines

Thermal Fatigue Evaluation

Plant Modification Guidelines

10 International Technology Exchange 11a Interim Fatigue Management Guidelines 11b Thermal Fatigue Management Guidelines 12 Develop & Delivery of Training 13 ASME Section XI WG Activities

2-37

MRP LICENSE RENEWAL SUPPORT


Recent Addition to Scope Established to Prepare Criteria for Addressing Environmental Effects of Fatigue Beyond 40 Years Scope and Approach Under Development

2-38

SUMMARY

Fatigue in Operating Plants is Not a New Issue


Utilities have generally been effectively dealing with fatigue as issues surface Industry groups respond to perform research/evaluation as required

Even though fatigue not identified as safety issue, utilities need to take appropriate actions to avoid forced outages due to fatigue failure

2-39

3
ACTIVITIES OF THE OECD-NEA IN THE AREA OF THERMAL FATIGUE IN LWR PIPING
E. Mathet OECD Nuclear Energy Agency Le Seine St Germain 12 boulevard des Iles F-92130 Issy-les-Moulineaux

3-1

Activities of the OECD-NEA in the area of Thermal Fatigue in LWR Piping

E. Mathet OECD Nuclear Energy Agency Le Seine St Germain 12 boulevard des Iles F-92130 Issy-les-Moulineaux

Abstract The OECD Nuclear Energy Agency has 27 Member countries. Under the auspices of the Nuclear Safety Division are two senior committees dealing with regulatory aspects (CNRA) and technological aspects (CSNI). Under these two committees activities relevant to thermal fatigue in LWR piping are carried out under the two Working Groups (WG's): WG on Operating Experience, WG on Integrity of Components and Structures. There is also co-operation with the former task group on thermalhydraulics application. These groups make recommendations to the senior committees. WG on Operating Experience is mainly concerned with the analysis of safety significant incidents, but it undertakes special studies as well. WG on Integrity of Components and Structures (Integrity and Ageing, IAGE) has activities in fracture mechanics, NonDestructive Examination and material degradation, as being the three aspects of structural integrity for metal reactor components. It also has sub-groups dealing with the ageing of concrete structures and the seismic behaviour of structures. In the area of thermal fatigue, it has already organised a Specialist Meeting in Paris in 1998, and it currently has a programme of work approved by CSNI. This paper provides background on the Agency itself, while focusing on the thermal fatigue activities within NEA carried out mainly through the Committee on the Safety of Nuclear Installations and the groups of technical experts on operating experience, structural integrity, and thermal hydraulic.

3-3

INTRODUCTION The Nuclear Energy Agency (NEA) is one of the 15 bodies that make up the Organisation for Economic Co-operation and Development (OECD), located in the Paris area in France. The Members of the OECD/NEA are a group of 27 like -minded, developed countries which at the end of 1999 operated 348 reactor units. Today, nuclear
energy accounts for 25 per cent of all electricity produced in OECD countries (17 per cent world-wide).

The mission of the OECD/NEA (Ref 1) is to assist its Member countries in maintaining and further developing, through international co-operation, the scientific, technological and legal bases required for a safe, environmentally friendly and economical use of nuclear energy for peaceful purposes, as well as to provide authoritative assessments and to forge common understandings on key issues, as input to government decisions on nuclear energy policy and to broader OECD policy analyses in areas such as energy and sustainable development. In carrying out this mission, the NEA facilitates exchange 'and comparison' of national regulatory practices, reviews selected technical and economic aspects of nuclear power development and sponsors joint international research projects. In addition, the NEA through its studies and the development of technical standards and codes of practice provides a source of objective information and advice to its Members. The broad direction of the programme of work of the OECD/NEA is set by the Steering Committee for Nuclear Energy. This Committee is assisted by a number of specialised standing committees such as the Committee on the Safety of Nuclear Installations (CSNI) and the Committee on Nuclear Regulatory Activities (CNRA). In a typical year, these two committees with the assistance of the NEA secretariat organise 12 to 15 workshops focused on well defined topics of current high interest to nuclear safety; produce about 35 to 40 reports - several of which are state-of-the art reports - and organise and co-ordinate a number of International Standard Problem (ISP) exercises with the aim of increasing confidence in the prediction of the computer codes used to assess the safety of nuclear power plants. In general, the emphasis of the OECD/NEA safety programme is related to safety research. The CSNI, (ref 2), is made up of senior scientists and engineers, with broad responsibilities for safety technology and research programmes. The technical fields of nuclear reactor safety interest into which the CSNI has designated specific Working Groups (items 1 to 4) and Special Expert Groups (items 5 and 6) are: 1. Risk Assessment, 2. Analysis and Management of Accidents,

3-4

3. Integrity of Components and Structures, 4. Operating Experiences, 5. Human and Organisational Factors 6. Fuel Safety Margins along with expert groups formed from time to time. This paper will focus specifically on those activities of Working Group on Integrity and Ageing (IAGE) relevant to thermal fatigue in LWR piping. IAGE WG deals with the integrity of structures and components, and has three sub-groups, dealing with the integrity of metal structures and components, ageing of concrete structures, and the seismic behaviour of structures. It should be noted by most activities are done in cooperation with other Working Groups, in particular the CSNI WG on Operating experiences which guides the operation of the Incident Reporting System (IRS) and also evaluates the events and identifies significant safety issues to be brought to the attention of CSNI. There is also co-operation with the former task group on thermalhydraulic applications. The IAGE working group on metal components has dealt with issues related more or less closely to thermal fatigue in piping by sponsoring: round robins or benchmarks on fatigue crack growth in plates and pipes under mechanical loading, on crack opening behaviour and leak rates in pipes, on pipe leak and break probabilities, and NDE in a range of components (the former PISC project, jointly with the CEC); Specialists meetings or workshops on Leak Before Break, reactor coolant system leakage and failure probabilities and experience with thermal fatigue in LWR piping caused by mixing and stratification; and a report on monitoring. NEA works in co-ordination with other international exercises such as the CEC, IAEA and WANO to avoid duplication and to ensure the appropriate participation. This coordination takes the form of liaison between the secretariats, joint sponsorship of meetings or projects, and exchange of future programmes of work. THERMAL FATIGUE IN LWR PIPING Thermal fatigue due to stratification and mixing of hot and cold water is a recurring phenomenon, but with a relatively low frequency. It can affect safety related piping of many different systems (ECCS, RHR, AFW, pressuriser surge line and safety / relief valve discharge lines) at various locations. It may happen that these phenomena are not included in the design conditions. Recently, similar events have occurred in sections of piping in the primary circuit and therefore the plant owners and the safety

3-5

authorities are now alerted over some conditions not being consistent with licensing basis and inspection commitments. Of particular concern are those problems arising from unanticipated thermal fatigue in unisolable piping connected to the reactor coolant system. None of the mentioned incidents led to radioactivity propagation to the environment, but the safety significance of these events results from the leakage of primary coolant through the second barrier but inside the reactor containment. Thermal fatigue problems can be seen as challenging issues for plant owners. Because of the high potential impact on safety, cost and radiation exposure, these issues have to be addressed more effectively. This is possible only with a very close co-operation between researcher, designer and plant owner and, among plant personnel, between maintenance and operation staff. Having all these people working together is the key point to keep the risk of thermal fatigue under control and, more generally, to ensure safe and effective plant operation. There is a need to encourage co-operation between the different disciplines to solve the problem. PAST ACTIVITIES OF THE GROUP The IAGE working group on the integrity of metal structures and components has an on-going activities in the area of thermal fatigue in LWR piping. In 1997 this IAGE WG decided a Specialist Meeting, (ref 3), on experience with thermal fatigue caused by mixing and stratification of hot and cold water would be timely. There have been many incidents (cracks) caused by stratification. The IRS had about 25 reports up to 1996, and during the year before there were several events with the same cause. Thermal fatigue cracks in BWR internals have been of concern for some time. Other events such as Dampierre-1 had raised questions about earlier assumptions. The Specialist Meeting was hosted by IPSN, France under the sponsorship of WANO and CSNI WG on Operating experiences and CSNI WG on Integrity and Ageing jointly. As the topic arises out of operating experience, it was essential to involve the utilities fully through the participation of WANO. There had been sessions on operating experience, thermal hydraulic phenomena, material and structural response, monitoring, inspection, mitigation and prevention and the safety implication. There had been about 70 participants. Outputs of the Specialist Meeting . Outputs of the Specialist Meeting were written down into conclusions and recommendations that were agreed by participants and approved by CSNI. Although some progress in the knowledge have been made since this meeting they are still valid and served as a basis to draw up the IAGE WG action plan. They are reproduced below. 3-6

"- There is a need to develop further accepted methods to identify locations with potential risk of thermal fatigue. There are proposals for simplified screening criteria based on semi-empirical models to determine the areas where there is risk of thermal fatigue, but these are not generally accepted. As there are many uncertainties, it is possible to consider a probabilistic treatment, although the data for this are also limited. As first step in this direction, the use of best estimate analysis methods should be considered, so that attention is focused on the areas most at risk for thermal fatigue, although these may be difficult to define. Then, at those locations, monitoring of temperature and of pressure if necessary should be implemented. In addition, periodic verification of the leaktightness of the nearest valves could further reduce the risk of thermal fatigue. - Monitoring of temperature fluctuations can be seen as an important part of the defence and, at present, it remains the most reliable method to avoid unanticipated incidents. At present, there are different strategies in use and no single method provides defence. There is a need for combining redesign and revised operating practices. A small internal cold water leakage into a hot section of pipe can lead to a quick propagation of cracks by thermal fatigue in some sensitive zones. Manufacturing process probably has a very important impact on the rate of crack initiation. It is not possible to draw up simple criteria for such parameters as allowable valve leak rates or limiting pipe diameters, as there is a great variety in the systems and operating procedures. - Concerning the experience from non-destructive examinations, ultrasonic testing gave numerous false calls, and for certain geometries and material conditions, performance of present NDE methods to detect fatigue cracks is limited. If a greater reliance is to be placed on NDE, the development of qualified methods is needed." Follow up actions approved by CSNI. Discussions at the Specialist Meeting held in Paris in June 1998, and within the IAGE working group have identified a need to develop screening criteria to guide monitoring and inspection. That is, there is a need to develop reliable methods to identify positions in the reactor coolant and associated systems, which are sensitive to thermal stratification and thermal mixing. It was proposed to work towards this objective, by preparing a report or technical position paper on present practices. Topics to be discussed are: Design for conservative estimates of thermal loads Design and operation practices that reduce thermal stratification/mixing problems
S S

3-7

Actions taken to mitigate thermal stratification/mixing problems Thermal stratification/mixing in conventional power plants Thermal stratification/mixing problems with Russian designed reactors

S S S S S S S S S S

Other topics that could be investigated subsequently include: Basic causes of fatigue cracking by thermal stratification and thermal mixing Effective inspection methods Development of improved and reliable methods for the detection of fatigue cracks especially in critical components like elbows Advanced methods for the detection of the early phases of fatigue cracking Consideration of benchmarks for comparing present design methods

Other action undertook by the CSNI was the co-sponsorship along with the US Nuclear Regulatory Commission of the "International Conference on Fatigue" host by EPRI in Napa, California on July 31 - August 2, 2000. CURRENT AND FUTURE ACTIONS. At its last meeting in June 2000 in Paris, the IAGE Working Group heard presentations from France, Germany and Japan showing this topic kept on being of high interest for both the industry and regulatory bodies. The group discussed thoroughly the significance of (1) screening criteria to determine critical locations, (2) management of interfaces between system engineers, thermal hydraulic and mechanical engineers, (3) code changes and (4) experimental information available after all those years of work. He then draw up a programme of work that would address the subject with an international perspective as it is the purpose of the OECD-CSNI. One aspect of the program is to get a broad picture of technical issues. Outputs of the "International Conference on Fatigue" and of the Specialist Meeting in Paris in 1998 would serve to update understanding of the issues by sharing views and problems. Some topics listed above in the follow-up actions paragraph would be then addressed. As to the other aspect, it will focus on actions that would complement and foster cooperation between countries. To start with, it was decided to conduct an international survey on: screening criteria, experiments available that address high cycle thermal fatigue,

3-8

the tendency for code changes, if any.

3 Proceedings of the Specialists Meeting on Experience with thermal fatigue in LWR piping caused by mixing and stratification, Paris 1998. OECD/NEA/CSNI/R(98)8, December 1998.

Future activities considered by the group after the above mentioned actions being completed would be to conduct international benchmarks on technical issues (e.g. thermal hydraulic, crack initiation and propagation, NDE techniques and capabilities). CONCLUSIONS The contents of this paper provide an overview of both the framework and structure of the OECD Nuclear Energy Agency and in particular the activities being performed in the area of thermal fatigue in LWR piping. On issues of thermal fatigue, the main NEA safety technology committee, the Committee on the Safety Nuclear Installations, organises relevant activities based on the aspects of operating experience, structural integrity and thermalhydraulics. In all these areas, groups of experts review at regular intervals the state of the art and organise studies, workshops, preparation of reports, etc., aimed at discussing issues of mutual concern and at arriving at common technical positions. The use of specialist groups and specialist meetings to examine in depth specific questions of nuclear safety and the use of international standard problems to provide a vehicle for discussion and comparison are the important characteristics of the OECD/NEA programme of work. Although the OECD/NEA obligations are primarily towards its Member countries, most of the information generated by its activities is made available to all countries. Furthermore, in line with the general policy of the OECD, there is a trend towards increasing participation by countries which are not currently Members of the OECD. In the area covered by this paper, the OECD-CSNI decided to promote a program of work by organising or sponsoring conferences and workshops and also by promoting exchange of information on critical topics by conducting a survey in 2000. Follow-up actions would be scheduled after completion of those tasks. References

1 The Strategic Plan of the Nuclear Energy Agency , OECD/NEA/NE(99)1, 1999 2 The Strategic Plan for the Committee on the Safety of Nuclear Installations . OECD/NEA/CSNI/R(2000)3,
January 2000.

3-9

International Conference on Fatigue NAPA, CA, USA July 31 - Aug 2, 2000 Organized by EPRI Co-sponsored by USNRC and OECD Nuclear Energy Agency/Committee on the Safety of Nuclear Installations (CSNI)
International Conference on Fatigue, NAPA July31 - Aug 2, 2000 Eric Mathet, IAGE Secretary July 31, 2000

3-10

OECD
s s s

Organization for Economic Co-operation and Development W. Europe, N. America, Japan 29 members - in 1996 the Czech Republic, Hungary, Poland and S. Korea joined x OECD is
a tool for intergovernmental co-operation mainly in the economic field a place for policy makers to compare point of views and experiences

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-11

NEA (Nuclear Energy Agency)


s s s

Division of OECD 27 Member Countries 80 staff

Mission is to assist its Member countries


x

in maintaining and further developing, through international co-operation, a peaceful and economical use of nuclear energy to provide authoritative assessments and to forge common understandings on key issues

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-12

NEA (Nuclear Energy Agency)/CSNI


s

Nuclear Safety Division provides secretariat for CSNI

CSNI works through four WGs and two Special Expert Groups (SEGs)

WG on Structural Integrity of Components and Structures

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-13

NEA Strenghts
s

Homogeneity in membership
x x x x

small club like-minded approach climate of mutual trust relatively non political

s s

Provides added value and is cost effective Strong scientific/technical/legal work


x x

narrow focus does not deal with proliferation, safeguards

Work methods flexible and responsive to member needs


Eric Mathet, IAGE Secretary July 31, 2000

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

3-14

Four Working Groups and two Special Expert Groups


s s s s s s
s

Risk assessment Analysis and management of Accidents Integrity of Components and Structures Operating experiences Human and organisational factors Fuel Safety Margins
Co-ordination with CEC, IAEA, WANO etc
Eric Mathet, IAGE Secretary July 31, 2000

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

3-15

COMMITTEE on the SAFETY OF NUCLEAR INSTALLATIONS (CSNI) Chairman: M. Livolant - Secretary: G.M. Frescura CSNI Bureau: A. Thadani, A. Alonso

Working Groups

CSNI Programme Group

Special Expert Groups

Risk Assessment

Analysis and Management of Accidents

Integrity of Components and Structures

Operating Experiences

Human and Organisational Factors

Fuel Safety Fuel Behaviour Margins

Chairman: Secretary: Secretary:

Chairman: Chairman: Secretary: Secretary:

Chairman: Secretary: Secretary:

Chairman: Chairman: Secretary: Secretary:

Chairman: Chairman: Secretary: Secretary:

Chairman: Secretary:

Software Reliability Fire Risk Assessment

Design Basis

Integrity of Metal Components & Structures Concrete Structures Ageing

Fuel Cycle Safety

Human Reliability Errors of Commission

Severe Accidents

Risk Monitor

Fission Product Behaviour

Seismic Behaviour of Structures

Passive Systems PSA

DATA Bases Incident Reporting System (IRS) ICDE Project -Common Cause Failure Data Computer-Based Safety Systems Fuel Incident & Reporting System (FINAS)

NEA Co-operation with CEEC and NIS

RASPLAV

HALDEN REACTOR PROJECT

SANDIA LOWER HEAD FAILURE PROJECT

CABRI WATER LOOP PROJECT

New CSNI structure

3-16

WG on Integrity of Components and Structures (IAGE)


Since 1996 has 3 sub-groups
x

Integrity of metal components and structures

x Aging

of concrete structures behavior of structures

x Seismic

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-17

Typical IAGE WG activities


Workshops, Specialists Meetings s Benchmarks/round robins/ISPs s State of the Art Reports, Topical reports
s

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-18

Activities on Thermal Fatigue

Conference on fatigue, IPSN,Paris, 1998 s CSNI Approved proposal (Dec. 99) s Co-sponsorship of the International Conference on Fatigue, EPRI, July-Aug 2000 s IAGE WG program of work (June 2000)
s
International Conference on Fatigue, NAPA July31 - Aug 2, 2000 Eric Mathet, IAGE Secretary July 31, 2000

3-19

OUTPUTS of the Conference on fatigue, Paris, 1998


x

need to develop further accepted methods to identify locations with potential risk of thermal fatigue; Monitoring of temperature fluctuations can be seen as an important part of the defence and, at present, it remains the most reliable method to avoid unanticipated incidents; If a greater reliance is to be placed on NDE, the development of qualified methods is needed.

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-20

CSNI Approved Proposal, Dec 1999

Based on: a need to develop screening criteria to guide monitoring and inspection.
I.e. a need to develop reliable methods to identify locations in the reactor coolant and associated systems, which are sensitive to thermal stratification and thermal mixing.
International Conference on Fatigue, NAPA July31 - Aug 2, 2000 Eric Mathet, IAGE Secretary July 31, 2000

3-21

CSNI Approved Proposal, Dec 1999


s

Topics to be discussed are


! !

Design for conservative estimates of thermal loads; Design and operation practices that reduce thermal stratification/mixing problems; Actions taken to mitigate thermal stratification/mixing problems; Thermal stratification/mixing in conventional power plants; Thermal stratification/mixing problems with Russian designed reactors.

! !

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-22

CSNI Approved Proposal, Dec 1999


s

Other topics that could be investigated subsequently include:


"

Basic causes of fatigue cracking by thermal stratification and thermal mixing; Effective inspection methods; Development of improved and reliable methods for the detection of fatigue cracks especially in critical components like elbows; Advanced methods for the detection of the early phases of fatigue cracking; Consideration of benchmarks for comparing present design methods.

" "

"

"

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-23

IAGE WG discussions
At its last meeting in June 2000, the Metal sub-group discussed thoroughly the significance of
x x

screening criteria to determine critical locations; management of interfaces between system engineers, thermal hydraulic and mechanical engineers; tendency for design code changes; experimental information available after all those years of work.

x x

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-24

Follow-up actions
Conduct an international survey on:
" "

screening criteria; experiments available that address high cycle thermal fatigue; the tendency for code changes, if any.

"

Conduct international benchmarks on technical


issues (e.g. thermal hydraulic, crack initiation and propagation, NDE techniques and capabilities).

International Conference on Fatigue, NAPA July31 - Aug 2, 2000

Eric Mathet, IAGE Secretary

July 31, 2000

3-25

4
FATIGUE OF REACTOR COMPONENTS: NRC ACTIVITIES
D. Kalinousky J. Muscara U.S. Nuclear Regulatory Commission Washington, D.C., 20555

4-1

Fatigue of Reactor Components: NRC Activities

D. Kalinousky J. Muscara U.S. Nuclear Regulatory Commission Washington, D.C., 20555

4-3

Fatigue of Reactor Components: NRC Activities

D. Kalinousky J. Muscara U.S. Nuclear Regulatory Commission Washington, D.C., 20555

Abstract The NRC has been dealing with the fatigue of reactor components since the late 1970s. The NRC has initiated a Fatigue Action Plan and has identified three Generic Safety Issues dealing with fatigue; GSI-78, Monitoring of Design Basis Transient Fatigue Limits for Reactor Coolant System, GSI-166, Adequacy of Fatigue Life of Metal Components, and GSI-190, Fatigue Evaluation of Metal Components for 60-Year Life. It has also supported research on fatigue at the Argonne National Laboratory. It has been found in the above studies that the risk of core damage from fatigue failures of primary system components is small and does not justify any additional actions by the licensees. Although, for a twenty-year license renewal period, an increase in the frequency of fatigue failures has been identified. This increase in failure frequency needs to be addressed during license renewal through an aging management program. The ASME design curves have also been found to be non-conservative under certain reactor coolant environments and loading conditions. This is an issue that the NRC staff has discussed with ASME Code committees over the last 12 years. The NRC has asked the ASME Code to take actions to address this issue and has recommended two acceptable methods for incorporating the effects of the environment in code fatigue analyses. The first method is to develop the design curve in the conventional manner using the data from specimens tested in reactor environments, and the second method is to use a fatigue life correction factor.

4-4

Introduction Concerns about fatigue in reactor components arose in the late 1970s with reports of fatigue crack initiation and growth occurring in operating reactor components. These occurrences were discussed in the NRCs IE Bulletin 79-13 (1). A further concern was the possibility of thermal or mechanical transients that were not included in the original fatigue design basis. NRC Bulletin 88-08 (2) and NRC Bulletin 88-11 (3) were issued to initially address this concern. To further assess these concerns, the NRC initiated several actions. It identified several Generic Safety Issues dealing with fatigue and developed a Fatigue Action Plan to resolve these issues; in addition, research was initiated on the various fatigue topics. This research has been used to resolve these generic issues and to identify issues for further investigation. Currently, there is ongoing fatigue research sponsored by the NRC under its Environmentally Assisted Cracking program. Fatigue Action Plan In 1993, the Commission instructed the NRC staff to treat fatigue as a potential safety issue for operating reactors. The staffs response was the development of the Fatigue Action Plan (FAP). The FAP was developed to resolve three principal issues: Many older vintage nuclear power plants have components of the reactor coolant pressure boundary (RCPB) that were designed to industry codes that did not require the explicit fatigue analysis presently required by the American Society of Mechanical Engineers (ASME) Boiler and Pressure Vessel Code. A concern was raised regarding the fatigue resistance of these components for the plant design life. Test data showed that the design fatigue curves of the ASME Code were not conservative for nuclear power plant primary system environments (4). A concern was also raised regarding the fatigue resistance of components designed using these ASME Code curves. The appropriate corrective action to be taken when the fatigue allowable limit is exceeded (calculated Cumulative Usage Factor (CUF) > 1) has been a subject of controversy. The staff identified a need to develop a staff position on this subject.

One of the tasks under the FAP was to conduct a survey to determine the number of operating plants that have a fatigue analysis of the vessel, primary system components, and piping. This was done by a review of the available NRC licensing documentation. The survey found that approximately 40 percent of the operating plants have primary system components designed to a piping code that did not require a formal fatigue analysis. 4-5

To address the technical issues, the FAP called for the fatigue evaluation of a sample of RCPB components. The Idaho National Engineering Laboratory (INEL), under contract to the NRC, conducted assessments of sample components selected from seven power plants. The components were selected from an EPRI report (5) listing typical areas in PWRs and BWRs for which high fatigue usage factors have been calculated. In this analysis, interim fatigue design curves developed at Argonne National Laboratory (ANL) were used (4). The NRC published the results of the INEL component evaluations in NUREG/CR-6260 (6). The evaluation of the sample of components with high design usage factors, using the interim design fatigue curves, indicated that the ASME fatigue criteria (CUF<1) could be met for the majority of components when conservatisms in the original component design were taken into account. A detailed finite element analysis on two of the components indicated that additional conservatisms could be removed. In addition, data from transient monitoring instrumentation indicated that design transients for some components were conservative. Even considering detailed finite element analysis and plant transient monitoring, there are still a limited number of piping locations where the ASME Code fatigue criteria may be exceeded at the end of design life. The staff took the position that when the CUF for a component exceeded 1, Generic Letter (GL) 91-18, Information to Licensees Regarding Two NRC Inspection Manual Sections on Resolution of Degraded and Nonconforming Conditions and on Operability, (7) would apply. The GL stipulates that the failure to meet a code criterion specified in the FSAR is a nonconforming condition, which requires a corrective action. Since the license basis code is specified in the FSAR, the guidance in GL 91-18 is applicable. On the basis of the accomplished FAP tasks, the staff found that no immediate staff or licensee action was necessary to deal with the fatigue issues addressed. The FAP was completed in September of 1995 (8). Generic Safety Issue - 78 Generic Safety Issue (GSI) 78, Monitoring of Design Basis Transient Fatigue Limits for Reactor Coolant System, was developed in 1992 to determine whether transient monitoring was necessary at operating plants. The goal of the transient monitoring addressed by GSI-78 was to provide assurance that components do not exceed their licensing basis during the lifetime of the plant. The GSI-78 transient monitoring concern was addressed as part of the FAP. The resolution of GSI-78 was accomplished by focusing on the evaluation of risk from fatigue failure of selected reactor coolant system components over a 40-year period. A GSI-78 resolution package was presented to the Advisory Committee on Reactor 4-6

Safeguards with a discussion of the results of a study using the PRAISE code. This study demonstrated that the risk from fatigue failure of the reactor coolant system components was very low. Thus, GSI-78 was closed in February of 1997. Generic Safety Issue 166 GSI-166, Adequacy of Fatigue Life of Metal Components, was established in response to questions raised by the staff in SECY-93-049, Implementation of 10 CFR Part 54, Requirements for Renewal of Operating Licenses for Nuclear Power Plants, (9) about reassessing the fatigue design of metal components. The output from the FAP and GSI78 provided essential technical information for the resolution of GSI-166. In resolving GSI-166, the staff obtained the records of transient monitoring from the seven plants selected in the FAP assessments. Based on these records and conservatisms identified in the component analyses, the staff did not believe there existed a significant safety concern that current licensing basis fatigue criteria had been exceeded at operating plants over a 40-year life. Thus, no action was required for the licensees as a result of the evaluations performed under GSI-166 and the GSI was closed in February of 1997. Generic Safety Issue 190 GSI-166 and GSI-78 only applied to the original 40-year plant life. To address the possibility of a 20-year license renewal period, the staff opened GSI-190, Fatigue Evaluation of Metal Components for 60-Year Life, analyzing the same components from the reactor coolant system as for GSI-166, but for a 60 year life in the LWR environment. GSI-190 addressed the environmental effects on design basis fatigue transients, studying the probability of fatigue failure, and the associated core damage frequency (CDF) for a 60-year plant life. It did not address all aspects of fatigue related degradation, including those outside the design basis. Events involving thermal or mechanical stresses that result in unanticipated cyclic loads are outside the scope of the design basis transients and were not studied under GSI-190. The approach to resolving GSI-190 built upon the approaches used in the FAP and the previous GSIs. The Pacific Northwest National Laboratory (PNNL), under contract to the NRC, made use of the most recently developed fatigue curves in simulated LWR environmental conditions from Argonne National Laboratory (ANL) (10, 11, 12). PNNL performed calculations of the probability of component failure and the CDF associated with these failures. These results are reported in NUREG/CR-6674, Fatigue Analysis of Components for 60-Year Plant Life (13).

4-7

Probabilistic fatigue calculations using the pc-PRAISE code were performed on 47 sample components from 6 locations in seven example plants (five PWR and two BWR plants). During this work the staff identified that the pc-PRAISE code could not model large aspect ratio cracks (ratio of crack length to crack depth), nor could it model the joining of several small cracks to make a large crack which could subsequently propagate through the wall thickness of the component. The staff concluded that the effect of large aspect ratios was an important factor in fatigue analyses and there was a need to modify the pc-PRAISE code to model this. The modified pc-PRAISE program (version 4.2) is available via the NRCs ADAMS system, in a package titled pcPRAISE. Using the modified pc-PRAISE program, PNNL performed a probabilistic analysis for crack initiation and through wall crack growth in the components mentioned above for 40 and 60-year plant lives, considering both air and LWR (water) environments. Calculations for the air environment and the 40-year life were done to provide a baseline from which to compare the results of a 60-year life, incorporating the effect of the revised fatigue curves coupled with the modified pc-PRAISE program. An evaluation was performed to estimate the conditional CDF from the fatigue failure of these components. Given a through wall crack, the objective was to estimate the conditional probabilities of a small leak, of a large leak, and of a pipe break. Data on pipe failure events indicate that only a small fraction of through-wall flaws result in large leaks or breaks. Probabilistic fracture mechanics models, like the one contained in the pc-PRAISE program, predict that fatigue failures will usually be in the mode of small leaks rather than large leaks or breaks. From the conditional probabilities of small leak, large leak, and pipe break, the conditional CDF was estimated for the seven example plants based on results extracted from probabilistic risk assessments (PRA). The major findings from the calculations include the following points: Many of the components have cumulative probabilities of crack initiation and cumulative probabilities of through-wall cracks that approach unity within the 40 to 60-year time period. However, some components, often with similar values of fatigue usage factors, show much lower failure probabilities. This is related to the fact that the usage factors address only crack initiation and do not address the specific factors for each component that determine how likely it is that an initiated crack will grow through wall. The maximum failure rate is in the range of 10 -2 through-wall cracks per reactor year, and those failures were associated with high CUF locations. The 10 highest component failure rates for 40-years in air, 40-years in water, and 60-years in water are shown in Figure 1. Failure rates for other components having much lower failure probabilities are changed by as much as an order of magnitude from 40 to 60-years, but these components make relatively small overall contributions to the cumulative CDF estimates. 4-8

The maximum CDF based on these calculated failure rates is about 10 -6 per year. These maximum values correspond to components with very high cumulative failure probabilities, and the failure rates do not change much from 40 to 60 years. The range of CDF was between 10 -6 to 10-15. The 10 highest CDF rates for 40-years in air, 40-years in water, and 60-years in water are shown in Figure 2.

Based upon these low CDFs, the staff concluded that they could not be used as a basis for a cost/benefit backfit analysis to justify imposition of a new regulatory requirement on operating reactors. However, the calculations supporting resolution of this issue, which included consideration of environmental effects, indicates the potential for an increase in the frequency of pipe leaks as plants continue to operate. Thus, the staff concluded that, consistent with existing requirements in 10 CFR 54.21, licensees should address the effects of the coolant environment on component fatigue life as aging management programs are formulated in support of license renewal. NRC Research ANL has been conducting a research program for the NRC on environmentally assisted cracking since 1978. As part of this program, the environmental effects on fatigue crack initiation in pressure vessel and piping steels has been studied since 1986. Several key variables have been identified that influence the steels fatigue life; such as: steel type, sulfur content, strain amplitude, strain rate, temperature, and dissolved oxygen content. Test data from ANL and Japan have shown potentially significant effects of LWR environments on the fatigue resistance of pressure vessel and piping steels. Based on the existing fatigue strain vs. life (S-N) data, interim fatigue design curves have been developed that account for environmental effects and the results were reported in NUREG/CR-5999. A more rigorous statistical analysis of the available data was performed and statistical models were developed in NUREG/CR-6335 for estimating the fatigue lives of carbon and low-alloy ferritic steels, austenitic stainless steels, and Alloy 600 in LWR environments. The results have been used to estimate the probability of fatigue cracking. As more experimental data became available, the statistical models have been modified and/or optimized with a larger fatigue database. The ANL work performed more recently on fatigue in LWR environments has been summarized in topical reports NUREG/CR-6583 (14) for carbon and low-alloy steels, and NUREG/CR-5704 (15) for wrought and cast austenitic stainless steels. The reports present an overview of the existing fatigue S-N data on these steels and describe the influence of key service and material parameters on fatigue life, the mechanism of crack initiation in LWR environments, and procedures for incorporating environmental effects in ASME Section III fatigue evaluations. Two possible procedures have been identified for incorporating the effects of LWR environments into fatigue evaluations. The first procedure calls for environmentally adjusting the ASME design curve, such that it takes into account the shorter fatigue life 4-9

of steels in the LWR environment. The design curve is developed in the conventional manner by applying the factor of 2 on stress or 20 on cycles to the S-N test data developed from specimens tested in LWR environments. The second procedure, originally developed in Japan by Higuchi and Iida (16), and later proposed in the U. S. by GE and EPRI (17) uses a fatigue life correction factor, F en, which is the ratio of the fatigue life in air at room temperature to that in water at service temperature. Either procedure provides acceptable results and the NRC staff can support their use in the ASME Code. Further research to be sponsored by the NRC at ANL over the next few years will provide information on such topics as the synergistic effects of surface finish and environment on fatigue life; the combined effects of loading sequence and environment; and fatigue crack initiation in sensitized stainless steel. A comprehensive evaluation of stainless steel fatigue test specimens will be performed to explain why environmental effects are more pronounced in low dissolved oxygen than high dissolved oxygen water. Several of these planned research tasks are: Fatigue tests on low-alloy steel specimens with differing surface finishes and treatments will be carried out in air and water environments. This will provide information on the effects of surface roughness on the fatigue life. Metallographic evaluations of stainless steel specimens will be conducted to help explain why environmental effects are more pronounced in low dissolved oxygen water than in high dissolved oxygen water. The metallographic evaluations will also help to understand the mechanism of fatigue crack initiation in a LWR environment. Fatigue tests will be carried out on sensitized Type 304 stainless steel. The effect of intermediate temperatures on the fatigue life of austenitic stainless steels will be investigated. Existing fatigue data does not include the combined effects of environment and sequential loading. Therefore, fatigue tests on low-alloy steels in high dissolved oxygen water will be conducted to investigate these effects on fatigue life.

Need to Address Environmental Effects and NRC Interaction with ASME Results from studies in Japan and those at ANL illustrate potentially significant effects of LWR coolant environments on the fatigue life of carbon and lowalloy steels and of austenitic stainless steels (Figures 3 and 4). Under certain loading and environmental conditions the reactor water environmental effects alone substantially exceed the reductions in the current design curve to account for the differences between specimen

4-10

tests and component behavior. In addition, the current Code design curve for stainless steel in air does not accurately represent the available experimental data and thus includes a reduction of only 1.5 and ~10 from the mean air curve, not the 2 and 20 originally intended (Figure 5). Based on these studies and findings, NRC and ANL staff have been recommending to ASME Code committees and to the Board on Nuclear Codes and Standards (BNCS) since the late 1980s that the ASME Code fatigue design curves need to be updated to incorporate the effects of reactor coolant environments. The need to address this issue continues to become more important in light of the aging of plants, changes in other parts of the fatigue analyses addressed by the Code, and the renewal of plant licenses for an additional 20 years of operation. Ample data now exist to support the required changes. Although the existing fatigue design curves are non-conservative for certain LWR conditions and environments, the current design procedures, e.g., stress analysis rules, cycle counting, etc., are conservative enough that the overall assessment of fatigue life has been conservative. However, the Code permits new improved approaches for fatigue evaluations, e.g., finite element analyses, fatigue monitoring, improved Ke factors, etc., which can significantly reduce the conservatism in the other elements of the present design methods. To ensure that the overall assessment maintains a degree of conservatism consistent with that chosen for the fatigue limit in air, the current understanding regarding environmental effects should be incorporated into the ASME Code for fatigue evaluations. The ASME Code fatigue design curves, given in Appendix I of Section III, are based on straincontrolled tests of small polished specimens at room temperature in air. The fatigue design curves were developed from the mean curves for the experimental data by reducing the fatigue life at each point on the mean curve by a factor of 2 on strain or 20 on cycles, whichever was more conservative. This reduction was intended to account for data scatter (heattoheat variability), effects of mean stress or loading history, and the differences in surface condition and size between the test specimens and actual components. As explicitly noted in Subsection NB3121 of Section III of the Code, the effects of the LWR coolant environment are not addressed in the current Code fatigue design curves. The Subsection states that the owner's design specifications should provide information on any reduction to fatigue design curves necessitated by environmental conditions. In 1991, the ASME BNCS requested the PVRC to examine the existing worldwide fatigue strain vs. life (SN) data and develop recommendations for ASME. The PVRC committee on cyclic life and environmental effects has evaluated the issue and, at its June 15, 1999 meeting in Columbus, Ohio (18), the committee endorsed a method for incorporating the effects of LWR coolant environments into the ASME Code fatigue evaluations. The Executive Director of PVRC has transmitted their recommendations and approach to implement environmental fatigue procedures to the ASME BNCS by

4-11

letter from Hollinger to Ferguson dated October 31, 1999. The methodology, proposed by EPRI/GE (17), is based on the use of fatigue life correction factors (F en) to account for environmental effects on fatigue life. The PVRC recommendation also defines an effective F en, Fen/Z, where Z is a factor that constitutes the perceived conservatism in the Code design curves, i.e., the portion of that factor of 20 that is judged not actually necessary to account for data scatter, surface finish, etc. PVRC proposed the factor Z as 3.0 for carbon and lowalloy steels and 1.5 for stainless steels. Recent reports from Japan (19) and also findings from the ANL work (20) indicate that the entire margin of 20 is expended by factors other than the environmental effects. Also, the effective F en approach presumes that all other uncertainties have been anticipated and accounted for. Therefore we do not believe it is prudent to apply the factor Z to F en. The NRC sponsored work at ANL has also studied methods for incorporating environmental effects into ASME Code Section III fatigue evaluations (NUREG/CR6583, NUREG/CR-5704). Two procedures have been proposed; (a) use of environmentally adjusted fatigue design curves or (b) use F en to adjust the current ASME Code fatigue usage values for environmental effects. Although estimates of fatigue lives based on the two methods may differ somewhat because of differences between the ASME mean curves used to develop the current design curves and the bestfit curves to the existing data used to develop the environmentally adjusted curves, the NRC can support either of these methods as they provide an acceptable approach to account for environmental effects. In December 1999, the NRC again requested the ASME Code to address the effects of the LWR environment on fatigue life, and to update the current design curve for stainless steel in air. NRC staff and contractors are participating in Section III and XI activities to address and incorporate the effects of environment into ASME Code fatigue evaluations. The Section III Subgroup on Fatigue Strength has begun to revise the fatigue design curves. ASME Section XI Task Group on Operating Plant Fatigue Assessments is considering the possible inclusion of a methodology for addressing the effects of the environment on fatigue life in its revision of Non-Mandatory Appendix L. Thermal Fatigue In 1997, an event occurred at Oconee 2, which involved the unit being shut down due to cracking and leakage from a weld location in the 2-inch diameter ( NPS 2 ), Class 1 portion of a combination makeup and high-pressure injection line (equivalent to a portion of the High Pressure Safety Injection (HPSI) system as designated in the ASME Code).

4-12

After the failed nozzle assembly was removed from service and a metallurgical examination performed, it was determined that the crack consisted of a 360 degree inside surface flaw with minimum depth of 30 percent through-wall. The crack had also penetrated completely through-wall over an arc length of 77 degrees. The licensee for Oconee attributed the cracking to thermal cycling and flow-induced vibration. As a result of this event, the inservice inspection (ISI) requirements given in Section XI of the ASME Code for butt-welded connections in HPSI piping were reexamined. Subsection IWC of Section XI of the ASME Code requires that Class 2 HPSI piping down to NPS 2 receive both a volumetric and a surface examination as part of a facilitys ISI program. Vessels, piping, pumps, valves and other components NPS 1 and smaller in HPSI systems of pressurized water reactor plants are, however, specifically exempted from this requirement. Subsection IWB of Section XI of the ASME Code requires only surface examination for Class 1 butt-welded piping connections less than NPS 4, with the one exemption provision excluding piping of NPS 1 and smaller from examination. Therefore, for the HPSI system, the inspection criteria for Class 2 piping between NPS 4 and NPS 2 (or NPS 1 ), inclusive, are more comprehensive than those for Class 1 piping of the same size range. Moreover, the examination requirements specified for Class 1 HPSI piping may not be adequate to detect thermal fatigue cracking initiating on the inside diameter in a timely manner since volumetric examinations are not required. In response to issues raised by the NRC on the inspection criteria and the thermal fatigue event at Oconee 2, the industry proposed an initiative to address the issue of thermal fatigue in nuclear power plant piping. The program was initiated under the Electric Power Research Institute (EPRI) Pressurized Water Reactor (PWR) Materials Reliability Project (MRP). The goal of the Thermal Fatigue Issue Task Group (established/approved in March 1999 by the MRP) is to provide a consistent set of guidelines and a methodology for addressing thermal fatigue issues. The NRC and the EPRI MRP periodically meet to discuss progress and issues related to thermal fatigue. It is expected that the task group will develop screening, monitoring, inspection, and management guidelines addressing thermal fatigue in the 2000-2001 timeframe. Because thermal fatigue events that are not part of the design basis were not evaluated under the generic safety issues on fatigue, and because thermal fatigue cracking events have recently been reported in other plants, including France and Japan, the NRC will follow closely the work, guidelines and implementations being developed under this industry initiative. Conclusion Analysis of the fatigue issues through the resolved GSIs and the completed FAP have resulted in no new requirements being imposed on the licensees for their current 4-13

licensing period. For license renewal, it has been found that the increase in through wall crack frequency with an additional 20-year period of operation makes it necessary for the licensees to consider fatigue in their aging management programs. Research sponsored by the NRC has provided results that have helped resolve generic issues related to fatigue. It is providing data, results and methodologies that could be used to update the ASME Code for incorporating the effects of the LWR environment in fatigue analyses or for the NRC to develop its own guidance. References 1. U. S. Nuclear Regulatory Commission, IE Bulletin 79-13, Cracking in Feedwater System Piping, October 16, 1979. 2. U. S. Nuclear Regulatory Commission, NRC Bulletin 88-08, Thermal Stress in Piping Connected to Reactor Coolant Systems, August 4, 1988. 3. U. S. Nuclear Regulatory Commission, NRC Bulletin 88-11, Pressurizer Surge Line Stratification, December 20, 1988. 4. S. Majumdar, O.K. Chopra, and W.J. Shack, Interim Fatigue Design Curves for Carbon, Low-Alloy, and Austenitic Stainless Steels in LWR Environments, NUREG/CR-5999, 1993. 5. Electric Power Research Institute, Report TR-100252, Metal Fatigue in Operating Nuclear Power Plants, April 1992. 6. A.G. Ware, D.K. Morton, and M.E. Nitzel, Application of NUREG/CR-5999 Interim Fatigue Curves to Selected Nuclear Power Plant Components, NUREG/CR-6260, 1995. 7. U. S. Nuclear Regulatory Commission, GL 91-18, Information to Licensees Regarding Two NRC Inspection Manual Sections on Resolution of Degraded and Nonconforming Conditions and on Operability, November 7, 1991. 8. SECY-95-245, Completion of the Fatigue Action Plan, September 25, 1995.

9. SECY-93-049, Implementation of 10 CFR Part 54,Requirements for Renewal of Operating Licenses for Nuclear Power Plants, March 1, 1993. 10. J. Keisler, O.K. Chopra, and W.J. Shack, Statistical Analysis of Fatigue Strain-Life Data for Carbon and Low-Alloy Steels, NUREG/CR-6237, 1994.

4-14

11. J. Keisler, O.K. Chopra, and W.J. Shack, Fatigue Strain-Life Behavior of Carbon and Low-Alloy Steels, Austenitic Stainless Steels, and Alloy 600 in LWR Environments, NUREG/CR-6335, 1995. 12. O.K. Chopra and F.A. Simonen, Private Communication, Updated Fatigue Design Curves for Austenitic Stainless Steel in LWR Environments, 1998. 13. M.A. Khaleel, F.A. Simonen, et al., Fatigue Analysis of Components for 60-Year Plant Life, NUREG/CR-6674, 2000. 14. O.K. Chopra and W.J. Shack, Effects of LWR Coolant Environments on Fatigue Design Curves of Carbon and Low-Alloy Steels, NUREG/CR-6583, 1998. 15. O.K. Chopra, Effects of LWR Coolant Environments on Fatigue Design Curves of Austenitic Stainless Steels, NUREG/CR-5704, 1999. 16. M. Higuchi and K. Iida, Fatigue Strength Correction Factors for Carbon and LowAlloy Steels in Oxygen-Containing High-Temperature Water, Nucl. Eng. Des., 129, 1991. 17. Electric Power Research Institute, Report TR-105759, An Environmental Factor Approach to Account for Reactor Water Effects in Light Water Reactor Pressure Vessel and Piping Fatigue Evaluations, December 1995. 18. Welding Research Council Program Report, Vol. LIX No. 5/6, May/June 1999.

19. M. Higuchi, Fatigue Curves and Fatigue Design Criteria for Carbon and Low-Alloy Steels in High-Temperature Water, Probabilistic and Environmental Aspects of Fracture Mechanics, PVP Vol. 386, 1999. 20. O.K. Chopra and J. Muscara, Effects of Light Water Reactor Environments on Fatigue Crack Initiation in Piping and Pressure Vessel Steels, ICONE-8, Baltimore, MD, April 2000, Paper No. 8300.

4-15

Through Wall Crack Frequency


1E+00 1E-01 1E-02 1E-03 1E-04 1E-05 1E-06 1E-07
1 2 3 4 5 6 7 8 9 10

Component Air: 40 Yr Water: 40 Yr Water: 60 Yr

Figure 1: Through Wall Crack Frequencies

1. GE Old Recirc RHR return line T 2. Westinghouse New - RHR line inlet transition 3. CE New - Surge line elbow 4. Westinghouse New - Charging nozzle nozzle 5. CE Old - Surge line elbow 6. Westinghouse New - RPV outlet nozzle 7. B & W - RPV outlet nozzle 8. CE Old - RPV outlet nozzle 9. GE New RHR line straight pipe 10. GE Old Core spray system safe end

4-16

Core Damage Frequency


1E-05 1E-06 1E-07 1E-08 1E-09 1E-10 1E-11 1E-12
1 2 3 4 5 6 7 8 9 10

Component Air: 40 Yr Water: 40 Yr Water: 60 Yr

Figure 2: Core Damage Frequencies

1. Westinghouse New - RHR line inlet transition 2. CE New - Surge line elbow 3. CE Old - Surge line elbow 4. Westinghouse New - Charging nozzle nozzle 5. Westinghouse New - RPV outlet nozzle 6. B & W - Makeup/HPI nozzle safe end 7. B & W - RPV outlet nozzle 8. CE Old - RPV outlet nozzle 9. B & W - Decay heat removal line reducing T 10. Westinghouse Old - RPV outlet nozzle inside surface

4-17

Figure 3: S-N Data for Carbon Steel in Water

Figure 4: S-N Data for Stainless Steel in Water

4-18

Figure 5: Current ASME Mean Curve

4-19

Fatigue of Reactor Components: NRC Activities


Joe Muscara Doug Kalinousky
U. S. Nuclear Regulatory Commission Office of Nuclear Regulatory Research Washington, DC 20555

INTERNATIONAL CONFERENCE ON FATIGUE OF REACTOR COMPONENTS 31 July 2 August, 2000 Napa, California
4-20

Introduction
GSI-78, GSI-166, & Fatigue Action Plan GSI-190 Research program accomplishments Need to address environmental effects NRC interaction with ASME Future activities
4-21

GSI-78
Monitoring of Design Basis Transient Fatigue Limits for Reactor Coolant System
Originated in May 1983 Determine if transient monitoring was necessary at operating plants Resolved as part of FAP by analyzing risk Risk of fatigue failure over 40 years is low

No regulatory actions necessary


4-22

GSI-166
Adequacy of Fatigue Life of Metal Components
June 1993 Resolved by identifying conservatisms in component analysis Based on a risk analysis, no significant safety concern that licensing criteria had been exceeded for 40 years

Resolved with no licensee actions required


4-23

Fatigue Action Plan


Originated: July 1994; Completed: Sept 1995 Used to resolve GSI-78 & GSI-166 Three principal issues
Some components did not require a fatigue analysis ASME fatigue design curves not conservative in a LWR environment What to do when CUF>1
4-24

GSI-190
Fatigue Evaluation of Metal Components for 60-Year Life Carried on GSI-78 & GSI-166 for a 20-year license renewal period Considered environmental effects Estimated CDF due to fatigue failures from design basis transients
4-25

GSI-190 (cont)
Modified pcPRAISE (available in ADAMS)
Initiation: multiple sites, varying times Growth: from initiation until failure Linking: can form large aspect ratio cracks

Maximum failure rate, 10-2 per year Maximum CDF, 10-6 per year Address environment on component fatigue life in aging management programs
4-26

Research Program Accomplishments


NRC initiated studies at ANL in 1986 on the effects of environment and loading on the fatigue life of LWR materials Obtained fatigue life data for different materials and test conditions
material: carbon steels; A106B, A333-B low-alloy steels; A533-B, A302-B SSs; types 304 & 316NG cast SS; CF8M (with & w/o thermal aging) air & water 288 C and room temperature low DO; <10 ppb high DO; 400-800 ppb 1.0 - 0.32% 0.4 - 0.00004%/s
4-27

environment: temperature: dissolved oxygen: strain range: strain rate:

Characterized growth of small cracks in smooth fatigue test specimens to develop a mechanism of crack initiation in CSs and LASs

Research Program Accomplishments (cont)


Developed interim design curves for different operating parameters; such as, strain rate, DO, S, & temperature (NUREG/CR-5999, 93)
SS curve based only on high DO data, interim curve considered bounding for low DO environment Results used for fatigue evaluations of selected reactor components (NUREG/CR-6260, 95)

Developed statistical models from all existing data for estimating fatigue life (& probability of fatigue cracking) in carbon & lowalloy steels, austenitic SSs, and Alloy 600 in LWR environments (NUREG/CR-6335, 95)
Results used to assess fatigue life and risk for 40-year plant operation (GSI-78 & GSI-166)
4-28

Research Program Accomplishments (cont)


Statistical models were updated with evolving research results
NUREG/CR-6583, 98: CSs & LASs with additional data on DO NUREG/CR-5704, 99: additional data in low DO for wrought SSs & new data on cast SSs

Results from above reports were used in resolution of GSI-190


Evaluation of fatigue life and risk for 60-year plant operation Leak frequency for 40 - 60 year operation

From the above studies and data two procedures have evolved for incorporating effects of LWR environments into fatigue evaluations
Environmentally adjusted design fatigue curves Fatigue life correction factor, Fen (Japan/EPRI)
4-29

Need to Address Environmental Effects and NRC Interaction with ASME


Historical Perspective
Results from studies in Japan (Higuchi & Iida, 91) and ANL (NUREG/CR-4667, 90; NUREG/CR-6583, 98; NUREG/CR-5704, 99) illustrate potentially significant effects of LWR coolant environment on fatigue life of steels The current code design curve for stainless steel in air does not accurately represent the available experimental data (Jaske & ODonnell, 78) and thus includes a reduction of only 1.5 and ~10 from the mean air curve, not the 2 and 20 originally intended

4-30

Need to Address Environmental Effects and NRC Interaction with ASME


S-N Data for Carbon Steel in Water

4-31

Need to Address Environmental Effects and NRC Interaction with ASME


S-N Data for Stainless Steel in Water

4-32

Need to Address Environmental Effects and NRC Interaction with ASME


Current ASME Mean Curve

4-33

Need to Address Environmental Effects and NRC Interaction with ASME


Historical Perspective (cont)
Although existing code fatigue design curves are non-conservative for certain LWR conditions, design practices for fatigue life have been conservative (e.g., higher design loads, larger number of cycles) With improved approaches for fatigue evaluations (e.g., finite element analyses, fatigue monitoring, etc.) environmental effect should be incorporated into code fatigue evaluations to ensure a degree of conservatism consistent with that originally intended

4-34

Need to Address Environmental Effects and NRC Interaction with ASME


Historical Perspective (cont)
Since late 80s, NRC staff have been involved in discussions with ASME Code committees, PVRC, and technical community to address issues related to environmental effects on fatigue In 1991, ASME BNCS requested the PVRC to examine existing worldwide fatigue S-N data and develop recommendations In 1995, resolution of GSI-78 and GSI 166 established that
Risk to core damage from fatigue failure of RCS is very small; no action required for current plant design life of 40 years The NRC staff believe that the fatigue issues should be evaluated for extended period of operation for license renewal (under GSI4-35 190)

Need to Address Environmental Effects and NRC Interaction with ASME


Historical Perspective (cont)
In 1999, resolution of GSI-190 established that:
Based on probabilistic analyses, sensitivity studies, interactions with industry, and different approaches available to manage the effects of aging, no generic regulatory action is required However, calculations supporting resolution of this issue, service experience (NUREG/CR-6582), and nature of age related degradation indicate potential for increase in frequency of pipe leaks as plants continue to operate Consistent with requirements of 10 CFR 54.2, aging management programs for license renewal should address component fatigue including the effects of coolant environment

In December 1999, the NRC requested the ASME Code to address the effects of the LWR environment on fatigue life, and to update the current design curve for Stainless Steel in air
4-36

Need to Address Environmental Effects and NRC Interaction with ASME


Guidance for Managing Environmental Effects on Fatigue
The NRC will review aging management programs for fatigue in support of license renewal on a case-by-case basis until acceptable industry guidance is available We believe it would be useful for the ASME Code to develop methods for incorporating the environmental effects in fatigue life analyses
NRC staff and contractors are working with ASME code committees in Section III and XI for incorporation of such methods

In the absence of ASME action in this area, the NRC may have to develop the needed regulatory guidance
4-37

Need to Address Environmental Effects and NRC Interaction with ASME


Methodology Available for Incorporating Environmental Effects into Fatigue Evaluations
Two methods are available
Using environmentally adjusted fatigue design curves By applying environmental correction factor, Fen, to existing ASME Code fatigue analyses

Environmentally adjusted fatigue design curves and correlations for calculating Fen are contained in NUREG/CR6583 and NUREG/CR-5704 PVRC steering committee on Cyclic Life & Environmental Effects endorsed Fen approach (May/June 1999); and transmitted recommendations to ASME BNCS (Oct 1999) NRC can accept either method

4-38

Future Activities
Complete NRC Research
Characterize growth of small cracks in SS specimens
Better understanding of different behavior with respect to DO

Evaluate the combined effects of loading sequence and environment on fatigue Effect of surface roughness / finish Effect of intermediate temperature (EAC data)

Other Activities
Follow the industry research on thermal fatigue being conducted by EPRI under the PWR Materials Reliability Project Follow the new industry research to be conducted by EPRI to address environmental fatigue issues and component testing Continue to work with ASME Code committees to incorporate effects of 4-39 the environment into code fatigue analyses

5
A REGULATOR'S VIEW ON THE FATIGUE ISSUE
Louis Van der Wiel Ministry of Environment The Hague, Netherlands

Oral Presentation only. No presentation slides or technical paper available.

5-1

A REGULATORS VIEW ON THE FATIGUE ISSUE

Louis Van der Wiel Ministry of Environment The Hague, Netherlands

Oral Presentation only. No presentation slides or technical paper available.

5-3

THERMAL FATIGUE I

6
THERMAL FATIGUE IN FRENCH RHR SYSTEM
C. Faidy, T. Le Courtois Electricit de France EDF-SEPTEN E. de Fraguier Electricit de France EDF-CNEN J-A Leduff, A. Lefranois FRAMATOME J. Dechelotte Electricit de France EDF-DPN

6-1

6-3

6-4

6-5

6-6

6-7

The conclusions of these fracture analyses are : damage mechanism confirmed : high cycle thermal fatigue no fabrication defect deep cracks (one through wall crack and some others up to 80% of the wall thickness) along longitudinal and circumferential welds, starting at the root of the weld on the inner surface lot of small cracks in the weld area, mainly in the ground counterbore areas and high compressive residual stresses regions (less than 3 mm depth) just few cracks in the base metal, generally in ground area, with very limited depth

Fatigue evaluation
In order to evaluate the lead factors of the different degradations, we reviewed the design and construction data and the operating experience of the RHRS. The leak occurred on the first industrial cycle, at intermediate shutdown, with RHRS connected and a RCS temperature of 180C. CIVAUX 1 had more than 1500 hours of operation with a large T(over 150C) in this mixing tee. If we compare the different similar plants and the 2 trains of each of them, we have information for duration between 310 and 2800 hours (table 1).

N4 Plants 1 2 3 4 Table 1 :

Hours of operation Train A / B 2700 / 2800 1600 / 1700 1500 / 1500 450 / 310

Maximum crack depth in the mixing area Nominal thickness : 9.3 mm between 3.2 and 5.8 mm between 0.9 and 4.2 mm between 0.6 and 7.3 mm + 1 through wall between 0. and 1.8 mm

Maximum crack depth in different location of the mixing area for different similar plants

The maximum crack depth is in a reasonnable agreement with the number of hours of operation at large T (over 150C) with one critical location on plant 3 in the elbow (through wall crack after 1500 hours) and more limited degradations on plant 1 with more than 2800 hours of operation. The welding process is in general TIG weld, except for some longitudinal welds that used plasma root pass (figure 6) without weld metal. No fabrication defect has been found in any location.

-8-

6-8

6-9

6-10

6-11

Conclusions
A reasonable understanding of the CIVAUX 1 event has been obtained through a large and detailed analysis program. A new design has been proposed for N4 plants. All repair pieces are designed and fabricated with specific requirements (compressive residual stresses on the inner surface and polished surface finished), with that improvement the expected life of the new design will be over 4000 hours at large T. Nevertheless, in-service inspection and monitoring are required with shorter periods by French Safety Authority to include safety factors. A specific attention has been done to all mixing tees with large T in all safety class piping systems. A step by step procedure is yet under application, a screening criteria has been proposed based on simple model and code fatigue curves. A large complementary 5-years Research and Development program has been launched in 1999, the first results are expected end of this year. It will support all our past analysis and has to proposed some reliable procedure for flaw evaluation under high cycle thermal fatigue. All the French Codes (RCC-M and RSE-M) will be periodically updated to take into account the last developments on this subject : high cycle thermal fatigue. _________________________________

- 12 -

6-12

7
RESULTS OF THERMAL STRATIFICATION MEASUREMENTS AT NUCLEAR POWER PLANT BEZNAU

Suresh Sahgal NOK, Nuclear Power Plant Beznau Switzerland

7-1

7-3

7-4

7-5

7-6

7-7

7-8

7-9

7-10

8
LEAKAGE FROM CVCS PIPE OF REGENERATIVE HEAT EXCHANGER INDUCED BY HIGH CYCLE THERMAL FATIGUE AT TSURUGA NUCLEAR POWER STATION UNIT 2

Takeshi Sakai The Japan Atomic Power Company Tsuruga Power Station Myoujin-cho 1-Banchi Tsuruga-shi, Fukei-ken, Japan

8-1

LEAKAGE FROM CVCS PIPE OF REGENERATIVE HEAT EXCHANGER INDUCED BY HIGH-CYCLE THERMAL FATIGUE AT TSURUGA NUCLEAR POWER STATION UNIT 2
Takeshi Sakai The Japan Atomic Power Company Tsuruga Power Station Myoujin-cho 1-Banchi Tsuruga-shi, Fukei-ken, Japan ABSTRACT On July 12, 1999 while Tsuruga-2, PWR 4-loop plant, was operating at full power (1,160 MWe), unidentified leakage inside the primary containment vessel was detected. As the leakage was identified, the plant promptly started to proceed to cold shutdown. Visual inspection after an isolation of the CVCS (Chemical & Volume Control System) revealed that the leakage was from a connecting pipe between the middle and lower stage in the CVCS regenerative heat exchanger. The CVCS regenerative heat exchanger has three shells, i.e. the upper, the middle and the lower shell. Each heat exchanger shell has an inner cylinder containing a heat exchanger tube bundle. Reactor coolant is cooled while streaming inside the inner cylinder, however, the temperature of the coolant which flows outside the inner cylinder as a bypass flow keeps high. These two coolant flows are mixed around the outlet of the inner cylinder. Thermal-hydraulic mock-up tests simulating internal flows in the heat exchanger were conducted along with thermal and structural analyses. As a result, the mechanism induced cracks was determined to be highcycle thermal fatigue.

1. Introduction On July 12, 1999 while Tsuruga-2, PWR 4-loop plant, was operating at full power (1,160 MWe), the containment sump alarm was also initiated, indicating an increasing leakage rate. As a result of a visual survey inside the PCV, it was recognized that the source of leakage was the CVCS (Fig. 1). The leakage was from a connecting pipe between the middle and lower stages of the CVCS regenerative heat exchanger, as shown in Fig.2. The detailed investigation of the cracked pipe revealed that 12 circumferential and longitudinal cracks including a through-wall crack were identified, as shown in Fig.3. A lot of cracks were found on the inner surface of the middle stage heat exchanger shell. The fracture surface of crack a in the connecting pipe is shown in Fig.4. The fracture surfaces of the circumferential and longitudinal cracks had metallic crystal structure marks and beach marks, which are characteristics of high-cycle, low alternating stress level fatigue. The cracks on the inner surface of the middle stage heat exchanger shell were found to be many, small and multi-directional characteristics of thermal fatigue. As the result of crack investigation, it was believed that all cracks in the connecting pipe and the heat exchanger shell were due to high-cycle thermal fatigue.

8-3

Refueling Water Storage Tank

Containment Vessel

Containment Spray Containment Spray System Hx High Pressure Safety Injection System Pressurizer Accumulator Tank Steam Generator Pressurizer Relief Tank To Turbine

Letdown Inlet

Charging Outlet
From Feed Water Pump

Residual Heat Removal System Hx

Connecting Pipe
RCP Recirculation Sump Reactor Vessel CVCS Regenerative Hx

Flow

Elbow Letdown Outlet

Containment Sump CH/SIP CHP Primary Demineralized Boric Acid Tank Water Tank Volume Control Demineralizer Tank Chemical and Volume Control System

Charging Inlet

Fig.2 CVCS Regenerative Heat Exchanger

Fig.1 Outline of Tsuruga Power Station Unit-2 Reactor System

180

270

90

180

Unit (mm)

n d e f b Weldment

a ~ d j , k, n e ~ i
i k

Longitudinal Crack Circumferential Crack

c Flow

h a

Flow

Fig.3 Cracks in the Connecting Pipe

8-4

Crack

Downstream side The initiation site had no defect. Metallic crystal structure marks on the fracture surface.

Striation like mark in the fracture surface.

Initiation B Initiation C Initiation A Micro simple Inner surface Crack opening length 47mm

Fracture surface Fracture surface was flat. Beach mark on the fracture surface. Crack was started from multi-crack initiation sites.

Upstream side Grow up to big crack from small crack. Crack without branch. Austenitic material without defect. Cold worked layer on the surface by grinding.

Fracture surface

Fig.4 Fracture Surface of Longitudinal Crack a in the Connecting Pipe

#2 Support Ring 180

#1 Support Ring

(Top)

90

(Bottom)

Outlet Nozzle

270

180

Welding Line 154.5 574.5 Inner Surface of Middle Stage Shell Area of UT indications Area of PT indications PT indication Area of UT Impossible 0 100mm

(Top)

Fig.5 The Results of the Liquid Penetrant Examination in the Middle Stage Shell

8-5

2. Structure of Regenerative Heat Exchanger

The regenerative heat exchanger consists of 3 horizontal shells and connecting pipes. Fig.6 shows the structure of the damaged middle stage shell. Each shell has an inner cylinder containing the tube bundle to guide the main flow effectively. The inner cylinder has six support rings to support the internal structure and to limit the bypass flow in the annulus between the inner cylinder and shell to as small as possible. The inner cylinder containing a heat exchanger tube bundle is inserted into the shell in the fabrication process. It results in a small gap between each support ring and the shell. The measured gap between a support ring and the shell was approximately 1.5mm (approximately 3mm in diameter), which allowed a large bypass flow in the annulus between the inner cylinder and the shell, i.e. about 40 percent of the total flow. The bypass flow rate was estimated by the heat-balance calculation based on the actual operating temperature of inlet and outlet of the regenerative heat exchanger. The temperature of the bypass flow is very high (about 250C), but the velocity in the annulus is only 7cm/s. On the other hand, the mean temperature of the main flow passing inside the inner cylinder is approximately 170C as a result of heat-exchange by the tube bundle. As the result of reviewing the structure, it was identified that the two coolant flows (i.e. the lower temperature main flow and the higher temperature bypass flow) could be mixed around the outlet of the inner cylinder, and that there could be a potential for relatively high thermal stress due to the temperature difference. Therefore, thermal-hydraulic mock-up tests simulating internal flows in the heat exchanger were conducted.

Bypass Flow Calculated Value Based on Design Pressure Drop (Approximately 23%) Calculated Value Based on Operating Data (Approximately 40%)

Bypass Flow 0.04m/s 0.07m/s

Bypass Flow (Narrow Part) 1.3m/s 1.6m/s

Let Down Line Inlet 2.6m/s 2.6m/s

Inner Cylinder Inlet 0.8m/s 0.6m/s

Main Flow 0.4m/s 0.3m/s

Inner Cylinder Outlet 0.7m/s 0.5m/s

Let Down Line Outlet 2.4m/s 2.4m/s

Charging Line Outlet Pass Partition Thermal Shield Plate

Let Down Line Inlet

#6 Support Ring Inner Cylinder

Bypass Flow

#1 Support Ring #2 Support Ring Inner Cylinder Outlet

Inner Cylinder Inlet Charging Line Inlet Bypass Flow (Narrow Part)

Main Flow Let Down Line Outlet Fixed Support Leg Sliding Support Leg

Fig.6 Structure of the Regenerative Heat Exchanger (the middle stage)

8-6

3. Investigation of the Causes 3.1 Thermal-Hydraulic Mock-up Tests Thermal-hydraulic mock-up tests were performed in order to understand the internal flow behavior in the heat exchanger (hereinafter Hx). The mock-up test facility, as shown in Fig.7, was fabricated, to represent half of the middle stage Hx and the connecting pipe between the middle and lower stages. All the dimensions were same as the actual Hx in order to simulate the internal flow. Table 1 shows the test conditions compared with the actual flow conditions. Therefore, it was considered that the Richardson number, which is the ratio of buoyancy and inertial force, was a key parameter to simulate the internal flow. The velocity in the mock-up test facility was determined so that the Richardson number could be the identical between mock-up test and actual Hx. The shell and the connecting pipe were made of acrylic resin to observe the flow pattern inside the Hx. And the inner cylinder and the support rings were made of stainless steel in consideration of the actual thermal conduction. The acrylic resin for the shell and the connecting pipe was selected, because it was estimated that the thermal conduction of acrylic resin is almost the same as the actual condition by the insulation outside of the shell and the connecting pipe. Two types of flow patterns, shown in Fig.8, were observed as a result of the test. Flow patterns were identified to depend on the gap at the bottom of the #2-support ring (hereinafter #2-gap). The gap is defined as the distance between the bottom of the #2-support ring and inner surface of the shell. Flow pattern 1 was observed in the case when the #2-gap was relatively small. In this flow pattern, a cold water region appeared in the lower part of the shell between the #2 and #3 support rings. This cold region was produced by a cold main flow, i.e., water in this region was stagnant and cooled by thermal conduction to main cold flow inside the inner cylinder via the inner cylinder wall. In the region between the #1 and #2 support rings, hot bypass water flowed downward from the top of the #2 support ring to the connecting pipe. And cold water flowed downward from the outlet of the inner cylinder to the connecting pipe symmetrically toward the 90 and the 270 sides of the shell. The temperature distribution of the connecting pipe is shown in Fig.8. The 270 side was hot and the 0 side was cold. On the other hand, flow pattern 2 was observed in the case the #2-gap was relatively large. In this flow pattern, the cold water region disappeared in the lower side of the shell between the #2 and #3 support rings, because the hot bypass flow blew out the cold water produced by the cold main flow through the inner cylinder. In the region between the #1 and #2 support rings, hot bypass water flowed asymmetrically. Hot bypass water flowed on the 90 side and all cold water from the outlet of the inner cylinder flowed on the 270 side of the shell. Therefore, hot water flowed from the 0 side of the connecting pipe so that hot water was rotated towards downstream in the connecting pipe, as shown in Fig.8. As figures viewing some crosssections of the connecting pipe shown in fig.8, hot water occupied in the 0 side at the inlet, in the 90 side at the outlet of the elbow. It was also obtained from the mock-up tests showing a transition region between the two flow patterns. Flow pattern 1 occurred if the #2-gap was less than 1.5mm and flow pattern 2 occurred if the #2-gap was larger than 1.7mm, so that the transition region was between 1.5mm and 1.7mm of the #2-gap. Temperature fluctuations were recorded. The largest fluctuation at the shell and the connecting pipe was 23C and 14C respectively, as shown in Fig.9. The period of the fluctuations was almost several seconds. This data was used in a stress analysis.

8-7

Location of Thermocouples #3 Support Ring #2 Support Ring

Table 1 Comparison of Flow Conditions Between the Mock-up and the Actual Hx
Main Flow Bypass Flow Nondimensional Density Difference (/) Representative Size*(D) Representative Velocity*(u) Ri * Actual Hx Flow Rate Temperature Flow Rate Temperature 17.1m3/h 170C 11.4m3/h 250C Richardson Number
Ri = g / D u2

Main Flow Bypass Flow Nondimensional Density Difference (/) Representative Size*(D) Representative Velocity*(u) Ri

Mock-up Flow Rate Temperature Flow Rate Temperature

6.8m3/h 30C 4.6m3/h 70C

#1 Support Ring

0.112 0.08m 0.070m/s 18

0.018 0.08m 0.028m/s 18

Representative size and velocity correspond to the annulus gap and the velocity of the bypass flow in the annulus between the inner cylinder and the shell.

Middle Stage Heat Exchanger Location of Thermocouples Lower Stage Heat Exchanger

Fig.7 Mock-up Test Facility Flow Pattern 1


Lower Offset Cold Water Region
270 180 270 180

Flow Pattern 2
#2 #1
270
0 90

#3

180

Head
90

Upper Offset Cold Water Region Vanishes

#3

#2

#1
270
0 270 180 270 180 90

180

Head
90

Hot Water Cold Water


0 90 270

Hot Water
0 180
0 90

270

2.5mm (180)

Cold Water
0

270 180

0.5mm (180) 1.5mm (270)

90 270 1.5mm (90)

0 90

1.5mm (270)

1.5mm (90) 270 0.5mm (0) 180 90 0 180

270 180 0 90 90

Offset condition

Offset Condition

270 2.5mm (0) 180 90 0 180

270

180 0 90

90

Fig.8 Flow Patterns in the Regenerative Heat Exchanger

250

250

Temperature (C)

Temperature (C)

200

T 23C

200

T 14C

150 0:00

150 0:01 0:02 0:03 Time ( hour : min. ) 0:04 0:05 0:00 0:01 0:02 0:03 0:04 0:05 Time ( hour : min. )

(a) Shell

(b) Elbow

Fig.9 Temperature Fluctuation Data 8-8

3.2 Simulation of Flow Pattern Change Mechanism The flow pattern change mechanism was studied based on the results of the mock-up tests. Fig.10 shows the flow pattern change mechanism. The gaps at the bottom of each support ring are relatively small because of gravity. From the investigation of the actual Hx, the #2-gap was 0.6mm and the #4 support ring touched the inner surface of the shell. Therefore, in an initial condition, the bypass flow could be stagnant in the lower portion between each support ring without a region between the #1 and #2 support rings. This condition produces cold water in the lower portion of the shell as shown in Fig.10 (STEP 1). Deformation of the shell occurs due to the temperature difference between the top and the bottom of the shell (STEP 2). Each bottom gap increases by this deformation so that cold water produced in the lower portion of the shell starts to be swept by bypass flow through the bottom gap (STEP 3). In the next step (STEP 4), as the top and the bottom temperature of the shell become uniform, each gap starts to decrease. Then, deformation of the shell starts to vanish because of this temperature difference decreasing (STEP 5). As the bottom gap decreases, cold water region starts to be produced again, and then condition inside of the Hx returns to STEP 2. A simulation was conducted in order to evaluate this flow pattern change mechanism and to obtain its frequency and the amount of gap change. Fig.11 shows the finite element model used for the simulation. In each section between each support ring, the thermal-hydraulic conditions were input. Some thermal-hydraulic conditions were obtained by the mock-up tests. There were two important parameters. One was the change rate of temperature differences between the top and the bottom of the shell. The experimental data is shown in Fig.12. As the #2 gap decreases (i.e. the #2 support ring moves downwards), the flow of hot water through the lower gap decreases and the cold water region is produced. It results in increasing the temperature differences between the top and the bottom of the shell. The change rate of temperature differences becomes positive. On the other hand, the rate becomes negative when the #2 gap increases. The important point was that one #2 gap had two values of the rate. Another important parameter was the lag time of the changing temperature of the bottom of the shell after finishing the support ring position change. Fig.13 shows experimental data on the lag time . After lifting the support ring, the temperature increased in lag time after the end of lifting support ring (Fig.13 (a)). And also, the temperature decreased in lag time after the end of lowering support ring (Fig.13 (b)). Fig.14 shows the results of the computer simulation. According to the result, the #2-gap changed between about 1.4mm and 1.8mm and this range was wider than the transition region which says two flow patterns could occur repeatedly. The period was about 500 seconds.

8-9

STEP1 (Initial state)


Cold water region grow up in the lower position between each support ring

#1 Support Ring #2 Hot Cold

#3 Hot Cold

#4 Hot Cold

#5 Hot Cold

#6

Thermal Shield Plate

STEP2
Hot Cold Hot Cold Hot Cold Hot Cold

#2 Support Ring #1 Support Ring

#3 Support Ring

#4 Support Ring

#5 Support Ring

#6 Support Ring

Thermal Shield Plate

STEP3

Gap Increasing
Hot Cold Hot Cold Hot Cold Hot Cold

STEP4
Uniform Temperature Hot Cold Hot Cold Hot Cold

Center of the Model Shell Inner Cylinder Sliding Support Leg (Rotation Free)

Fixed Support Leg (Rotation Free)

STEP5

Deformation of the shell almost vanishes

Gap Decreasing
Hot Cold Hot Cold Hot Cold

Hot Cold

Fig.11 Finite Element Model (Half Model)

Repeating STEP2 to STEP5

Sliding Support Gap Decreasing Leg

Fixed Support Leg

Fig.10 Flow Pattern Change Mechanism


0.4 0.3 0.2 0.1 d(T) dt (C/sec) -0.1 + -0.2 - -0.3 -0.4 0 0.5 1 1.5 2 2.5 3 (mm) - 0 0.5 1 1.5 2 2.5 3

: Decreasing Gap : Increasing Gap

Lower Gap of the #2 Support Ring

Fig.12 Experimental Data on the Change Rate of Temperature Difference between Top and Bottom of Hx. Shell

68 67 66 180 180

68 67 66 65.5C End of lowering

Temperature (C)

65 64 Start of lifting support ring 63 62 61 60 59 58 0 59.5C 100 End of lifting

Temperature (C)

0.15C/s(Rate converted to the actual condition) 0 61.6C

65 64 63

0.24C/s(Rate converted to the actual condition)

180

62 Start of lowering 61 support 60 59 58

0 60.4C

200 Time(sec)

300

400

100

200 Time(sec)

300

400

(a) Gap Increasing from 0.5mm to 1.5mm

(b) Gap Decreasing from 2.5mm to 0.5mm

Fig.13 Experimental Data on Lag Time

8-10

2.0 1.8 1.6 Lower Gap (mm) 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0 500 1000 1500 Time (sec) Bypass Flow:40% 2000 2500 3000 #2 #3 #4 Flow Pattern 2 Transition Region Flow Pattern 1

Fig.14 Results of the Simulation of Flow Pattern Change Mechanism

3.3 Evaluation of Crack Initiation and Propagation Crack initiation and propagation analyses were conducted to evaluate the fatigue life due to thermal stress produced by the flow pattern change. This was based on the results of the thermal-hydraulic mock-up tests, the simulation of flow pattern change mechanism and the metallurgical investigation. The thermal stress for the crack propagation analysis was calculated by superposition of the lower frequent temperature change due to flow pattern change and the higher frequency temperature fluctuation due to the mixture of the bypass flow and the main flow. The high average stresses due to operating pressure, thermal transient and residue stress by welding and so on were applied to the cracked pipe and shell. The stress analysis and the experimental method estimated the average stresses. The fatigue life decreased in consideration of the effect of average stress by the hardness and the surface finish investigation at the cracked pipe and shell. The average stresses were compared with the thermal stress due to the flow pattern change. It was evaluated that stresses due to the above-mentioned flow pattern change mechanism were high enough to cause fatigue initiation. Results of the evaluation of crack initiation and propagation, as shown in Table 2, revealed that both cycles of thermal stress were in the order of 10 5. It was determined from the flow pattern change mechanism simulation (approximately 500 seconds) that the cycles calculated by using the plant operating time of Tsuruga-2 (approximately 95,000 hours) were on the order of 105. The calculation results corresponded to the cycles obtained by the crack initiation and propagation analyses. It was assumed that the cause of the cracked pipe and shell was the high-cycle thermal fatigue, because the results of thermal-hydraulic mock-up tests and the stress analysis corresponded to the fracture surface of cracks and the plant operating history.

8-11

Table 2 Evaluation Results of Crack Initiation and Propagation


Cracks Investigation Results of Evaluation Results (105 cycles) Fracture Surface Fatigue Life Depth of Number of Total Fatigue Life for Cracks Beach Fatigue Life for Initiation (1) Propagation (mm) Marks 12.4 11.6 9 12.0 8 6 8 2.0~ 0.3~ 1.9~ 2.4~ a 5.0~6.8 3.1~5.0 1.6~2.0 1.7~2.4 7.0~ 3.4~ 3.5~ 4.1~

a d e Shell Note(1)

Depth values except for crack investigated.

were based on the cracks

4. Conclusion As a result of the investigation, the cause of the leakage from the connecting pipe was considered to be as follows; (1) Flow out of the lower temperature bypass flow region occurred repeatedly at the lower part of the shell, which yielded a cyclic deformation of the shell due to thermal expansion and shrinkage. (2) This cyclic deformation caused a cyclic change of the gap between the inner cylinder support ring and shell, and consequently the cyclic change of the flow pattern at the region where the bypass flow and main flow mixed. (3) Superposition of lower frequent temperature change due to the change of flow pattern and higher frequent temperature fluctuation due to the mixture of the bypass flow and main flow caused high-cycle thermal fatigue cracking.

5. Reference (1) T.Hoshino, T,Aoki, T.Ueno and Y.Kutomi, Leakage from CVCS Pipe of Regenerative Heat Exchanger Induced by High-Cycle Thermal Fatigue at Tsuruga Nuclear Power Station unit-2, ICON-8615, presented at 8 th International Conference on Nuclear Engineering, Baltimore, MD USA (April 2000)

8-12

8-13

8-14

8-15

8-16

8-17

8-18

8-19

8-20

8-21

8-22

8-23

8-24

8-25

8-26

8-27

8-28

8-29

9
OPERATING EXPERIENCE REGARDING THERMAL FATIGUE OF UNISOLABLE PIPING CONNECTED TO PWR REACTOR COOLANT SYSTEMS

Paul Hirschberg Arthur F. Deardorff Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557 John Carey EPRI Project Manager 1300 Harris Boulevard Charlotte, NC 28262

9-1

OPERATING EXPERIENCE REGARDING THERMAL FATIGUE OF UNISOLABLE PIPING CONNECTED TO PWR REACTOR COOLANT SYSTEMS

Paul Hirschberg Arthur F. Deardorff Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557 ABSTRACT

John Carey EPRI Project Manager 1300 Harris Boulevard Charlotte, NC 28262

The U.S. Nuclear Regulatory Commission issued Bulletin 88-08 in response to thermal fatigue failures that occurred at three plants in unisolable portions of piping systems attached to the reactor coolant loop piping. The failure mechanism was thermal stratification caused by valve leakage, and thermal cycling, caused by turbulence penetration. The EPRI Materials Reliability Project (MRP) is currently undertaking a project to develop improved guidelines for screening, evaluation, inspection, and monitoring of thermal fatigue. One of the first steps in developing such guidelines has been to collect operating experience data from domestic PWR plants, to gain an understanding of the quantity and severity of thermal fatigue damage that has actually occurred. By reviewing the failure locations, mechanisms, monitoring results, and corrective actions, more effective tools that address the root causes of actual failures can be developed. The operating experience data collected consists of two groups: major leak events in domestic and foreign plants, and precursors to leakage, such as thermal monitoring results and observed anomalies in domestic plants. Both sets of data will be available to utilities for review in a database format. The emphasis in this paper will be on the latter group, reviewing the experiences of domestic utilities in implementing monitoring and inspection activities in response to the Bulletin. Each PWR plant was contacted to determine what thermal and other monitoring was implemented, the results, whether leaks or cracks were found, modifications that were made, and unusual occurrences observed that were attributed to thermal fatigue. The scope was limited to the Bulletin 88-08 applicable piping systems. This paper summarizes the results of this survey. INTRODUCTION NRC Bulletin 88-08 [1] cited two types of thermal fatigue failures: 1) Inleakage events, in which a leaking isolation valve in the high pressure injection system allowed colder fluid to leak into and stratify in the unisolable portion of the reactor coolant system. The upstream pressure of the injection system, which is driven by the charging pumps, is higher than RCS pressure, thus allowing inleakage. Although not fully understood at the time, cracking was caused by turbulence penetration from the RCS flow into the branch pipe, resulting in thermal cycling between the hot RCS flow and cold leakage flow. The events that precipitated the Bulletin occurred at the Farley and Tihange plants. Outleakage events, in which intermittent leakage out of the stem leakoff line of the isolation valve in the Residual Heat Removal system allowed hot RCS fluid to leak out of the unisolable portion of the RCS. The intermittent leakage caused stratification and thermal cycling in the unisolable section, resulting in a crack. A leak event at the Genkai plant was the source of the concern.

2)

9-3

The Bulletin required all PWR plants to: 1) Review all systems connected to the reactor coolant loop that are normally stagnant to identify unisolable sections that are potentially susceptible to cracking from the thermal fatigue mechanisms described in the Bulletin 2) In locations that may have been subjected to high thermal stresses, perform nondestructive examinations of the welds, heat affected zones, high stress points, and geometric discontinuities to assure that there are no existing flaws 3) Implement a program to assure that thermal fatigue cracking will not occur in these lines by either: a) monitoring the lines to measure thermal stratification and evaluating the results against acceptance criteria, b) preventing pressure upstream of isolation valves from exceeding RCS pressure, or c) installing permanent modifications to enable the system to withstand the stresses. Since the issuance of the Bulletin, other leakage events have occurred, mostly either in foreign plants, or due to scenarios that were different than those anticipated by the Bulletin. Domestic plants responded to 88-08 in a variety of ways, either instrumenting some or all of the lines with thermocouples, installing pressure monitoring systems, measuring valve leakage, or installing various piping system modifications to preclude thermal fatigue. Insight has been gained from the results of the temperature monitoring performed at these plants. This paper summarizes the results of a comprehensive review of the operating experience related to the Bulletin 88-08 thermal fatigue issue, both from the leak events worldwide, and the domestic monitoring experience and anomalies observed in the course of operation. DATA COLLECTION METHODOLOGY AND SCOPE Over 30 references were reviewed to obtain the details of the leak events that have occurred worldwide. This included NRC documents, licensee event reports, conference proceedings, NUREG reports, previously published EPRI reports, operating experience reports, and individual plant failure analyses. Three databases previously compiled by EPRI [2, 3, 4] were searched for relevant items. The NRCs Bibliographic Retrieval System was another source of event information. The primary source of information for individual plant thermal monitoring experience was a questionnaire that was sent to all domestic PWR plants. The questionnaire solicited information on systems monitored, degree of stratification observed, NDE examination results, leaks and cracks found, other monitoring, and modifications made to reduce thermal fatigue. Responses to the questionnaire were followed up with telephone conversations to obtain additional details. Other information was obtained from ASME technical papers, stress analyses performed by Structural Integrity Associates, operating experience reports, plant discrepancy reports, and conversations with cognizant plant engineers. Other thermal fatigue and thermal stratification events have occurred which are beyond the scope of this review. Stratification in the Pressurizer Surge line and Steam Generator Feedwater nozzles are managed by other means and were not included in the scope of the MRP program. Some foreign plant leak events occurred in systems connected to the RCS but were in isolable locations, and were also not included. RESULTS OF REVIEW - GENERAL Worldwide, there have been fourteen leak events recorded in unisolable portions of stagnant systems connected to the Reactor Coolant system [5, 6] in PWR plants. Only seven of these were caused by the mechanisms described in Bulletin 88-08. Of these, after the original Farley crack that precipitated the Bulletin, all occurred in foreign plants. From this standpoint, the actions taken by domestic plants in response to the Bulletin were effective in preventing additional failures. The remaining seven leaks were due to other scenarios not anticipated in the Bulletin but are nevertheless related to the original issue. Four of these pertained to design features that are somewhat unique, and the remaining three had essentially a common root cause. These will be discussed below under the relevant systems.

9-4

Domestic plants actions in response to Bulletin 88-08 varied. Two-thirds of the plants instrumented one or more lines to measure thermal stratification. In most cases, one or two operating cycles were monitored, and if minimal stratification was observed, monitoring was discontinued. One-fourth of the plants performed valve leakage monitoring, either by direct leakage measurement at a test connection, or by applying pressure in a closed section and measuring pressure reduction, or by fluid inventory balancing. In most cases, such monitoring is continuing. Only one plant monitors high pressure injection system pressure upstream of the isolation valve to assure it does not exceed RCS pressure. However, most CE plants, which do not have safety injection systems that operate above RCS pressure, monitor pressure to prevent RCS pressure from entering lower pressure systems due to backleakage through the check valves. About 40% of the plants took some action to mitigate thermal fatigue, such as rerouting piping, adding isolation valves, relocating valves, or improving valve maintenance to prevent leaks. One of the requirements of the Bulletin was for the plants to review their systems and determine which were susceptible to the thermal fatigue issue. The following were the systems considered by the plants to be the most susceptible and /or required temperature monitoring: 1. 2. 3. 4. 5. 6. 7. 8. 9. Auxiliary Spray High Pressure Safety Injection Alternate Charging Charging / Makeup Intermediate Head (Hot Leg) Injection RHR / Shutdown Cooling / Decay Heat Suction Main Pressurizer Spray Low Pressure Injection Drain Lines / Excess Letdown (after 1995)

RESULTS OF REVIEW BY SYSTEM High Pressure Safety Injection This is the system in which the original 88-08 failures occurred, at the Farley and Tihange plants. The system is driven by the charging (or HPI) pumps at pressures higher than RCS pressure. The injection nozzles are smaller diameter than the lower pressure injection nozzles, as low as 1 on the Westinghouse four loop plants. The lines enter the RCS either from above or from the side. At Farley, the crack occurred in the heat affected zone of the weld between the first elbow and the horizontal pipe, about 3 feet from the RCS cold leg nozzle, on a 6 inch NPS line. The crack was caused by hot RCS fluid from turbulence penetration interacting at the bottom of the pipe with cold valve leakage fluid that had stratified. The failure was caused by high cycle fatigue, however the current understanding of turbulence penetration is not sufficient to be able to predict the cycling frequency or the depth of penetration with any certainty. The general rule of thumb is that turbulence penetration effects extend between 5 and 25 pipe diameters into the branch pipe. Four other leak events occurred that were of a similar nature. At Tihange, the crack was located about 2 feet from the RCS hot leg nozzle, in the elbow base metal. At Obrigheim, the crack was located even closer to the RCS, in the nozzle to elbow weld. It had as a contributing cause a deep notch in the circumferential weld that originated in the fabrication process. At Dampierre 2, the crack was in the weld between the check valve and straight pipe just upstream of the hot leg nozzle. At Dampierre 1, the leak occurred in the base metal of the horizontal run between the hot leg and the check valve, about 2 feet from the nozzle. After the failed pipe was replaced, another crack occurred in the same location only nine months later. The replaced section was found to have high residual stresses, and the isolation valve had not stopped leaking.

9-5

A somewhat different failure was found at Biblis, where the crack occurred at a tee that connects a hot and a cold injection line. There were contributing causes of a bound up snubber placing high tensile stress on the pipe, and the presence of pump mechanical vibrations. This system design is not found in domestic plants. It should be noted that none of the safety injection cracks occurred in 1 nozzles as used in the Westinghouse four loop design. Also, Combustion Engineering plants can be considered as not susceptible to inleakage in the safety injection system because the system is not driven by the higher pressure charging pumps; the highest pressure in the system is about 1500 psi, from the SI pumps. Babcock and Wilcox plants have thermal sleeves in the high pressure injection nozzles, protecting most of the region subject to thermal fatigue. The results of temperature monitoring in U.S. plants for this system indicated very little thermal stratification during normal operation unless the isolation valve was leaking. The entire pipe from the loop nozzle to the isolation valve generally remained hot. There was only one case where significant stratification was reported: this was due to valve leakage and resulted in a top to bottom temperature difference of 215F, with cycling having a 2-20 minute period. The valve leakage was corrected and the stratification disappeared. Although there was little thermal stratification during normal operation, several plants reported stratification during plant heatup for a different reason. During periods when not all reactor coolant pumps are running, a loop with the pump turned off experiences a rise in static pressure. These plants reported that the associated check valve did not seat tightly and RCS flow leaked backwards past the check valve and into the other loops. This resulted in stratification gradients of up to 170F. This stratification does not pose a concern because it only occurs for a short time and has only a limited number of thermal cycles. Only one plant installed some form of pressure monitoring on this system. Pressure is checked by a weekly surveillance and is relieved when system pressure exceeds RCS pressure. Two plants installed a second isolation valve to prevent inleakage. Intermediate Pressure Safety Injection (From SI Pumps) This system is fed by the safety injection pumps, which operate at about 1500psi, and usually injects through six inch nozzles in the hot legs. Babcock and Wilcox plants do not have this system, and Westinghouse three and two loop plants use the high pressure injection nozzles for both functions. Combustion Engineering plants use the same nozzle for intermediate head injection, accumulator injection, and low head (shutdown cooling) injection. A number of plants measured a moderate amount of stratification on these lines but not due to valve leakage. Most of the lines connect to the reactor coolant loop from above, followed by a horizontal pipe run. Top to bottom temperature gradients of up to 100F were reported in this horizontal section, due to natural convection from the RCS fluid rising to the top of the pipe, and heat loss to the ambient cooling the fluid at the bottom. The stratification appeared to be constant with no cycling, and therefore is not a concern for a thermal fatigue failure. There were incidents of check valve backleakage, which caused the upstream piping to pressurize to RCS pressure. Most of the CE plants have installed pressure control systems that relieve to the reactor coolant drain tank if backleakage causes the pressure to exceed SI pump pressure. No failures have occurred in these lines. Low Pressure Safety Injection (RHR and Accumulators) The low pressure injection nozzles usually perform the combined function of accumulator injection (at about 600 psi) and RHR / Shutdown Cooling injection (at about 350 psi). On B&W plants, Decay Heat Removal / Core Flood injection is done directly into the reactor pressure vessel. The low pressure nozzles are the largest injection nozzles, between 10 and 14 inches in diameter.

9-6

No leaks have occurred in these lines. Of the plants that monitored these lines, two reported observing thermal stratification due to natural convection, with no valve leakage and no cycling. These lines tend to have longer horizontal runs with the isolation valve located a longer distance from the loop nozzle. One plant had a check valve leak backwards during heatup, causing stratification in the other lines, when reactor coolant pumps were being cycled. In the B&W plants, only a small amount of stratification occurred in the lines connected to the reactor vessel, with no cycling, as flow in the vessel past the nozzle is insufficient to cause turbulence penetration. None of the above are concerns for thermal fatigue, as there is insufficient cycling to cause a failure. RHR Suction / Shutdown Cooling / Decay Heat Drop The RHR Suction line typically drops vertically from the RCS Hot Leg, then travels horizontally for some distance to the isolation valve. Bulletin 88-08 Supplement 3 was issued because of a leak at the Genkai plant. The internals of the isolation valve alternately shrunk and expanded causing intermittent outleakage of RCS fluid through the stem packing and leakoff line. The crack, which occurred at the weld between the first elbow downstream of the hot leg nozzle and the horizontal run, was concluded to have been caused by the intermittent leakage periodically introducing hot fluid at the elbow. The horizontal run loses heat to the ambient, thus causing the outleakage to stratify. Had the leakage not been intermittent, there would not have been enough thermal cycles to cause a failure according to the root cause described in the bulletin. The elbow was 9 pipe diameters from the loop nozzle, well within the range of turbulence penetration from the hot leg flow. In hindsight, it is possible that the cracking was assisted by turbulence penetration causing periodic incursions of hot leg fluid into the horizontal portion of the system, which resulted in the horizontal run periodically stratifying, and the elbow being submitted to cyclic thermal shocks. Most of the responses to the 88-08 Bulletin focused on the valve leakage issue, particularly on stem packing leakage. A number of plants implemented improved valve maintenance programs to reduce packing leaks. Although these programs were effective in stopping packing leaks, they may not have addressed the real root cause of thermal stratification and thermal cycling in the RHR system. The results of thermal monitoring indicated that numerous plants measured thermal stratification in the RHR suction line without any valve leakage occurring. Top to bottom gradients of up to 350F were measured at some plants. A more common temperature gradient was 120F, with the upper temperature cycling +/- 40, 5-15 cycles per day. The cause of this stratification is turbulence penetration of the hot leg fluid extending into the horizontal pipe run, which stratifies due to natural convection. In order to stratify, the length of the vertical run has to be short enough for the hot fluid to reach the horizontal run, but not so short that the horizontal run is always hot, if the isolation valve is located close to the elbow. A number of plants have long horizontal runs, which increases the propensity for stratification due to heat losses. Some plants noted stratification only during heatup or cooldown, probably because changes in RCS flow during those periods changed the depth of turbulence penetration. One plant noted that during power reductions, the greater the reduction in power, the longer was the turbulence penetration distance. Another plant noted that the turbulence penetration was not cyclic but manifested itself in a random helical pattern. Charging / Alternate Charging / Makeup Two leak events occurred in plants having the Babcock and Wilcox Makeup system nozzle design. At Crystal River and Oconee 2, the thermal sleeve became loose and allowed turbulent mixing of hot RCS fluid with cold makeup fluid behind the sleeve. In the B&W design, makeup is not heated by a regenerative heat exchanger and thus the temperature difference between makeup and RCS flow is large. The thermal sleeve was installed with a press fit, and plastic deformation due to hot and cold transients caused it to loosen. The B&W plants have installed an improved thermal sleeve design and implemented a periodic inspection program, such that the problem is considered resolved.

9-7

Westinghouse plants typically run either charging or alternate charging, but not both. The concern is that the line that is not in operation could experience inleakage from the charging pump discharge and have the same mechanism for thermal cycling at the high pressure safety injection lines. Of the B&W plants, one uses two nozzles for normal makeup while the others use one. CE plants use both charging paths, so this issue is not a concern for those plants. The results of thermal monitoring on these lines indicated very little thermal stratification. Either there was no valve leakage, or the distance from the isolation valve to the loop nozzle was short enough that the stagnant fluid did not cool significantly between the regenerative heat exchanger outlet and the loop nozzle. The B&W plant that uses two makeup nozzles reported stratification of up to 325F during heatup, due to backflow caused by pressure differences in the two lines when reactor coolant pumps are not running. Pressurizer Spray (Main and Auxiliary) The auxiliary spray system in Westinghouse and CE plants draws from the charging system and is stagnant during normal operation. If the isolation valve leaks, charging system pressure can produce flow toward the tee with the main spray system. The fluid in the auxiliary spray system is typically cold because the piping length is long enough to cool to ambient. The main spray comes from the reactor coolant loop cold legs, thus there is the potential for a significant temperature difference between the main and auxiliary spray near the tee, and thermal stratification cycling could result if there is unsteady mixing of the flows from the main spray and the leakage. The results of temperature monitoring indicated that thermal stratification was minimal during normal operation, unless the isolation valve was leaking. Two plants reported top to bottom temperature gradients of 200F caused by valve leakage. When the leakage was corrected, the stratification disappeared. Testing and analysis by one plant indicated that the flow velocity of the main spray line was insufficient to cause turbulence penetration to enter the auxiliary spray line. Therefore, although it is possible to have some stratification in the auxiliary spray line from natural convection during normal operation, it is not a concern due to the lack of a cycling mechanism. During heatup and cooldown, several plants noted significant stratification occurring in these lines. When none or an insufficient number of reactor coolant pumps are running to generate sufficient main spray flow, hot steam from the pressurizer enters the lines. This occurs despite the presence of the goose neck loop seal in some of the pressurizer designs. When auxiliary spray was used, a top to bottom temperature gradient of up to 260F was measured. In addition, auxiliary spray tended to be cycled on and off repeatedly especially during heatups, which results not only in thermal stratification cycles but also thermal shocks at the pressurizer nozzle and auxiliary to main spray tee. One plant noted significant stratification during letdown isolations. Apparently there was some isolation valve leakage of charging flow, which was significantly cooler due to the lack of letdown flow in the regenerative heat exchanger. One leak event occurred in Finland at the Loviisa plant. The crack was in a Z-type isolation valve in the auxiliary spray line, vertically above the main spray tee. During heatup and cooldown, pressurizer steam entered the valve outlet, and auxiliary spray entered the inlet; the hot outlet was at a lower elevation than the cold inlet, and stratification cycling occurred internally in the valve. There had also been a preexisting material inclusion defect in the valve that facilitated cracking. Reactor Coolant Loop Drains / Excess Letdown The reactor coolant loop drain lines and the excess letdown line were not recognized as being susceptible to developing unisolable leaks due to thermal stratification cycling in any of the plant reviews in response to Bulletin 88-08. However, four leak events have occurred in these lines. At Three Mile Island, a leak occurred in a cold leg drain line, in the weld between the first elbow downstream of the loop nozzle and the horizontal pipe run. At Oconee, a leak happened in almost the exact same location, this time near the center of the elbow extrados. The TMI drain line was 1 NPS, with an increase to 2 in
9-8

the horizontal run; the distance from the inside surface of the reactor coolant loop to the crack was 14. Similarly, at Oconee the drain line was 1 and the distance to the crack was 13. The crack was caused by turbulence penetration from the RCS intermittently extending into the horizontal run. In both cases the pipe was not insulated, which made it easier for the horizontal portion to stratify. The vertical pipe length was just right for the hot fluid to periodically extend into the horizontal run, but not so short as to keep the horizontal pipe warm all the time. At Mihama, the same mechanism caused a leak in the excess letdown line. The line was 2 NPS and the crack was located 15 from the reactor coolant loop inside surface. One difference was that the Mihama line was insulated; in this case, however, the length of the horizontal run to the isolation valve was very long, resulting in an equivalent amount of heat loss to ambient as a shorter, uninsulated line. There is also evidence that there was an unsteady stratification layer in the elbow. The Mihama elbow was also found to have had high tensile residual stress induced in the fabrication process. The corrective action for this event was to shorten the vertical run such that the turbulence penetration boundary was well into the horizontal run, away from the elbow, which is a point of stress concentration. The other leak event occurred at Loviisa, in a cross tie line between a hot leg drain and a cold leg drain. The crack was in a weld between a reducer and the tee that joins the two drains. Leakage past the cross tie valve allowed hot leg fluid to flow into the cold leg drain. Intermittent thermal expansion of the valve internals caused thermal cycling at the tee. This system design is not found in domestic plants. Since none of the domestic plants initially identified the RCS drains as a potentially susceptible system, none of them performed any temperature monitoring. However, one plant had a leak occur in the excess letdown line in a similar location as in Mihama. In this case the crack was in the heat affected zone of the weld between the first off elbow and the horizontal run, 13 from the reactor coolant loop. The utility concluded that the cause was a thermal interference between a flange on the line and a floor support plate, which generated a stress cycle each time the reactor coolant loop heated up or cooled down. At another plant, a crack was found in a hot leg nuclear sampling system isolation valve. The crack was caused by admitting 600F samples on a daily basis into a normally stagnant line. CONCLUSIONS Thermal stratification and cycling is still a concern for fatigue cracking in lines connected to the reactor coolant loop. However, the susceptible systems can probably be limited to high pressure safety injection, RHR suction, and reactor coolant loop drain lines, or similar lines. More work needs to be done to improve the understanding of turbulence penetration effects, particularly how to determine the penetration distance and the rate of cycling. The details of the worldwide leak events and the domestic plant monitoring experience are being made available in database format on EPRI WEB. The database will be kept up to date by adding new operating experience as submitted by users. REFERENCES 1. Thermal Stresses in Piping Connected to Reactor Coolant Systems, Bulletin 88-08 and Supplements 1, 2, and 3, U.S. NRC, 1988 1989. 2. EPRI Fatigue Management Handbook, TR-104534 Vol. 2, EPRI Palo Alto, December 1994. 3. Nuclear Reactor Piping Failures at U.S. Commercial LWRs: 1961 1997, Report TR-110102, EPRI Palo Alto, December 1998. 4. Piping System Failure Rates and Rupture Frequencies for Use in Risk Informed Inservice Inspection Applications, Report TR-111880, EPRI Palo Alto, March 1999. 5. Experience with Thermal Fatigue in LWR Piping Caused by Mixing and Stratification, Specialists Meeting Proceedings, OECD Nuclear Energy Agency, Paris, June 1998. 6. Assessment of Pressurized Water Reactor Primary System Leaks, NUREG/CR-6582, INEEL, December 1998.

9-9

OPERATING EXPERIENCE: THERMAL FATIGUE OF UNISOLABLE PIPING CONNECTED TO PWR REACTOR COOLANT SYSTEMS
Paul Hirschberg, Art Deardorff Structural Integrity Associates John Carey EPRI July 2000
9-10

NRC Bulletin 88-08


Inleakage past a closed isolation valve between the charging system and the safety injection nozzle caused an unisolable leak at Farley and Tihange Outleakage through the RHR suction isolation valve leakoff line caused cracking in the first elbow off the reactor coolant loop (Genkai) Scope of Bulletin is the ASME Class 1 piping that is attached to the Reactor Coolant Loop and is not isolable from the loop, and is usually stagnant and subject to cyclic thermal stratification
9-11

NRC Bulletin 88-08 - Actions


PWR plants identify systems susceptible to thermal stratification / cycling mechanisms Perform ISI examinations of welds in susceptible locations Assure cracking will not occur by either:
Instrumenting lines to monitor for thermal stratification Monitor pressure upstream of SI isolation valve Valve leakage monitoring Hardware modifications

9-12

Other Thermal Stratification Issues Not in Scope


Pressurizer Surge Line stratification Steam Generator Feedwater nozzle stratification

9-13

EPRI MRP Industry Operating Experience Review


Collect details of Bulletin 88-08 related leak events that occurred worldwide Survey domestic utilities to obtain results of monitoring programs, identify conditions that could be precursors to thermal fatigue cracking, and capture non-reportable events and observations Present information in a database format, to be maintained on EPRI WEB

9-14

Sources of Operating Experience Data


Questionnaire sent to all domestic PWR plants, responses followed up by telephone Three databases previously compiled by EPRI (Risk informed ISI, SKI, Fatigue Management Handbook) NRC Bibliographic Retrieval System (LERs) Operating experience reports Published technical papers Evaluations by utility and consultants NUREGs, conference proceedings

9-15

Results of Review - General


14 leak events have occurred worldwide:
Only 7 were due to scenarios cited in Bulletin 88-08 After Farley, none of these occurred in domestic plants Of the remaining 7 events:
4 were in domestic plants 2 were specific to B&W plants 2 in designs unique to European plants 3 had a common root cause

9-16

General Results - Plant Responses to Bulletin 88-08


67% instrumented one or more systems to measure thermal stratification
Most monitored for only one or two cycles

27% perform some form of valve leakage monitoring 12% have means to monitor and control system pressures 40% took some action to prevent thermal fatigue failures: modified piping geometry, improved valve maintenance, added or changed valves

9-17

General Results - Systems Considered by Plants to be Susceptible to Thermal Fatigue


1. 2. 3. 4. 5. 6. 7. 8. 9. Auxiliary Spray Safety Injection from charging pumps Alternate Charging Charging / Makeup Safety Injection from SI pumps RHR Suction / Shutdown cooling / DHR drop Main Pressurizer Spray Low Pressure Injection RCS Drain lines / Excess letdown (after 1995)

9-18

Safety Injection from Charging Pumps Leak Events


Farley (1987) - HAZ of elbow to pipe weld Obrigheim (1986) - nozzle to elbow weld Tihange (1988) - elbow base metal Dampierre 2 (1992) - valve to pipe weld Dampierre 1 (1996) - straight pipe base metal (cracked again in 9 months) Biblis (1995) - tee connecting hot and cold injection lines None in Westinghouse 1 1/2" nozzles

9-19

Farley and Tihange

In-leakage from higher pressures in charging system Interactions between cold in-leakage and hot Reactor Coolant System fluid caused thermal cycling
9-20

Safety Injection from Charging Pumps Monitoring Experience


Monitoring indicated that stratification was insignificant under normal conditions unless a valve was leaking Stratification occurred during heatup; several plants reported backflow through the check valve when the associated RCP was off and others were running. Cross flow to other loops caused stratification, max. 170 F CE plants are not subject to inleakage Only one plant installed a pressure control system to prevent upstream pressure from exceeding RCS pressure
9-21

Safety Injection from SI Pumps and RHR / Accumulator Injection


No leak events Several plants reported a moderate amount of stratification
Most cases caused by natural convection; no cycling Maximum T = 200 F in long straight run, no leakage One case was due to a leaking check valve which allowed outleakage during heatups CE plants have pressure relief systems to prevent overpressurization from check valve backleakage B&W plant lines injecting into vessel have negligible stratification with no turbulence penetration

9-22

RHR Suction / Shutdown Cooling / Decay Heat Removal


Only leak event was at Genkai
Attributed to intermittent outleakage of RCS hot leg fluid through isolation valve stem packing May have also been due to turbulence penetration of hot leg fluid into horizontal run

9-23

Genkai

Out-leakage occurred at isolation valve leak-off line Cyclic stratification occurred in horizontal piping

9-24

Hypothetical Turbulence Penetration Cycling

9-25

RHR Suction / Shutdown Cooling / Decay Heat Removal


Many plants reported significant thermal stratification, up to 350 F
Most cases were due to turbulence penetration and not the Genkai leak-off line mechanism One case due to a packing leak in the isolation valve Some reported cycling but the period was long (0.2 - 8 days) Some cases were significant only during power reduction; the greater the power reduction, the longer the turbulence penetration distance

9-26

Charging / Makeup / Alt. Charging


Leak Events:
Crystal River (1983) and Oconee 2 (1997) In Makeup / HPI system, B&W plants only Caused by leakage past thermal sleeve allowing turbulent mixing of cold makeup flow with RCS flow Managed by a B&W Owners Group program

9-27

Charging / Makeup / Alt. Charging


Monitoring Experience:
No stratification in charging / makeup except when both makeup trains in use, and crossflow occurs during heatups due to pressure differences from pump cycling Alternate charging had insignificant stratification Some plants use both charging and alternate charging, or equalize their use

9-28

Pressurizer Spray / Auxiliary Spray


Leak event at Loviisa
Crack in auxiliary spray isolation valve due to internal thermal cycling of pressurizer steam and cold auxiliary spray Pre-existing material defect Valve design not used in U.S.

9-29

Pressurizer Spray / Auxiliary Spray Monitoring Experience


Stratification is not significant during normal operation unless isolation valve leaks No turbulence penetration from main spray into aux. spray During heatups and cooldowns, significant stratification (260 F) occurs when RCPs are off and main spray flow is insufficient to fill the pipe. Pressurizer steam refloods the line, and cycling occurs with each auxiliary spray injection. One plant reported higher auxiliary spray stratification (236 F) during periods when letdown is isolated; isolation valve leaked by and charging flow is colder
9-30

Reactor Coolant Loop Drains


Leak Events:
TMI 1 - 1995 Oconee 1 - 2000

Caused by high cycle fatigue due to turbulence penetration Vertical length was about 8 - 9 diameters Horizontal portion of drain line was uninsulated Horizontal portion stratifies due to loss of heat to ambient and intermittent incursion of reactor coolant

9-31

TMI Cold Leg Drain

Isometric of drain line showing details of leak area and location of damaged support
9-32

Excess Letdown
Excess Letdown Leak Event:
Mihama - 1997

Caused by high cycle fatigue due to turbulence penetration, similar to drain line events Vertical length about 8 - 9 diameters Horizontal portion of drain line was insulated, but isolation valve was far away (18 feet) Modification was to shorten the distance from the RCL to the first elbow, to avoid the turbulence penetration boundary

9-33

Modification to Avoid Turbulence Penetration

15.5 in.

15.75 in.

Low Temperature Region Low Temperature Region Interface at elbow pipe results in high thermal stress

Interface at horizontal pipe results in low thermal stress

Before Modification
9-34

After Modification

10.24 in.

High Temperature Region

High Temperature Region

RCL Drains / Excess Letdown Monitoring Experience


One plant had a leak in 1995 on the Excess Letdown line at the first elbow. It was in a similar location to the Mihama event. It was attributed to thermal interference during heatup / cooldown On an RCL drain / sampling line, one plant reported finding a crack in 1994 due to thermal fatigue, caused by admitting 600 F samples into a stagnant line on a daily basis

9-35

Conclusions
Mechanisms for thermal stratification and cycling induced fatigue failures are still present Concern primarily in high pressure safety injection, RHR Suction, and RCL drains Need ongoing monitoring to prevent valve inleakage RCL drains and excess letdown lines should be insulated

9-36

10
AUXILIARY FEEDWATER LINE STRATIFICATION AND COUFAST SIMULATION
Authors : STPHAN J.-M., MASSON J.C. lectricit de France - Industry Branch - Research and Development Division Mechanics and Components Technology 77818 Moret-sur-Loing CEDEX - France

10-1

AUXILIARY FEEDWATER LINE STRATIFICATION AND COUFAST SIMULATION

Authors : STPHAN J.-M., MASSON J.C. lectricit de France - Industry Branch - Research and Development Division Mechanics and Components Technology 77818 Moret-sur-Loing CEDEX - France

ABSTRACT : EDF conducted in the last years different studies on the stratification problems in pipes. One of them consisted of subjecting a metallic full-scale mockup of the steam generator feedwater system to different regimes of stratification. Very useful data were obtained on thermal and mechanical effects of stratification. Then 4000 cycles of fatigue were applied between two stable states of stratification. The 2 levels of the stratification interface were chosen to obtain the maximum possible variation of stresses during the cycle. At the end of the tests, destructive examinations revealed small cracks due to fatigue. The usage factors were calculated using elastic and cyclic elastic-plastic computations. Considering the small depths of the cracks (1.4 to 2 mm) the usage factors (1.3 to 1.9) can be accepted, even if margins are weaker than those obtained in other fatigue studies of mockups subjected to thermal shocks. 1 - INTRODUCTION A thermal stratified flow is characterized by 3 superposed horizontal layers : a cold layer at the bottom of the pipe, a hot layer at its top and an interface between these two layers, which shows a vertical temperature gradient. The conditions required for the creation of a stratified flow are mainly the low velocities of fluids. They exist in certain pipings of nuclear power plants especially in the steam generator (SG) feedwater lines, in some dead ends (safety injection, residual heat removal,), and in some pressurizer surge lines [1, 2, 3]. In the case of a SG feedwater line, the cold flow is generated by water supplied to the steam generator through the auxiliary feedwater system during hot shutdowns. The hot flow results from the heating of water in the internal feedwater ring. A counter-current recirculation can be established in the upper section of the feedwater line. The incoming cold flow does not always suffice to prevent the hot water from flowing backward [1]. The main disturbances due to thermal stratification on nuclear units have been piping displacements and fatigue cracks on pipings. If large lengths of pipe are stratified, the displacements or the forces applied on supports may be considerable (TROJAN power plant [2]). Cracks result from the fatigue damage caused by the variations of the stratified state. In the case of the SG feed line, the variations are due to modifications of the feedwater flow rate. The main cracks mentioned have been noticed in COOK2 (1978) [1], SEQUOYAH (1992), and DIABLO CANYON (1992). Studies of thermal stratification have formed the subject of articles for congresses or reviews [4, 5, 6]. The study presented in this paper concerns the stratification in the SG feedwater system. It is original in that, at the end of a knowledge acquisition phase, a cycling of the mock-up under fatigue conditions has been carried out. 2 - DESCRIPTION OF THE COUFAST MOCK-UP The COUFAST mock-up is made up of a 90 elbow connected to 2 straight pipes (figure 1). Its dimensions and material are similar to those used in the SG feedwater system of the French PWR 900. It differs principally in the length of the horizontal section, and in the downward section which is vertical instead of being tilted at 45. It has been built according to the rules applicable to the French nuclear constructions [13]. The only exceptions are the 3 welds (among 5) in the horizontal section. Their designs differ deliberately from the standard rules,

10- 3

either in the taper angle or the radius, or in the shot-peening finish (figure 2). The aim was to study the effects of geometry or shot-peening on the fatigue resistance. The cold water is injected in the lower part of the mock-up through a diffuser to minimize the disturbance in the elbow. The hot water inlet nozzle is set four meters away from the elbow. The outlet nozzles are set at the end of the horizontal section at a distance of 6 meters from the elbow. When the stratification takes place in the mockup, a part of the hot water flows toward the elbow in a counter-current direction. It forms the upper hot layer of the stratified flow. This flow reproduces the flow back from the SG ring toward the elbow. As this flowrate is not known, it was necessary to investigate various hot water inflows, combined to various cold water inflows, to check the situations existing in the PWR plants. Mechanically speaking, the model is supported in 2 points. The mock-up is embedded on a concrete bed-plate on the horizontal section about 5 meters away from the elbow. A weight-support in the middle of the vertical section takes up partially the weight of the mock-up (figure 1). It was adjusted to give no momentum at the fixed point when the mock-up is filled with cold water. The connections with the water feed loop leave the model free to bend under effect of thermal stratification. 3 - THE INSTRUMENTATION OF THE MOCK-UP A great number of thermocouples (about 60) are set on the outer wall, particularly on the elbow and its weld, and in front of the 3 thermocouple rods (figure 3). The inner wall is fitted with 3 sets of 6 thermocouples in front of the thermocouple rods n3 and 4 (fig. 4). One of them emerge in the water but the others are embedded at a depth of 0.5 mm in the inner wall. They are very close to each other, spaced by only 0.5 mm. These 3 sets are used to measure the heat exchanged between the fluid and the inner wall in the different layers of the stratified flow. The vertical profiles of fluid temperature are measured along three vertical diameters (figure 3). Two opposite vertical rods, each bearing 18 thermocouples (1 per cm), are set in each diameter (figure 5). Fifty strain gauges are set on the outer wall as shown on figure 4. The follow-up of cold thermal shock tests has made it necessary to set 25 additional strain gauges and 10 associated thermocouples on the inner wall. 4 - THE PERFORMING OF THE TESTS The nominal thermal-hydraulic conditions for testing were as follows: - pressure of 80 bars, - cold water at 60C with flow rates ranging from 0.1 to 9 m /h, - hot water at 280C with flow rates ranging from 0.1 to 5 m /h. The thermal-hydraulic capacities of the water feed loop have limited the sum of the cold and the hot inflows to 10 3 m /h. Figure 6 shows the flow diagram of the SUPERBABETTE loop which feeds the model. Four tests have been carried out at hot water temperatures other than 280C so as to get closer to stratified states measured on the plants. Complementary tests of fast thermal shocks were made at the end-of-life of the model. Some modifications of the loop were made to obtain an instantaneous increasing of cold water inflow. 5 - TESTS AT STABLE STRATIFIED STATES Some of the results presented in this paragraph have already been published [7]. 5.1 - Thermal-hydraulic aspects The tests have been carried out by keeping constant the cold and the hot incoming flow rates until the temperatures in the model had stabilized. Half an hour at least is necessary to let the temperatures stabilize. The stratification level stabilizes in a short period of time, but the vertical thermal gradient in the stratification interface increases slowly. Because thermal stresses depend on the thermal gradient, it is necessary to wait for a stable state to obtain the maximum stress values. Different combinations of incoming flow rates have been investigated as shown in figure 7. Below a cold flow rate 3 3 of 5 m /h, the level of stratification is independent of the hot inflow (if this one is over 0.5 m /h). Above a cold 3 flow rate of 5 m /h, results are more and more dependent of the hot inflow, and the comparison is not satisfactory
10- 4
3 3

with in situ measurements. This discordance is assumed to result from the diameters of the outlet nozzles which are too small. The position of the stratification interface and the maximum vertical thermal gradient (see the profile in figure 8) 3 both increase as the cold flow rises (see also 5.3). When the cold water inflows is below 0.5 m /h the cold water 3 stream is heated by the wall. For example, in the case of a cold flow rate of 0.1 m /h at 60C, the temperature of the bottom increases to 100C just after the elbow and then rises as a function of the distance from the elbow. This makes the thermal stresses smaller for very low cold inflows. No significant variations in the temperature of the stratification interface have been noticed during the tests. The maximum peak-to-peak variation was 30C at a frequency ranging from 0.1 to 1 Hz. Thermal-hydraulic calculations have been made successfully on the model configuration [8]. 5.2 - Thermomechanical aspects A heat exchange coefficient cannot be easily defined in the zone of the stratification interface because the fluid temperature varies abruptly with the altitude. However the measurements on the mock-up enable a determination of a global transfer coefficient . Its value was about 3500 W/m/C, that is much higher than those determined in the other layers of the flow. No quantitative explanation has yet been found. At the interface level, the circumferential thermal gradients on the wall are softened. Their ratio relative to the thermal gradient in the fluid, is about 1/2 on the inner wall and 1/4 on the outer wall (less for low stratifications). The global displacements of the model were measured at the point where the support takes up the weight. They are considerable and can reach 100 mm in the vertical direction and 50 mm in the horizontal direction. The measurements of stresses is reported in the chapter 8. 5.3 - Comparison with in situ measurements The figure 9 summarizes the comparison with measurements made on the SG feedwater system in actual PWR plants. The positions of the stratification interface are quite comparable. The thermal gradients exceed those measured in the plants, because of care taken to obtain stabilization. As indicated in 5.1 the thermal-hydraulic behavior of the COUFAST mock-up is representative of the behavior of SG feedwater systems, only if the cold 3 3 water flowrate is less than 5 to 6 m /h (and the hot water flowrate more than 0.5 m /h). Then, with unchanging temperatures, the height H of the interface level correlates remarkably with the cold water flowrate. Based on the COUFAST results and on the measurements in 3 EDF plants, it has been possible to establish a simple mathematical model which gives the relative height H/D depending on the Froude number (figure 9). 6 - TESTS OF QUICK ESTABLISHING OF STRATIFICATION In the SG feedwater systems the flow rate may change frequently and quickly. So it was advisable to evaluate the value of stresses on the inner wall in these conditions. Modifications of the facilities were achieved to apply cold thermal shocks to the mock-up. Four tests have been carried out, starting from two different initial states and for two cold water flow rates. In a first phase a cold front spreads out horizontally over a height about 8 cm (figure 10). Then, in a second phase, the cold front rises up, turning into a horizontal stratification, while the vertical thermal gradient gradually increases in the fluid, reaching temporarily 180C/cm. Thermal variations measured during the thermal shocks (maximum peak-to-peak = 80C) are small. The additional stresses measured on the inner wall during thermal shocks are weak (about 10 MPa). From the fatigue viewpoint, they do not cause any other significant damage because they set up in zones which differ from those affected during stable stratified states. 7 - CYCLING TESTS UNDER THERMAL FATIGUE CONDITIONS 7.1 - Choice of the cycling conditions The most damaging cycle (figures 7 and 11) has been defined on the basis of measurements made during stable stratified tests, and was confirmed by numerical studies. It was found to be a cycle between 2 stratified states corresponding to positions of the stratification level of about 50 and 70 (azimuthal angle from the bottom). The

10- 5

greatest amplitude of stresses in the straight section (about 320 MPa) is reached on the inner wall at an azimuthal angle of 60. Water chemical conditions during these cycling tests were imposed to comply with the maximum values authorized for SG feedwater systems in the French nuclear power plants. Stratification can occur only when the steam generator is supplied through the Auxiliary Feedwater System as during hot shutdowns.

In this case the ultimate values are : pH (at 25C) : 8.8 to 9.3, oxygen content lower than 100 g/l, cationic conductivity lower than 2 S/cm.

7.2 - Realisation of cycling This cycle has been applied about 4000 times on the model in four series of about 1000 cycles. Non destructive examinations (gamma radiographic on welds and ultrasonic on the elbow) have been made prior to the cycles and after each series. No evolution has been noticed and no indication was found. The respect of the oxygen content is of a prime importance for fatigue life of ferritic steels. The readings have proven that the maximum value reached was about 30 ppb (or g/l) during only a very short period of time. The average value was about 5 ppb, which corresponds to the maximum admissible value under normal operating conditions (power > 25% nominal power). The condition imposed in French PWR plants has been respected and so it is not useful to apply the correction proposed for example by EPRI (environment factor) when, among other things, the dissolved oxygen in water is higher than 100 ppb [9, 10, 11]. 7.3 - Examination after cycling At the end of its life, the destructive examinations of the model revealed small cracks in the welds tapers. The cracks were situated on each weld in the area affected by the variations of the stratification interface position*.
reference of weld height / diameter (H /D) crack lengths (mm) depths (mm) {left / right) S1 0.27 57 1.35 / 1.25 S2 0.2 30 1.4 / 2.0 S3 0.19 33 1.95 / 1.8 S4 0.17 42 2.8 / 0.75 S5 0.13 (39) (1.0)

* Note : Taking into account the curvature of the mock-up induced by the stratification, the heights of the cracks have to be corrected. Reporting the relative heights to the location of the rods 3 & 4, H/D* = 0.19 which is exactly the average of the 2 stratification levels.

The biggest crack is 2.8 mm deep and is located on the most severe taper. The weld S5 which is shot-blasted doesn't show any crack in the tapers, but only one (1 mm) at the junction with the weld fillet. 8 - NUMERICAL SIMULATIONS OF THE TESTS 8.1 - Numerical simulation of static tests All the calculations described later were made with the Code_Asterdeveloped by EDF R&D Division [12]. The work first consisted of validating the optimum meshing in the straight section of the piping before starting a complete 3D calculation. We used a 2D finite element model with generalized plane deformations. The measurements in the section of the 3 & 4 rods (figure 3) were used for the validation because this section included many thermocouples. It was possible to compare the fluid temperature profiles, the temperatures on the inner and the outer wall, and the stresses measured by strain gauges at 6 points on the outer wall. The validation is made for seven stratified states giving different positions of the stratification level. Using the specific transfer coefficient (see 5.2) the temperature profiles calculated on the inner and the outer wall are close to those measured (fig. 12). Calculated and measured stresses are also in good conformity (fig. 13). On the inner wall, calculated stresses are the highest on the flank just below (120 MPa in traction) and just above (200

10- 6

MPa in compression) the stratification level. They depend on the level of the stratification in amplitude but more significant is the variation of the position of the extremum of these stresses which follows the level of the stratification. Consequently, the maximum stress amplitude is reached when stratification is cycled between two stable stratified states. The optimal states were determined to be the S_43 and the S_26 (see figure 7). A complete calculation using a 3D finite element model has then been made for the same seven stratified states. The comparison between stress measurements (located in the current zones) and calculations was satisfactory, too. The highest stresses calculated were in the weld tapers, so the following computations concentrated on the simulation of these tapers (3 geometric shapes) by 3D elastic finite elements. It should be noted that the stress intensification factor Kt depends on the level of stratification (1.5 to 2). For the conditions of the tests, Kt 1.8 close to the factor K3 = 1.7 tabulated in ASME III & RCC-M [13] for as-welded joints. 8.2 - Numerical elastic calculations of the fatigue generated by thermal cycles For fatigue calculation it is essential to know the stress amplitude range at the location where it is the highest. Elastic computations performed for the 3 different shapes of the tapers allowed to select these locations for the chosen cycle of stratification (between the states S_43 - S_26). The variation of stresses during a cycle is obtained by the difference between stresses calculated for the 2 stabilized stratified states. The vertical profiles of temperature for these 2 states were imposed as loading conditions (figure 11). The mesh used for the computation of the mock-up (1/4 of the model with regard to symmetry conditions) is shown on figures 14 and 15. The boundary conditions were "no longitudinal displacements" in the center of the weld (plane of symmetry), and "free end" displacements at the other end ( lengthened to simulate an continuous pipe). The vertical profile of the stress intensity (Tresca) variation during a cycle shows (figure 16) a steep peak which is situated just between the minimal and maximal heights of the stratification interface for the 2 endings of the cycle. The interpretation of the results was made according to the chapter B 3200 of the RCC-M (or ASME III) because in the chapter B3600 (Piping analysis) it is not convenient to introduce non-axisymmetric loading (as stratification). It gives the following results for the weld S2 (standard weld on the straight section) : - range of total stress - range of primary plus secondary stress Sp : 540 MPa Sn : 420 MPa

The stresses mentioned above were calculated taking into account the dependence of thermo-elastic characteristics (Young's modulus and coefficient of expansion) on the temperature. They lead to a usage factor of about 1.3 for the 4000 cycles carried out on the mock-up. We used also the new rule introduced in the RCC-M code for austenitic stainless steels (June 1994) to evaluate its impact (The specific rule for ferritic steels is under development). The plastic concentration factor Ke increases then from 1.3 to 1.5. Thus the usage factor becomes 1.9 which is in accordance with the small depth of the cracks. Without dependence of characteristics on temperature, the usage factor are respectively 0.7 and 1.4 (see the recap in the table 8.3).
Note : The main advantage of the new rule is to be more realistic and less sensitive to methods of calculation. Using the new rule, the usage factor of the weld with an excessive tapering (S3 with an angle of 20) decreases from 8.7 to 4, while it increases from 1 to 1.5 for the normal tapers S1 and S2 (these calculations were made with the same hypotheses and methods which were differing from those of the recent calculations reported herein).

8.3 - Numerical elastic-plastic calculations of the fatigue generated by thermal cycles Considering the important role of the factor Ke (for simplified elastic-plastic analysis), elastic-plastic computations have been made according to RCC-M method (calculation of the range of the deformation). The elastic-plastic cyclic behavior of the A42 ferritic steel (~ SA 106 carbon steel) has been estimated at room temperature by fitting different curves of ferritic steels given in literature. It has been precisely measured at the maximum temperature of the cycle (280C) on specimens of the same material used for the mock-up making. The figures 17 and 18 present the cyclic curves (used just as they are in the non linear isotropic modeling) and the linearization used in the linear kinematic hardening modeling. The incremental computations take into account the internal pressure (first stage). The second stage consists in going from the isothermal cold state to the first stratification state (S_26) with an intermediate state (see figure 11) to avoid an abnormal plastification. Then the fatigue cycles (S_26 - S_43 - S-26) are applied. Four different cases are calculated. The first uses an isotropic modeling and the material behavior (stress-strain curve) is defined by averaging for each value of strain the stresses at 20C and 280C. The cyclic stress-strain

10- 7

curve at the most loaded point is shown on figure 19. The behavior becomes completely elastic during the second part of the first cycle (plastic adaptation) so the computations were not carry on beyond. The second case takes into account the dependence of the behavior curves on the temperature by interpolating (linearly) between the 2 basic curves. The figure 20 shows a plastic adaptation from the very first cycle. The third case is a calculation using a linear kinematic model with a cyclic curve taken as the average of the 2 basic curves. The computation needed four cycles to obtain a stabilization (this was probably due to numerical problems). The figure 21 shows the plastic accommodation (non linear open cycle) and the figure 22 shows the evolution of the equivalent strain (based on Tresca and assigned with an algebraic sign). The fourth case with the linear kinematic model used the dependence of curves on temperature (linear interpolation of the parameters of the 2 models at 20C and 280C). There is again a plastic accommodation but after optimization of the numerical parameters, the stability is obtain from the first cycle (S_26 - S_43 - S_26). The following table gives the results in term of equivalent strain range and the usage factors obtained by full elastic-plastic analysis and by simplified elastic-plastic analysis (called "elastic" in the last column ; see 8.2) :
reference of calculation isotropic at average T kinematic at average T isotropic depending on T kinematic depending on T
pl el equivalent strain usage Ke* = / range (%) factor (full elastic-plastic)

RCC-M rule for Ke (Sn) before 1994 new (austenitic) before 1994 new (austenitic)

Ke factor usage factor "elastic" "elastic" analysis 1.19 1.48 1.32 1.5 0.69 1.37 1.28 1.9

0.29 0.29 0.32 0.41

0.60 0.62 0.78 1.75

1.13 1.13 1.15 1.46

It is well known that the isotropic modeling of the cyclic behavior gives an underestimation of the actual behavior of the structure, and the kinematic modeling gives an overestimation. The Ke factor used following an elastic calculation is determined as being the upper boundary of the rate (Ke*) between strain range obtained by elastic-plastic calculation and that obtained by elastic calculation. This is always verified for the new rule introduced in the RCC-M. For the previous rule (the same as in ASME Code) the Ke factor is under the elastic-plastic Ke* obtained with the kinematic model depending on temperature. Only the usage factor obtained by this last modeling is upper 1 and could be considered as conservative. 9 - CONCLUSION The study of the COUFAST mock-up was conducted in full scale (dimensions, temperatures, pressure) of the steam generator feedwater systems. Much important knowledge was acquired during the tests at stable stratified states (temperature profiles in the fluid, thermal transfer coefficients on stratification interface). The elastic calculations were validated by comparisons with measurements. After that, fatigue tests were conducted by cycling between the 2 stable stratified states giving the most severe alternating stress range (very severe in comparison with actual situations in plants). The destructive examinations after 4000 cycles revealed small cracks in the tapers of welds. The usage factors were calculated with different methods. They are acceptable in respect with the depth of cracks if they are calculated with the new rule introduced in the RCC-M. Elastic-plastic computations are conservative only by using a kinematic hardening model for material cyclic behavior. It is strongly recommended to take into account the dependence of behavior on temperature. REFERENCES [1] US Nuclear Regulatory Commission : Cracking in feedwater system piping, Bulletin 79-13, June , 1979. [2] USNRC : Thermal stresses in piping connected to Reactor Coolant Systems, Bulletin 88-08, June 1988. [3] USNRC : Pressurizer Surge Line Thermal Stratification Bulletin 88-11, December 20, 1988. [4] Deardorff A.F., Kim J.H., Roidt R.M. : Thermal stratification in Nuclear reactor piping system, ICONE, Japan, November 1991 [5] Wolf L., Schygula U., Geiss M., Handjosten E. : Thermal stratification tests in horizontal feedwater pipelines, NUREG/CP-0091, Volume 5, 1987. [6] Wolf L., Hfner W., Schygula U., Geiss M., Handjosten E. : Results of HDR-experiments for pipe loads under thermally stratified flow conditions, Nuclear Engineering and Design 137 (1992)

10- 8

[7] Stphan J.M., Caron J. : COUFAST : Experimental study of mechanical consequences of thermal stratification on a piping elbow, ASME PVP , New Orleans, June 1992 [8] Pniguel C., Stphan J.M. : Thermal hydraulic and mechanical behaviour of an elbow in presence of a stratified flow, NURETH-5, Salt Lake City, September 1992 [9] Chopra O. K. : Status of Fatigue Issues at Argonne National Laboratory, EPRI Conference on Operating Power Plant Fatigue Issues & Resolutions, Snowbird (Utah), August 1996 [10] Gosselin S. R. : An environmental Factor Approach to Account for Reactor Water Effects in LWR Pressure Vessel and Piping Fatigue Evaluations, EPRI Conference, Snowbird, August 1996 [11] Mehta H.S., Gosselin S.R. : An environmental Factor Approach to Account for Reactor Water Effects in LWR Pressure Vessel and Piping Fatigue Evaluations, PVP-Vol. 323, Volume 1, ASME 1996 [12] Lvesque J.R. : Code_ASTER , EDF DER 1996, http://www.edf.fr/der/html/produits/production [13] AFCEN : Rgles de Conception des Chaudires nuclaires-Mcanique (RCC-M), 1993, AFNOR, Paris [14] Masson J.C., Stephan J.M. : Fatigue induced by thermal stratification - Results of tests and calculations of the COUFAST model, OCDE workshop on thermal stratification, Paris, june 1998

10- 9

AUXILIARY FEEDWATER LINE STRATIFICATION AND COUFAST SIMULATION


J.M. STEPHAN , J.C. MASSON EDF/R&D Division Mechanics and Technology of Components
International Conference on Fatigue of Reactors Components July 2000

10-10

Plan General overview of COUFAST model and of testing conditions Stable states of stratification Fatigue cycling Numerical interpretation Conclusion

10-11

Thermal Hydraulics phenomena on PWR pipings 1.1 - Different types Thermal Stratification Mixing Zones Turbulent Penetration

COOK 2 DIABLO CANYON SEQUOYAH


10-12

CIVAUX

FARLEY TIHANGE DAMPIERRE

Thermal Hydraulics phenomena in PWR pipings 1.2 - Main difficulties : - Local phenomena - Inside piping phenomena - Difficulties to obtain thermal characteristics with CFD calculations

Needs for experimental evaluation in complement of on site measurements


10-13

Exemple : Thermal Stratification on COUFAST model 2.1 - General overview of COUFAST model and of testing conditions 2.2 - Stable states of stratification 2.3 - Fatigue cycling 2.4 - Numerical interpretation 2.4 - Discussion on methods for prediction of fatigue 2.5 - Conclusion

10-14

weight supported

WELDS S1 S2 S3 S4 S5

hot water ! outlet " inlet (280C)

" embedded 280C stratification interface cold water (60C) cold water outlet 406.4 mm 21.4 mm

Diameter Thickness

cross-section

Ferritic A42 steel

! cold water inlet (60C)

General overview of the COUFAST mock-up


10-15

Water Chemistry

pH (at 20C) : oxygen content :

9,0 +/- 0,3 < 31 g/l < 2 S/cm

Cationic conductivity :

10-16

results of destructive examinations

reference of the weld


crack length (mm) crack depth (mm) (elbow side) crack depth (mm) (oulet side)

S1

S2

S3 (20 )
33 1.95

S4

S5

(elbow) (straight)
57 1.35 30 1.4

(r = 0) (shot)
42 2.8 (39) (1.0)

1.25

2.0

1.8

0.75

10-17

weight supported

WELDS S1 S2 S3 S4 S5

hot water ! outlet " inlet (280C)

" embedded 280C stratification interface cold water (60C) cold water outlet 406.4 mm 21.4 mm

Diameter Thickness

cross-section

Ferritic A42 steel

! cold water inlet (60C)

General overview of the COUFAST mock-up


10-18

2D calculations of stable states of stratifications 1 - to adjust the values for heat exchange coefficients cold water (forced circulation) : hot water (natural circulation) : interface
(special exchange)

750 W / m2K-1 2000 W / m2K-1 3500 W / m2K-1

2 - to optimize the mesh for 3D calculations

32 elements on the lower quarter of the cicumference 2 elements through the thickness

10-19

3D calculations of stable states of stratifications - 7 states of stratification to analyse the influence of the height of the stratification level the sharpness of the gradient

- 4 different geometries to compare the elbow and the weld S1 connecting it to the pipe the tapers in welds S2, S3 and S4

10-20

main results for stable states of stratifications

- good adjustment of the thermal model - validation of mechanical calculations - parametric studies (4 geometries - 7 states)
guided us to chose the most appropriate cycle for fatigue gave us indications on stress intensification in the tapers

10-21

stress intensification factors

geometry stratif. height 55 mm 70 mm 95 mm 115 mm 125 mm 150 mm 190 mm

S1
1.98 1.78 1.78 1.77 1.75 1.50 1.72

S2
1.78 1.77 1.73

S3
2.63 2.60 2.56 2.66

S4
1.88 1.88 1.92

10-22

comparison of Ke and usage factors for average T plastic calculations


reference of calculation isotropic average T kinematic average T usage factor 0.60 0.62

elastic calculations

Ke*

RCC-M rule Ke (Sn) before 1994 new rule


(austenitic)

Ke factor 1.19 1.48

usage factor 0.69 1.37

1.13 1.13

Ke* = epl / e el

10-23

comparison when depending on temperature

plastic calculations

elastic calculations

reference of calculation isotropic


depending on T

Ke*

usage factor 0.78 1.75

RCC-M rule Ke (Sn) before 1994 new rule


(austenitic)

Ke factor 1.32 1.5

usage factor 1.28 1.9

1.15 1.46

kinematic
depending on T

Ke* = epl / e el
10-24

results of plastic calculations

behavior's model for calculation


isotropic average temperature linear kinematic average temperature isotropic
depending on temperature

equivalent strain range (%) 0.29 0.29 0.32 0.41

usage factor 0.60 0.62 0.78 1.75

linear kinematic
depending on temperature

10-25

CONCLUSION

- much knowledge on stratification - validation of thermal and elastic calculations - severe fatigue cycling small cracks

- usage factors acceptable if using the new RCC-M rule


- elastic-plastic calculations conservative if kinematic modeling

10-26

11
FATIGUE EVALUATION IN PIPING CAUSED BY THERMAL STRATIFICATION
Eduardo Maneschy and Rodolfo Suanno Eletronuclear S.A. Universidade do Estado do Rio de Janeiro Rio de Janeiro, RJ Brazil

11-1

FATIGUE EVALUATION IN PIPING CAUSED BY THERMAL STRATIFICATION Eduardo Maneschy 1 and Rodolfo Suanno 1,2 Eletronuclear S.A. 1 Universidade do Estado do Rio de Janeiro 2 Rio de Janeiro, RJ Brazil

ABSTRACT This paper presents the fatigue evaluation of the thermal cycling stratification in the residual heat removal (RHR) auxiliary line of Angra 1 nuclear power station. Since the phenomenon is plant specific, the investigation is conducted through an experimental study (temperature measurements) and numerical analysis (heat transfer and stress calculations). The temperature profile is obtained using thermocouples installed on the external surface of the pipe, at five circumferential locations. These data are used as an input to a transient heat transfer model in order to obtain the temperature distribution through pipe section. The temperature difference in the section produces local stresses (due to the non-linear profile across the pipe section) that are superposed to the global ones (moments due to restrained thermal expansion) to obtain total response. A three-dimensional finite elements model is used for the thermal and structural analyses, being the numerical solution carried out by ANSYS program. The part of the RHR considered is class 1, and the fatigue calculation to obtain the residual life of the piping is performed according to the ASME Code Section III Subsection NB-3200 requirements.

INTRODUCTION Since the late 80s thermal stratification in piping has been one of the most important concerns in the nuclear industry. At least two plants, Farley 2 (USA) in 87 and Tihange 1 (Belgium) in 88, have reported primary coolant leakage due to cracks in elbow piping of an emergency core cooling system. The damage is attributed to high cycle thermally induced fatigue mechanisms, which produces through-wall cracks in welds (Farley) and base metal ( ihange). At Genkai 1 (Japan) in 88 a similar defect was T reported in the weld of an elbow caused by thermal cycling in the unisolable part of the RHR system, [1]. Additional information of these events is available in the USNRC Bulletin 88-08, [2]. Dampierre 2 in 92 and Dampierre 1 in 96, both in France, [3] and [4], and Oconee 2 (USA) in 97, [5], also developed through-wall crack in unisolable parts of piping system due to thermal stratification. The current industry data indicate that the thermal cycling is critical for fatigue and appears to depend on piping configuration and on the power level of the plant. This phenomenon eventually is responsible for cracks in horizontal section of the piping caused by intermittent valve leakage and/or turbulent penetration, [6]. As a result of these problems, and to avoid future leakage, corrective and preventive actions were considered in many plants. To take into account the thermal stratification, a transient not included in the original design specification, a temperature monitoring program was implemented in Angra 1. The pressurizer surge

11-3

line was the first system selected to be investigated. The analysis was conducted by Westinghouse in 1992, and concluded that stratification load would not impact the original design, [7]. Recently, in response to a question raised by the Brazilian Licensing Authority, a screening was performed in Angra 1 to find out which systems could be susceptible to stratification. The evaluation indicated the RHR suction line between the reactor coolant system and isolation valve as a potential location. In order to quantify the severity of the phenomenon, a temperature monitoring system was installed on the piping outside surface. Since the result of the plant measurements has shown a high level of thermal cycling, a structural analysis was conducted to evaluate the stratification effect on the lifetime of the line. The present paper focuses on experimental and numerical approaches adopted to perform the fatigue assessment of the RHR piping. The results of the monitoring program are used as an input to heat transfer and stress evaluation. Since the length of horizontal run upstream of the isolation valve is short, stratification induces local thermal stress through the pipe wall. The resulting stress should be combined with global stresses caused by bending deflection (restrained thermal expansion). Both thermal and structural analysis are developed employing three-dimensional finite elements models. Numerical solution is performed by ANSYS program, [8], and the stress analysis is conducted according to ASME Section III Subsection NB-3200 Code, [9].

MONITORING TEMPERATURE IN ANGRA 1 Angra 1 is a 657 MW Westinghouse two loops plant. The RHR system comprises a 8 inches sch 160 pipe (outside diameter is 219 mm and thickness is 23 mm) made of stainless steel (ASME SA 376 TP 304). The isolation valve is located in the horizontal section, approximately ten times the outside diameter away from the connection with main reactor coolant piping (hot-leg). Part of the layout is shown in Fig.1. The temperature monitoring system consists of thermocouples mounted on the outside surface of the piping, in a section located 0,20 m upstream of the valve. The sensors are installed at five different angular points around the pipe, according to the schematic representation indicated in the Fig.1. The data are obtained continuously at every two minutes since the end of 1998. Figure 2a is a typical result showing the temperatures measured at 100% power operation as a function of time. The result during a day is nearly the same since the beginning of the measurements program. For loop # 2 there are two distinct responses between the different fluid layers. The thermocouple at the top of the pipe shows that the response is cyclic, and the temperature changes from 225o C to 300o C in approximately 40 minutes. The remaining thermocouples indicate constant temperatures, with values at the four circumferential locations equal to 198o C, 167o C, 157o C and 144o C. The cyclic response is strongly dependent on the power level of the plant, as can be observed comparing Figure 2a (100% operation) and Figure 2b (60% operation). A smaller level of stratification, without oscillation, is observed in loop # 1. The temperature distribution at 100% power is plotted as a function of relative height (sensor location normalized to the outside pipe diameter) and depicted in the Fig.3. Configurations that correspond to the fluctuations between maximum (300o C) and minimum (225o C) temperature at the top are shown. Notice

11-4

RCS (Hot Leg)

0,2 m

Section AA

60

60

Pipe cross section AA ( Thermocouples positions)

Fig. 1 Piping layout and thermocouples locations

11-5

2a 100% power operation

350 300 250 200 150


Bottom Top

Temperature ( C)

100 50 0 00:00 04:00 08:00 12:00 Time (hr)


2b 60% power operation

16:00

20:00

00:00

300 250 Temperature ( C) 200 150 100 50 0 4:00 8:00 12:00 Time (hr) 16:00 20:00 0:00
Bottom
o

Top

Fig. 2 Measured temperatures at pipe outside diameter as a function of hours of day

11-6

1 Relative height (sensor location/outside diameter)

T min

T max

0,5

0 0 -0,5 50 100 150 200 250 300 350

-1 Temperature ( C)
o

Fig. 3 Maximum and minimum temperature stratification through pipe section (relative height = 1 top ; relative height = 1 bottom)

0 45 90 Section AA

Pipe cross section (viewed from valve side)


0

Fig.4 Piping layout for three-dimensional model

11-7

that top-to-bottom maximum temperature difference on the outside wall reaches approximately 156o C, being the profile through the section continuous. By contrast, a discontinuous one is observed in the pressurizer surge line. It is important to mention that at least two walkdowns inside containment building during operation were conducted, and a leakage through the isolation valve was difficult to be identified. Besides that, the level of the reactor coolant drain tank was also checked, and no reportable variation was noticed. It appears that the Angra 1 thermal cycling might not be produced by the active mechanism that occurred in Genkai 1, [1], but can be similar to those developed due to turbulent penetration induced by the hotleg flow, [6]. Accordingly, temperature fluctuations at the top are caused when a hot turbulent flow from the hot-leg penetrates into the RHR branch line and interacts with the cooler fluid at the horizontal part of the piping.

THERMAL ANALYSIS Because the instrumentation is mounted on the external surface of the pipe, the fluid condition inside is not known and the temperature distribution through the pipe section can not be determined. In order to overcome this difficult, a practical alternative should be adopted. In this paper the basic idea is to assume an initial condition for the fluid and verify, using transient heat transfer calculation, if the temperature response matches the values measured outside. To achieve this goal, an iterative procedure is necessary. Numerical method to solve the problem is based on the three-dimensional finite element method. The thermal model is constructed using the eight-node element (ANSYS SOLID70), with the pipe thickness meshed using four elements. The system is simulated in the stratified region and includes part of the vertical section of the line, elbow, horizontal portion and pipe-valve transition zone as schematically represented in Fig.4. The flow is restricted to a small area located in the upper part of the horizontal pipe, and the boundary condition is the temperature profile presented in Fig.3. At this area the water is supposed to heatup in five seconds from 225o C to 315o C (hot-leg temperature). The values in other points through the section are unchanged. The temperature history at the pipe is obtained adopting a heat transfer coefficient equal to 1600 W/m2 o C, a number selected in a conservative basis, and considered to cover a wide range of flow rates. These values are consistent with the measured temperatures and represent the final picks after several trials, which include variation in the fluid temperature and its time to reach maximum. The thermal model for the minimum stratification (initial stage, beginning of the transient) is represented in Fig.5. It can be noticed that at the external surface the temperature in the upper part is 225o C, and decays to the quantities readings in the other thermocouples (locations and results as shown in Fig.1 and Fig.3). The temperature distribution at final stage (end of the transient), when the value at pipe outside reaches 300o C, is shown in Fig.6. This is the maximum stratification and occurs 300 seconds after starting the calculation. The corresponding temperatures at these instants are reproduced in Fig. 7 and 8 as a function of pipe internal circumference. The numerical solution is obtained for three sections at the elbow and the section where the thermocouples were installed (see Fig.4). It can be observed that such as the measured data, the change in calculated temperature is limited to a small area located at the top portion of the pipe.

11-8

Fig. 5 Temperature distribution at the minimum stratification level (beginning of the transient, 225o C at top pipe)

Fig. 6 Temperature distribution at the maximum stratification level (end of the transient, 300o C at top pipe)

11-9

350

300

250 Temperature ( C)
o

200

150

100

Section AA Elbow 90 Degrees Elbow 45 Degrees Elbow 0 Degrees

50

0 0 100 200 300 Pipe circumference (mm) 400 500 600

Fig. 7 Temperature distribution around pipe internal circumference (beginning of the transient, 225o C at top pipe)

350

300

250 Temperature ( C)
o

200

150

100

Section AA Elbow 90 Degrees Elbow 45 Degrees Elbow 0 Degrees

50

0 0 100 200 300 Pipe circumference (mm) 400 500 600

Fig. 8 Temperature distribution around pipe internal circumference (end of the transient, 300o C at top pipe)

11-10

A close examination of the data confirms that the numerical simulation approximately reproduces the measured temperatures (see Fig.9), giving an indication that the through-wall solution is adequate for structural analysis. Notice that the highest stresses are induced during steady-state response, when the greatest top-to-bottom temperature difference is achieved. Therefore, the stratification analysis differs from thermal shock problems, when the stresses associated with through-wall gradients occurs during the transient portion of the loading, and tends to disappear when the steady-state is reached.

STRUCTURAL ANALYSIS The structural analysis is conducted with a three-dimensional finite element model, constructed employing a eight-node element (ANSYS SOLID45). The mesh is represented in Fig.10, and is built with 21120 solid elements. The thermal stresses are evaluated using the most unfavorable temperature gradient through the pipe section, derived from the heat transfer results, which occurs at the end of the transient. In order to save computer running time only a part of the piping layout is modeled, with boundary conditions imposed using substructuring technique. Therefore, the effect of the removed parts are calculated previously and used as stiffness matrices at the ends. This implies that the restraints are not rigid and the analysis considers the flexibility of the removed pipe. The applied load is represented by the temperature distribution along the piping, imposed in three different regions: from the hot-leg connection down to the vertical pipe (0,80 m before the elbow, see Fig.1) it is assumed as 315o C. Between this point and the pipe-valve interface the stratification profile as obtined by the thermal analysis is imposed. After the isolation valve a 40o C (ambient temperature) is adopted. Global stress resulting from the moments produced by the ends restraints is combined with local stress caused by the thermal gradient through the pipe section. The total stress is obtained summing up these two components. In the present work it is assumed that the mechanical loads (dead weight and seismic) are negligible when compared to the thermal stresses. Therefore, they are not included in the calculations. The input for stress analysis considers the values generated by the finite element solution and is performed using the rules of the ASME III NB-3200 Code [9]. Accordingly, primary plus secondary stress intensity range is evaluated first (shakedown criterion), followed by the peak stress intensity range calculation. Finally, the ASME fatigue curves are employed to determine the cumulative usage factor, which shall be less than the allowable value of 1,0.

RESULTS The study is performed considering the minimum and maximum stratification illustrated in Fig.3. Because the stress is proportional to the temperature difference, the values corresponding to the intermediates levels of stratification are obtained by linear interpolation. On the other hand, the original stress analysis performed by Westinghouse, [10], calculated the usage factor very small, which means that the design transients are negligible compared with thermal cycling loads. The analysis considering the upper part of the pipe at 225o C is performed first. The results of the calculation are presented in Fig.11. As next, the top of the pipe is assumed to be at 300o C. The analysis

11-11

Relative height (sensor location/outside diameter)

0,5

0 0 -0,5 50 100 150 200 250 300 350

-1 Temperature (oC)

Fig. 9 Measured () and numerical () data comparisom at the maximum stratification (relative height = 1 top ; relative height = 1 bottom)

Fig. 10 Solid model for structural analysis

11-12

is conducted and the resulting stresses are shown in Fig.12. The detailed observation related to the stress plots defines the horizontal pipe and isolation valve interface as the most critical location. Another component which is highly stressed is the elbow. The fatigue analysis is carried out using the stress range when the internal pressure and the temperature goes from one condition to another in time. Since the shakedown criterion is met, the concentration factor Ke is 1,0, and the alternating component is half of the peak stress range. The alternating stress intensity is amplified by 2,0, a number supposed to cover the effect of the structural discontinuity at the pipe-valve interface, and by the ratio of the modulus of elasticity given on the ASME fatigue curve to the value used in this analysis (Ec/Eh =1,11). Two cases have to be considered. The first one is when the operation changes from zero load condition (cooldown) to 100% power, an event assumed to occur 200 times. The pressure plus thermal fluctuation stress range is 329 MPa, plotted in Fig.13, and the usage factor is 0,008. The second case to analyze is when the temperature fluctuation at the top of the pipe takes place. During the transient the pressure is essentially constant and does not contribute to the total stress. At 100% power the stresses due to thermal cycling are determined considering the differences at two load conditions: transient down (225o C) and up (300o C) at the top part of the pipe, being the remaining values at other regions held constant (see Fig.3). This is supposed to occur nearly 36 times a day or at a rough guess, 420000 for plant life (40 years and availability factor of 0,8). As shown in Fig.14, the stress range at the pipe-valve location is 182 MPa. The calculated cumulative usage factor is 0,84. As it can be noticed the stress range is controlled by the temperature oscillation at the upper part of the pipe. However, the temperature amplitude at this portion is not constant (see Fig.2a) and, consequently, the above usage factor is overly conservative. A better life estimation is obtained if an appropriated cycle counting, such as rainflow method, is considered. If this technique is adopted, and neglecting the small fluctuations (less than 30o C), only 20 meaningful cycles per day have to be counted (235000 cycles during plant lifetime). As a result, the cumulative usage factor is reduced to 0,47. Despite of the above number being less than 1,0, a nondestructive examination was conducted during the last Angra 1 outage (May 2000). To investigate if a crack developed due to thermal cycling, the ultrasonic method was adopted, being the locations selected the elbow and the horizontal piping sections upstream the isolation valve. The examination was concentrated on the top of the pipe, the region where the highest stresses take place (see Figs.11-14). All welds and base metal were inspected, and no indications were found. Anyhow, the mitigation of the problem shall continue to be achieved. To reduce the stress fluctuations, operational and maintenance procedures, structural modifications or a combination among them, are highly desired. Even though this does not eliminate thermal cycling, the damage would be less severe.

CONCLUSION The thermal cycling stratification phenomenon at Angra 1 RHR piping was analyzed by an experimental and numerical approach. The data used in the study were obtained at 100% power operation, measured for a period from the end of 1998 to mid of 1999.

11-13

Fig. 11 Stress distribution due to minimum stratification level (beginning of the transient, 225o C at the top pipe)

Fig. 12 Stress distribution due to maximum stratification level (end of the transient, 300o C at the top)

11-14

Fig. 13 Stress distribution due to pressure and maximum stratification when the operational condition changes from zero load to 100% power

Fig. 14 Stress distribution at 100% when the stratification level changes from minimum (225o C) to maximum (300o C) at the top pipe

11-15

Since no leakage indication was found, it appears that the thermal fluctuation occurs due to turbulent penetration. Two other reasons support this idea: 1) the isolation valve location related to hot-leg is within the limit to turbulent penetration, which is assumed to be between 10 and 25 branch line internal diameters ; 2) the characteristic of the temperature distribution through the cross section shows a mixture between hot and cold fluid (by contrast, during leakage hot fluid is concentrated at the top of the pipe). However, additional investigation has to prove this statement. The temperature profile on the pipe surface was used as input to thermal and structural threedimensional finite element models. For the fatigue evaluation it was assumed 420000 cycles throughout the plant lifetime. This is a conservative estimation, based on stratification between maximum and minimum levels equal to 36 times a day. If the rainflow technique is adopted this value is reduced to 20 cycles per day. From the stress analysis results, the critical location is the interface between the pipe and isolation valve. At this point the calculated usage factor is equal to 0,84 (conservative) or 0,47 (realistic), both less than the allowable value. This indicates that the thermal cycling stratification observed in Angra 1 does not have impact on the integrity of the RHR piping.

REFERENCES 1. Shirahama, S., Failure to the Residual Heat Removal System Suction Line Pipe in Genkai Unit 1 Caused by Thermal Stratification Cycling, Experience with Thermal Fatigue in LWR Piping Caused by Mixing and Stratification, June 8-10, 1998, France, p.73-81. 2. USNRC, Thermal Stresses in Piping Connected to Reactor Coolant System, Bulletim 88-08 and Supplements 1, 2, and 3, U.S. Nuclear Regulatory Commission, June 22, 1988. 3. Navarro, G., Determination of the Thermal Loading Affecting the Auxiliary Lines of the Reactor Coolant System in French PWR Plants, Experience with Thermal Fatigue in LWR Piping Caused by Mixing and Stratification, June 8-10, 1998, France, p.358-362. 4. Gauthier, V., Thermal Fatigue Cracking of Safety Injection System Pipes Non Destructive Testing Inspections Feedback, Experience with Thermal Fatigue in LWR Piping Caused by Mixing and Stratification, June 8-10, 1998, France, p.436-453. 5. Lund, A.L. and M. Hartzman, USNRC Regulatory Perspective on Unanticipated Thermal Fatigue in LWR Piping, Experience with Thermal Fatigue in LWR Piping Caused by Mixing and Stratification, June 8-10, 1998, France, p.41-48. 6. EPRI, Thermal Stratification, Cycling, and Striping (TASCS), TR-103581, March 1994. 7. Westinghouse, Structural Evaluation of the Angra Unit 1 Pressurizer Surge Line, Considering the Effects of Thermal Stratification, WCAP-13685, Westinghouse, 1994. 8. ANSYS, ANSYS Finite Element Code Users' Manual-Revision 5.4, Swanson Analysis Systems Inc., Houston PA, 1997. 9. ASME, ASME Boiler and Pressure Vessel Code, Section III Subsection NB and Appendices, American Society of Mechanical Engineers, 1989. 10. Westinghouse, ASME Section III Piping Stress Analysis for the Angra Nuclear Generating Station Unit 1, WCAP-9630, Westinghouse, 1994.

11-16

THERMAL FATIGUE II

12
CONSIDERATION OF THERMAL FATIGUE AND CYCLE MONITORING OF A B31.1 PLANT
Kenneth Chang American Electric Power

Oral presentation only. No presentation slides or technical paper available.

12-1

Consideration of Thermal Fatigue and Cycle Monitoring of a B31.1 Plant

Kenneth Chang American Electric Power

Oral presentation only. No presentation slides or technical paper available.

12-3

13
MAIN RESULTS OF EDFS EXPERIENCE ON IN SITU MEASUREMENTS RELATED TO THERMAL FATIGUE ON PWR REACTOR COOLANT PIPING

Frdric DULCERE, Gilles NAVARRO Electricit de France, Research and Development Division 6, quai Watier - 78401 CHATOU FRANCE- 33 .(0)1.30.87.78.06 frederic.dulcere@edf.fr, gilles.navarro@edf.fr

13-1

INTERNATIONAL CONFERENCE ON FATIGUE OF REACTOR COMPONENTS


July 31 - August 2, 2000 Napa, California

Frdric DULCERE, Gilles NAVARRO Electricit de France, Research and Development Division
6, quai Watier - 78401 CHATOU FRANCE- 33.(0)1.30.87.78.06 frederic.dulcere@edf.fr, gilles.navarro@edf.fr

MAIN RESULTS OF EDFS EXPERIENCE ON IN SITU MEASUREMENTS RELATED TO THERMAL FATIGUE ON PWR REACTOR COOLANT PIPING

1. INTRODUCTION
The various incidents imputed to thermal fatigue, which occurred throughout the world on the auxiliary lines of the Reactor Coolant System (SIS, RHR, CVC), led EDF to urge a research program in order to determine the origins and the consequences of these problems. The stakes related to these problems concern, at the same time, the safety and the availability of the units. This program was based on instrumentations of the nuclear power plants in operation whose principal results are presented hereafter.

2. DEFINITION OF THE INSTRUMENTATIONS


Specific instrumentations were installed on the auxiliary lines of several power plants representative of the French nuclear capacity. The strategy of standardized plants enabled us to equip a limited number of installations while keeping a global sight of the phenomena affecting the whole of our nuclear plants. Thus, two 900 MW units (Blayais 1 and Cruas 4), one 1300 MW unit (Golfech 2), and more recently one 1400 MW unit (Chooz B1) were instrumented. The international experience feedback showed that the defects had occurred near or on the welds. Therefore, it is close to these that the majority of the sensors was installed. The choice of the positions of the transducers was defined according to four principal objectives: characterization of the thermal loadings supported by the auxiliary lines under operation, recording of the transients for comparison with the conception transients, highlighting of thermal cyclings, discovery of unexpected phenomena. The instrumentations mainly consisted of thermocouples welded on the external skin of the pipes (30 to 140 according to the site) and pressure pick-up measurements (approximately ten). For each

13-3

instrumentation, the principal signals of exploitation were also recorded in order to be correlated with the local data. For each instrumented unit, uninterrupted acquisitions lasted several cycles of operation in order to be ensured of the representativeness of the observations carried out. This monitoring was made with acquisition periods ranging between 1 and 20 seconds which made it possible to accurately record the consequences of the main operating conditions on the instrumented lines.

3. UNISOLABLE PIPING LOADINGS


A good knowledge of the unisolable piping damage constitutes a very significant stake from a safety point of view. The availability is the second objective as incidents, on such part of the lines, induce long stops of the units (More than 50 days for Dampierre 1 SIS crack). The incident that occured at Dampierre 2 in 1992 was in accordance with discovered crackings at Farley and Tihange. Then, the instrumentation program aimed at determining these pipes damage risks by improving our knowledge of the thermohydraulic phenomena which sit there.

3. 1. TURBULENT PENETRATION
The principal phenomenon affecting the unisolable piping of the non-discharging lines connected to the main primary system is related to the shearing of the principal flow at the nozzle. It results in the penetration of primary fluid in the auxiliary line and structure in a gimlet shape (vortex). The penetration length varies between 15 and 20 times the hydraulic diameter of the pipes. Even if this distance varies according to the thermohydraulic conditions of the flow and the lines geometry, the phenomenon was observed on the whole of the instrumented lines.

3. 2. CONSEQUENCES UNDER NORMAL OPERATION 3. 2. 1. HEAT PROPAGATION


According to the length of the unisolable piping and the vortex penetration depth, the head of the vortex can reach the first isolation valve. In this case, the temperature of the latter is close to the primary temperature. When the turbulent penetration does not reach the first isolation valve, two cases arise: either the head of turbulence is situated in a vertical pipe and the propagation of heat is stopped (the vertical pipe is hot at the top and cold at the bottom), or the head of turbulence is in a horizontal pipe and a free convection current forms and propagates heat to the valve or to the first vertical part going down. Figure 1 shows an example of conditioning of a horizontal auxiliary line where the vortex is relayed by a free convection current which leads to a very high temperature at the level of the valve.

13-4

Ho tL eg
302 C L1 14 D= 275 C 302 C L2 19 D= 272 C 265 C 301 C

L3 24 D=

Figure 1 : Heat propagation in the unisolable piping

The turbulent penetration does not lead to any fluctuation of temperature in so far as it only mixes hot fluid. If it meets a lower temperature fluid, coming for example from a leaking valve, slight cyclings can be observed. It is what the instrumentation highlighted, in particular on SIS lines and on draining of crossover leg lines. These cyclings are regular and do not vary with operating conditions. They do not present particular risks.

3. 2. 2. VORTEX INSTABILITY
An unexpected phenomenon particularly caught our attention. It relates to an accumulator injection line in Cold Leg (figure 2) where a complex thermohydraulic phenomenon creates repeated temperature shocks upstream of a vertical part (figure 3). Its appearance frequency is random and it even completely disappeared during several months. No correlation with operation of exploitation was found.

13-5

L3 = 21 D

L1 5 D =
RCP 321 VP

RPE

L2 = 18 D

Figure 2 : Instrumentation of the accumulator injection line These thermal shocks are attributed to length penetration variations of the primary vortex whose head is near the elbow. Turbulence head movements occur, which, from time to time, go beyond the elbow and generate hot shocks in the higher part of the horizontal pipe, then go back causing slow coolings of this part of the line.

Figure 3 : Example of thermal shocks For this line, and some others, the designer (FRAMATOME) carried out fatigue calculations. All in all, these calculations showed, that under normal operation, the integrity of the auxiliary lines was not called into question in spite of the difference between the loadings expected and those measured. Nevertheless, some lines (of which the line seeing the temperature random peaks) are today the object of a particular control program, because the fatigue resistance margins are from now on reduced.

13-6

BF 3

3. 3. DAMAGED VALVES CONSEQUENCES : FARLEY-TIHANGE PHENOMENON 3. 3. 1. SIMULATION OF A SIS/CVC LEAK AT BLAYAIS 1


With an aim at better understanding the risks related to a leak of the isolation valves between the main primary system (155 bars) and the CVC (180 bars), in particular at the origin of the incidents of Farley and Tihange, a specific installation was installed at Blayais 1 (figure 4). It consisted in the creation of a bypass of one of the isolation valves, equipped with two manual valves and a flowmeter, in order to simulate a leakage with variable flow from CVC towards the primary system.

Manual Regulating Valve

Flowmeter

HL 1
7 6 5 4 3 2 1

MD

HL 2

H H S I

7 6 5 4

HL 3
5 4 3 2 1

Figure 4 : Functional diagram of the injection test Injections staged between 0 and 300 l/h were carried out. They led to significant modifications of the lowest line thermal loadings, in this case the SIS line in Hot Leg 1. For the other lines, no sign of cold fluid was detected in the range of studied flow rate.

13-7

Hot Leg 1

Position of the manual valve

Cycling measured at the invert of the pipe


Section 1 Section 2
Section 3

Closed

1/2 tr

2 tr

Opened

Figure 5 : Cyclings evolution during the injection test

More precisely, the observation of the unisolable part of this line teaches us about the interaction of the primary vortex and the cold water leakage coming from the high pressure system. As the leak-flow increases, the area of interaction moves away from the valves. The more important the cold water leak is, the more the zone of interaction moves towards the RCS. In the range of studied flow rate, the amplitude of cyclings also increases with the leak-flow. Upstream of the valve, the cold fluid injection results in an important cooling of the pipe. The thermomechanical calculations based on these results enabled us to determine an approximation of the flow rate limit leading to welds damage.

3. 3. 2. FEEDBACK FROM THE DAMPIERRE 1 INCIDENT


Double cracking at Dampierre 1 in 1997 was especially surprising by its localization in the straight part of the pipe. It mainly showed that our knowledge of the phenomena was still imperfect. Until then, the welds had always been regarded as the most sensitive points of the auxiliary lines. In fact, the interaction between the primary vortex and the cold water leakage is very local and the instrumentation of Blayais 1 did not make it possible to extrapolate the results obtained at the two ends of the horizontal pipe to the straight part of the line.

13-8

Among the investigations carried out following the discovery of the cracks, temperature measurements during hot shutdown on the damaged line, allowed, by comparison with Blayais 1 injection tests data, to identify the origin of the problem as being the SIS/CVC isolation valve leakage. While waiting for the generalization of a modification equivalent to that of Tihange (pressure well), these temperature measurements are now carried out with each starting in complement of the tests aiming at highlighting a possible cold water leakage. A more precise instrumentation was installed on the SIS line in Hot Leg 1 of Dampierre 1 in order to improve our knowledge in the accused zone and to compare, in a wider way, the line loadings with those of Blayais 1. The valve separating the SIS and CVC circuits having been repaired, this instrumentation showed that without leak, thermal loadings are equivalent between the two units.

4. DEAD-END BRANCH LINE LOADS


The damages noted in the dead-end branch lines (in between the two isolating valves) are generally low, and few through-wall cracks were highlighted in these zones. The safety aspect is less critical there than on the unisolable part, and the principal stake concerns maintenance and plant availability.

4. 1.

HEAT PROPAGATION PHENOMENON

The instrumentations showed that, by conduction through the first isolation valve when it is hot, then by free convection, the temperature of most of auxiliary lines is high in the dead-end branch line. This propagation can extend over significant lengths on horizontal piping. Temperatures higher than 100C were recorded at more than 150 D from primary nozzle (figure 6). Only a vertical pipe going down can stop heat conduction. The insulation removal associated with a good ventilation of the building can also contribute to the reduction of the temperatures, but the result remains random.
107C 187C

L2 = 156 D
RCP 422 VP RPE RCP 418 VP

L 1 = 140 D

BU2

Figure 6 : Heat propagation on a draining of crossover leg line

13-9

4. 2. CONSEQUENCES 4. 2. 1. WATER/STEAM OR WATER/AIR DIPHASIC ENVIRONMENT


The main risk related to a high temperature in the dead-end branch line is a chemical risk because the temperature favors the process of corrosion. The probability that this phenomenon leads to defects would be extremely limited if piping contained only water because the chemical elements likely to corrode metal remain in weak concentration. The risk worsens when a diphasic environment is established in the dead leg. This diphasic environment can have two origins: A bad venting involving a water/air environment. This probability is related to the alignment and the vent valves positioning which do not always allow a perfect venting. A temperature in the dead leg higher than the saturation temperature, which generates a water/steam environment. Indeed, the pressure variations in some lines can be quite important (between 1 and 155 bars) when it is not imposed by the auxiliary line operation (example: pressure forced at 42 bars on the accumulator injection lines). In both cases, the presence of a water/steam or water/air interface causes the concentration of corrosive elements at the interface (boric acid for example). This phenomenon results in the appearance of level lines which can evolve in generalized corrosion according to the concentration of the corrosive elements and the environment oxygen content. These damages are slow and represent a stake mostly from a point of view of availability and maintenance costs. Besides, oxygen presence associated with a high level of stress (average stress or residual stress) favors the appearance of stress corrosion cracking. Unlike generalized corrosion, this degradation mode can be fast and involves several through-wall cracks (Bugey 3 SIS, draining of crossover leg lines of Cattenom 1 and Fessenheim 1).

4. 2. 2. THERMAL STRATIFICATION
The presence of a free convection current in the dead-end branch line causes the appearance of thermal stratification. Maximum amplitudes of 80C were observed. They are at the origin of strain and stress, the effects of which combine with the stress normally supported by the piping.

13-10

195 C

L4 29 D =
119 C 168 C

L5 31 D=
110 C

130 C

L6 72 D=
109 C 79 C L7 75 D=

Figure 7 : Stratification in the dead-end branch line

5. CONCLUSIONS
The whole of the instrumentations results helped us improve our knowledge of thermohydraulic phenomena affecting RCS auxiliary lines of units in operation. The unisolable pipes are subjected to a hot vortex influence which, combined with a cold water weak source, can generate thermal fatigue damage. The difficulty in having a perfect control of the cold water arrivals led EDF to urge a modification on all 900 MW units in order to remove this source of degradation. Concerning the dead-end branch line, the presence of a diphasic environment mainly related to high temperatures can lead to the appearance of degradations of chemical origin from level lines till stress corrosion cracking. From a lawful point of view the gathered data are used to check the fatigue behavior of sensitive auxiliary lines in spite of significant differences between the real loadings and those used for the design. These results enabled us to improve our inspection strategy and preventive maintenance of the auxiliary lines. They are also used to determine in a realistic way the stresses supported by a given line, or to justify the behavior of some defects discovered during inspections. The results of these instrumentations are finally used for future units design by taking into account several cycles of acquisition on the 900 MW units, but also on the 1300 MW and 1450 MW units, where some innovations resulted from the experience feedback of first measurement series.

13-11

14
EVALUATION OF OCONEE-2 HIGH PRESSURE INJECTION/NORMAL MAKEUP (HPI/NMU) LINE WELD FAILURE

Bret L. Boman Framatome Technologies Lynchburg, VA James R. Smotrel Framatome Technologies Lynchburg, VA J. Michael Davis Duke Energy Charlotte, NC Timothy D. Brown Duke Energy Seneca, SC

14-1

Evaluation of Oconee-2 High Pressure Injection/Normal Makeup (HPI/NMU) Line Weld Failure

Bret L. Boman Framatome Technologies Lynchburg, VA

J. Michael Davis Duke Energy Charlotte, NC

James R. Smotrel Framatome Technologies Lynchburg, VA

Timothy D. Brown Duke Energy Seneca, SC

ABSTRACT An evaluation of the 1997 Oconee Nuclear Station Unit 2 High Pressure Injection/Normal Makeup (HPI/NMU) line weld failure is described herein. Computational Fluid Dynamics (CFD) analyses formed an integral part of the root cause investigation. CFD results were used to test each of the postulated failure causes. Ultimately, the failure was determined to be the result of a loose thermal sleeve that allowed cyclic thermal mixing between cold makeup flow and hot reactor coolant system (RCS) water to occur. This cyclic mixing resulted in thermal fatigue that caused cracks to initiate and propagate. INTRODUCTION On April 21, 1997, the Oconee Nuclear Station - Unit 2 experienced a through-wall crack in the weld between the High Pressure Injection/Normal Makeup (HPI/NMU) piping and the nozzle safe-end, Figure 1. The crack caused a leak in the 2A1 line. Plant operators correctly diagnosed the leak and brought the plant to safe shutdown. Evaluations, repairs, and corrective actions were conducted to understand the failure and preclude further occurrences. SYSTEM/COMPONENT DESCRIPTION Each of the four RCS cold legs is equipped with highpressure injection piping for use during an emergency condition such as a loss-of-coolant accident for the reactor. Two of these 2- inch schedule 160 (66.6 mm ID) stainless steel piping lines provide the normal makeup flow to the RCS to replace the water being removed and purified via the letdown piping. The nozzle is comprised of two parts, the base nozzle and a safeend, Figure 1.

To prevent thermal cycling of the base metal, each nozzle is equipped with a 1.5-inch (38.1 mm) ID thermal sleeve. Whereas the thermal sleeve experiences thermal cycling as the makeup flow varies to meet the demands of the RCS pressurizer level control, it does not experience pressure loading and had been perceived to be immune from significant thermal fatigue. Two thermal sleeve designs have been employed in most B&W-designed plants, Figure 2. The old design was installed without prescriptive guidelines for maintaining a tight fit between the sleeve and nozzle. The upstream end of the thermal sleeve (away from the RCS) was contact-expanded into the nozzle safe-end. As a result of flow-induced vibration, thermal expansion/contraction, or other phenomena many of the old design thermal sleeves loosened. The new design thermal sleeves have been hard-rolled into the nozzle with a specified wall thinning to provide good contact force. To date, the new design thermal sleeves have not experienced any loosening. Unique features of the Oconee units system designs (relative to other B&W units) are that: (1) normal makeup is fed through two cross-tied makeup lines, (rather than one line in other B&W units) and (2) each makeup line has a warming or minimum flow line (WL) to ensure continuous, positive makeup flow. (Other B&W units employ a bypass valve around the makeup control valve to ensure a continuous minimum flow.) The weld failure, safe-end cracking, and thermal sleeve cracking are not unique to the Oconee units. HPI/NMU nozzles have experienced similar failures at other B&W units.

14-3

SYSTEM OPERATION Under full power conditions, the cold leg water temperature is approximately 557F (292C). Normal makeup flow is typically 17 gpm (1.1 l/s) but can vary between 3 and 35 gpm (0.2 and 2.2 l/s). Makeup flow is typically 90F (33C). Under emergency conditions, high pressure injection flow is usually in excess of 100 gpm (6.3 l/s) per nozzle and temperatures can be as low as 40F (5C). INSPECTION RESULTS After safe shutdown of the plant, the nozzle safe-end, thermal sleeve, and portions of the makeup piping were replaced. The removed components were examined, visually and via metallurgical evaluation. The weld crack was observed to have initiated at the inside surface. Its circumferential extent was 360 degrees on the inside surface and 77 degrees on the outside surface, Figure 3. It is believed that the cracking had propagated over a long time period. Other components in the system also experienced damage. The region of contact expansion between the sleeve and the safe-end was found to have loosened to the point of producing an annular gap of approximately 0.03 inches (1 mm.) The thermal sleeve was also severely damaged. Numerous deep cracks were observed and a window of missing thermal sleeve was identified. There was significant wear of the thermal sleeve on the downstream (RCS) side. Within the safe-end, both outboard of and beneath the thermal sleeve contactexpanded area , severe cracking was observed, with some cracks extending nearly 30 percent through wall. In addition, the makeup piping in the vicinity of the warming line was found to have shallow surface cracks or crazing. TEMPERATURE MEASUREMENTS Because of concerns relative to thermal fatigue due to unanticipated thermal stratification in RCS attached piping systems (USNRC Bulletin 88-08,Reference 2) temperature measurements were made on the Oconee Unit 1 HPI/NMU lines in 1989-1990. The measured data showed no evidence of significant thermal stratification except during plant heatups and cooldowns. During some combinations of reactor coolant pump (RCP) operation, hot flow from one cold leg to another cold leg would occur via the cross-tied makeup piping. A maximum top-to-bottom pipe temperature difference of 327F (181.7C) was recorded in the makeup line adjacent to the idle RCP. This cross-flow occurred due to: (1) the pressure gradient between cold legs during certain partial RCP operation, and (2) reverse flow through leaking, oversized check valves. Since the number of thermal cycles occurring with these thermal stratified flow conditions was relatively small, this was not perceived, at that time, to be a significant thermal fatigue damage mechanism.

Post-failure (1997 to present) temperature measurements (at the weld and on the piping on both sides of the replaced, properly-sized check valves) have not shown this cross-flow induced thermal stratification since the replacement of the check valves. In fact, current measurements do not show any significant thermal activity. Maximum line temperatures and top-to-bottom temperature differences have been less than 190F (87.8C) and 20F (11.1C), respectively. However, due to physical constraints, temperature monitoring cannot reveal thermal stratification and cycling within the thermal sleeve itself. POSTULATED CAUSES OF FAILURE After review of the inspection results, temperature measurements, and other B&W plant operating experiences, the following potential causes of the weld crack and other observed thermal fatigue damage were postulated. 1. Partial RCP operation during plant heat-up/cool-down with resulting back-flow through the HPI/NMU boundary check valve of the non-operating RCP cold leg and into the HPI/NMU nozzle of the operating RCP cold leg in the same loop: It was thought that this cross-flow could have caused hot RCS water to flow backward from the idle cold leg, into that makeup nozzle, stratify and mix with that makeup lines cold warming line flow, and exit the other makeup nozzle, after also stratifying and mixing with the second makeup lines warming line flow. Turbulent penetration during full power operation: It was thought that this mechanism could have caused thermal stratification and cycling in the thermal sleeve and at the weld. In-leakage of hot RCS water through a loose thermal sleeve gap between the thermal sleeve and nozzle ID: It was thought that this annular back-flow could have caused hot RCS water to cyclically mix with the cold makeup flow in the vicinity of the weld.

2.

3.

Cycling mechanisms for the latter two failure causes included makeup flow variations, system pressure and flow variations, and the basic unsteady nature of the turbulent flow regime. CFD ANALYSIS Because CFD could potentially provide greater insight into the failure than could be achieved strictly through post-failure temperature monitoring, three series of CFD analyses, corresponding to each of the postulated failure causes, were conducted. CFD would provide a continuous temperature distribution (especially in areas that could not be measured

14-4

(e.g., within the thermal sleeve)) and would allow parametric evaluations to be performed. A three-dimensional finite element model of the piping and nozzle was constructed and the CFD code ANSYS FLOTRAN5.3, Reference 3, was used to calculate the flow patterns and temperature distributions in the nozzle and thermal sleeve region. The model included of 120,000 finite element nodes and used a thermal, turbulent, incompressible solution of the Navier-Stokes and energy equations. The model consisted of a portion of the cold leg, thermal sleeve, warming line, and makeup line. The standard K- turbulence equations were used with the Viollet buoyancy terms activated, Reference 4. The steady state results (typically 500 iterations) continued to show some cyclic variation with additional iterations, suggesting unsteady flow. CFD RESULTS The CFD analysis results consisted mainly of water temperature distributions within the thermal sleeve, nozzle, and makeup piping. Series 1 Results - The cross-flow cases modeled the jet pump suction that acts on the end of the thermal sleeve in the cold leg with the operating RCP. This resulted in a redistribution of the flow rate in the warming lines and the potential for back-flow through a leaking HPI/NMU check valve from the loop with the non-operating RCP to the HPI/NMU line of the operating RCP. In the HPI/NMU line with the idle RCP, the leaking check valve will result in RCS water being drawn into the thermal sleeve and HPI/NMU line due to a venturi like effect1. At, and upstream of the warming line interface, significant thermal gradients can exist as the hot RCS water flows over the top of the colder warming line flow, which enters the bottom of the HPI/NMU line. This can lead to substantial thermal stratification in the HPI/NMU line and high thermal gradients in the vicinity of the warming line connection, Figure 4. For the HPI/NMU piping with the operating RCP, the hot flow combines with the colder warming flow resulting in substantial thermal gradients also in the vicinity of the opposite warming line connection, Figure 5. Thermocouple data recorded in 1989 and 1990 support this conclusion.

Series 2 Results - The turbulent penetration cases investigated the thermal hydraulic conditions for a range of warming line and normal makeup line flow rates. This series of cases evaluated whether turbulent penetration during full power operation alone could result in thermal conditions at the safe end HPI/NMU piping weld that could lead to the Unit 2 failure. The results of those analyses showed that turbulent penetration effects produce high thermal gradients within the cantilever portion of the thermal sleeves at normal HPI/NMU flow rates. Normal HPI/NMU flow per nozzle is approximately 17 gpm (1.1 l/s), but can vary between 3 and 35 gpm (0.2 to 2.2 l/s). Only at very low flow rates (< 1 gpm) were significant thermal gradients calculated in the safe end-HPI/NMU line weld region upstream of the thermal sleeve, Figures 6-8. This indicates that during full power operations (with a tight thermal sleeve), the safe end piping weld is unaffected by turbulent penetration, since it is very unlikely that HPI/NMU flow would be less than 1 gpm. An additional turbulent penetration case was developed to determine the effect of RCS geometry on the length and orientation of the turbulent penetration. At the attachment point of the HPI nozzle, the RCS cold leg is inclined downward at 45 degrees to the horizontal. It was thought that this geometry might explain the temperature distribution in the sleeve. Turbulent penetration cases conducted within this investigation showed the tendency of the hot water to enter at the bottom of the thermal sleeve (downstream side), and realign to the top portion of the sleeve as the penetration moved further inward. To observe the effect of gravity, an additional case was modeled in which gravity was realigned to be exactly opposite the direction of cold leg flow. Results of this case, Figure 9, demonstrated that hot RCS water again entered the downstream side, which was now at the top of the thermal sleeve but did not have to reorient itself as with the previous cases. The distribution of the flow and the length of the penetration, in this case, compared favorably with previous work by EPRI in the Thermal Stratification, Cycling, and Striping program, Reference 5. Series 3 Results - The annular back-flow cases were analyzed similarly to the turbulent penetration cases, except that an additional source of upstream RCS flow was modeled through the gap between the thermal sleeve and safe end. The results showed that if the annular back-flow is of the same magnitude as the HPI/NMU flow rate, some hot water could propagate upstream (a separate hydraulics analysis calculated leakage rates of 1 to 12 gpm (0.06 to 0.8 l/s) depending on the thermal sleeve gap height and eccentricity). This region between the warming line and the thermal sleeve is already very turbulent due to the merging flows from the HPI/NMU line and the warming line. Annular back-flow would increase the turbulence and further disrupt the flow field. Significant time-dependent fluctuations of the thermal gradients would be

The pressure difference between the operating pump and idle pump loop is comprised of a static pressure difference and a velocity pressure difference. The static pressure in the operating pump loop is greater than the static pressure in the idle pump loop. However, due to the large differences in cold leg velocity between the operating pump loop and the idle pump loop, the aspirating flow effect due to the velocity pressure difference overwhelms the static pressure difference. This causes the makeup flow to be drawn from the idle pump loop to the operating pump loop.

14-5

expected in this region, which contains the safe end to HPI/NMU pipe weld, Figure 10. These results are especially relevant since the pipe upstream of the safe end exhibited a broad region of multi-directional cracks in the piping base metal. CORROBORATION OF RESULTS FLOTRANs ability to predict thermally stratified flows was demonstrated by benchmarking to test data, Reference 6. In addition, following a similar weld failure at the Crystal River Unit 3 plant in 1982, weld temperatures were measured as a function of makeup flow. Results showed that even at the lowest makeup flow rate of 1.5 gpm, no stratified flow was observed. This is consistent with the CFD results that showed turbulent penetration does not reach the weld until the makeup flow decreases below 1 gpm. Also, in all cases where weld and safe-end cracking has occurred, a loose thermal sleeve has been present. Thus, the conclusion that the loose thermal sleeve created conditions for high cycle thermal fatigue is supported by plant experiences and the CFD results. CONCLUSIONS AND RECOMMENDATIONS The CFD analyses provided insight into the failures that would have been difficult, if not impossible, to achieve with post-failure measurements. The CFD analyses confirmed that no single mechanism appears to be clearly responsible for all of the conditions observed in the Unit 2 nozzle, thermal sleeve, and piping. Since turbulent penetration during normal power operations caused the largest thermal gradients in the thermal sleeve, this is likely the mechanism responsible for the damage noted to the thermal sleeve. However, the CFD analysis demonstrates that turbulent penetration does not lead to significant thermal gradients at the safe end to HPI/NMU line weld except at uncharacteristically low HPI/NMU flows. Annular back-flow through a loose thermal sleeve can produce significant thermal gradients in the highly turbulent flow region at the safe end to HPI/NMU line weld and is likely the primary mechanism leading to the weld crack. However, annular back flow would not typically produce large thermal gradients in the thermal sleeve or HPI/NMU piping upstream of the warming line. During unbalanced partial RCP operation, cross flow can result in substantial thermal gradients in the vicinity of both warming line connections, and appears to be the most likely candidate responsible for the noted degradation at those locations. To preclude subsequent failures, it is recommended that the thermal sleeves undergo regular inspection to ensure the integrity of the thermal sleeve/nozzle joint (i.e., no

continuous gap that would allow hot RCS water to mix with the cold makeup flow). It is noteworthy that none of the redesigned thermal sleeves have exhibited the loosening experienced by the old thermal sleeve design. In addition, the thermal sleeve should either be treated as a consumable, have an inspection program implemented to insure its integrity, or be provided with makeup flow rates sufficient to preclude thermal cycling within the sleeve. REFERENCES (1) Shah, V.N., et al, Assessment of Pressurized Water Reactor Primary System Leaks, NUREG/CR-6582, December 1998. USNRC Bulletin 88-08, Thermal Stresses in Piping Connected to Reactor Coolant Systems. FLOTRAN Users Guide, Revision 5.1, DN:S261:51 1st Revision, Sept. 30, 1994, ANSYS, Inc., Houston, PA. (This Users Guide is also applicable to version 5.3) Viollet, P.L., "Modelling of Turbulent Recirculation Flows," Nuclear Engineering and Design (1987), pp. 365-377. Electric Power Research Institute (EPRI) Report TR103581, Thermal Stratification, Cycling, and Striping (TASCS), March 1994. Smotrel J.R., "Turbulent Thermal Stratification In A Long Horizontal Pipe," Proceedings of Seventh International ANSYS Conference, May, 1996.

(2)

(3)

(4)

(5)

(6)

14-6

14-7

14-8

14-9

15
CURRENT ACTIVITIES ON GUIDELINES OF HIGHCYCLE THERMAL FATIGUE IN JAPAN
Takao Nakamura The Kansai Electric Power Co.,Inc.

15-1

15-3

Current Activities on Guidelines of High-Cycle Thermal Fatigue in Japan


International Conference on Fatigue of Reactor Components 31 July-2 August,2000 Silverado Country Club & Conference Center Napa,California

Takao Nakamura
Kansai Electric Power Co.,Inc.
TEL : +81-3-3591-9261 (ex.256) FAX : +81-3-3593-0586 E-mail : K541422@kepco.co.jp

The

15-4

Background (1)
Voluntary countermeasures have been made based on troubles in domestic and foreign countries
Thermal stratification in Pressurizer Surge line and SG Feedwater line Thermal stratification due to Valve leakage Thermal stratification due to turbulence penetration to the stagnant line, etc

Leakage due to thermal fatigue occurred in spite of the efforts


- Mihama-2 excess letdown line in 4/1999 - Tsuruga-2 connecting pipe of regenerative heat exchanger in 7/1999

MITI required utilities to reinforce countermeasures

15-5

Background (2)
Additional requirement in MITI ordinance No.62
New requirement ; Structures and components should be designed to prevent the failure due to high cycle thermal fatigue.

Detailed actions to satisfy the new requirement described in the MITI guideline
Guideline established on the basis of current utilities practice Voluntary countermeasures turned to regulatory guideline

More investigation is necessary for phenomena such as Civau-1 case and turbulence penetration New group organized in JSME to establish voluntary design standards to be quoted to MITI ordinance

15-6

High-cycle Thermal Fatigue Phenomena


Phenomena Cause Turbulence penetration Leakage from valve seat
Example of Experienced Plant

Mihama-2 in 1999 Farley-2 in 1987 Genkai-1 in 1988 Trojan in 1988 D.C.Cook-2 in 1979

Thermal Leakage from valve grand stratification


Surge Operation Feedwater flow

Thermal striping

Mixing of hot and cold water Civaux-2 in 1998

15-7

JSME Activities
Utilities - vendors joint research for PWR and BWR
2 year program starting from April 2000 Includes hydraulic tests and analytical investigation Detailed program being developed

Scope (Phenomena to be considered)


Thermal fatigue due to temperature oscillation in a mixing zone of high and low temperature water Thermal fatigue due to oscillation of thermal stratification caused by turbulence penetration flow to the stagnant pipe

Establishment of design standards by utilizing data from the joint research and actual plants

15-8

Schedule for Establishment of High-cycle Thermal Fatigue Evaluation Standards


2000 2001 2002

1. Thermal Striping
(1) Investigation of past studies (2) Investigation of types of confluence in actual plants
(Investigation of points to be evaluation)

(3) Planning Test , Fabrication of test loop and device (4) Thermal striping test
-Measurement of temperature fluctuation

(5) -Thermal hydraulic analysis (6) Development of evaluation method

2. Thermal Stratification in stagnant pipe


(1) Investigation of past studies (2)
Investigation of configulations of stagnant part and operating conditions in actual plants (Investigation of points to be evaluation)

(3) Planning Test , (4)

Fabrication of test loop and device

Thermal stratification test in stagnant pipe


-Observation of themal stratification by the test -Study correlation between major flow parameters and the penetration depth

(5) Development of evaluation method

3.

Development of high-cycle thermal fatigue evaluation standard

Guideline for Evaluation of High-cycle Thermal Fatigue R&D Efforts


Investigation of past studies Investigation of points to be evaluation (configurations, operating conditions)

15-9

Clarification of factors to cause temperature fluctuation Study of the scope and structure for standard Thermal striping test
Measurement of temperature fluctuation Test plan Thermal hydraulic analysis

Thermal stratification test in stagnant pipe


Test plan Measurement of thermal stratification by the test Study correlation between major flow parameters and the penetration depth

Study correlation between major flow parameters and characteristics of temperature fluctuation Develop evaluation method of the distribution of temperature and stress Develop evaluation charts

Develop evaluation method Preliminary evaluation of actual components Results of actual plant component inspection for validation

Develop evaluation method Preliminary evaluation of actual components

Development of the draft high-cycle thermal fatigue evaluation guideline

15-10

Thermal Striping Test (1) Investigation of Types of Confluence in Actual Plants BWR PWR -Confluence type -Pipe Diameter (D1,D2) -Flow Rate (U1,U2) -Temperature (T1,T2) -Upstream Turbulence (valve, elbow, diffuser)
(D1,U1,T1)

Basic 3 types
Same Diameter, Collision
(D1,U2,T2)

(D1,U2,T2)

Same Diameter, Confluence


(D1,U1,T1) (D2,U2,T2)

Different Diameter, Confluence


(D1,U1,T1)

15-11

Thermal Striping Test (2) Test Plan


Visual Test Pipe mat. Acrylic resin Fluid Temperature Fluctuation Test Metal (SUS)
(D1,U2,T2) (D2,U2,T2)

model
(D1,U1,T1) (D1,U2,T2) (D1,U1,T1) (D1,U1,T1)

Type A

Type B

Type C

Test Parameter

Flow Rate Ratio: U2/U1 , Main Stream Flow Rate : U1 Diameter Ratio: D2/D1(Type-C) , Upstream Turbulence (Valve,Elbow,Diffuser) , etc. (Constant) Temperature Difference T and Main stream pipe Diameter D1

Measurement

Fluid Temperature ,

Metal Temperature
Amplitude and Period of Fluid Temperature Fluctuation Amplitude and Period of Pipe Wall Temperature Fluctuation

Evaluation

Confirmation of Flow Condition, Fluid Temperature Fluctuation Profile

15-12

Draft Guideline for Evaluation of Thermal Stratification in Stagnant Pipe


Design Condition Length to Horizontal Part : L/d, Branch Pipe Diameter : d, Main Pipe Diameter : D Main Flow rate : V, Main Flow Temperature : T, Horizontal Part Length : l, etc.

Determination of L1
(Parameter) V, d/D, T, (density), (kinematic viscosity), etc. Major pipe

Tcr [by calculating] is maximum temperature that stress amplitude by thermal stratification is less than fatigue limit in whole pipe region.

T Tcr ?
Yes No

T Tcr

lim

Determination of L2
(Parameter) V, d/D, T, l , etc.
L2 is maximum length when penetration flow enough reach into horizontal part .

L/d L1/d ?
Yes No

L/d

L1: Penetration depth of cavity-flow


Yes

L/d L2/d ?
No

L/d [ L/d L2/d ]

OK

Individual Detail Evaluation

[ L/d

L2/d ]

15-13

Thermal Stratification Test in Stagnant Pipe 1. Investigation of Branch types and its Flow Conditions in Actual Plants 2. Planning Test Matrix
Parameter: V, T, D, d, L, l (horizontal part length) , Branch type, Upstream elbow, , ,etc.

3. Thermal Stratification Test


(1) Visual Test (acrylic pipe)
Case1 : main pipe temperature T1=branch pipe temperature T2 =RT Case2 : T2-T1= about 40

(2)Actual Temperature and Pressure Test (SUS pipe)

Evaluation of L1 and L2

15-14

Summary
- More investigation is necessary to evaluate high-cycle thermal fatigue quantitatively - Data (from researches including tests and calculations ) will be collected by the Japanese joint research in 2 years - JSME design standard will be established using collected Data - International information exchanging is highly expected

Draft Guideline for Evaluation of Thermal striping

Design condition : Diameter ratio D2/D1 Velocity ratio U2/U1 Yes Tin < Tcr ? No

Temperature difference tin Upstream elbow/valve Tcr : Critical temperature difference Tf = Tin AF

15-15

Consideration to attenuation by mixing Tf mod = Tf Yes

Chart I : AF (attenuation factor)

Mod. Factor for upstream elbow/valve Tf mod < Tcr ? No

Heat transfer coefficient h in steady flow Consideration to heat transfer enhancement in unsteady condition Chart II : EF (enhancement factor) Structural calculation assuming most effective frequency Chart III : *

alt = 1/2 Kt * ETf mod / (1-)


Yes

alt < lim ? No

(lim : Fatigue limit)

Consideration to spectrum of Tf using Chart IV : Power spectrum of Tf and Chart III Structural calculation considering power spectrum of Tf i : i-th frequency component exceeding fatigue limit

f i operation time UF = i allowable repetition for i


UF < 1 ? Yes No

Redesign

Acceptable design

Draft Guideline for Evaluation of Thermal striping


Chart I : Attenuation of fluid temperature fluctuation by mixing
1.0 AF (attenuation factor) Enhancement factor U2/U1 = 10 1 0.5 0.1 3 U2/U1 = 10 2 U2/U1 = 1 U2/U1 = 0.1

15-16

Chart II : Heat transfer enhancement in unsteady condition

0.0

5 10 Nondimensional axial distance

1 Reynolds number

Chart III : Thermal stress assuming most effective frequency


1.0 * Bi = 10 5 0.5 1

Chart IV : Power spectrum density of temperature fluctuation


Power spectrum density of temperature fluctuation

0.0

0.1

1.0

f*

10.0

100.0

.001

.1 Strouhal number

10.

FATIGUE MONITORING/EVALUATION

16
STATUS AND UPDATE OF THE EPRI FATIGUEPRO FATIGUE MONITORING PROGRAM
Stan T. Rosinski EPRI 1300 Harris Boulevard Charlotte, North Carolina 28262 Gary L. Stevens Arthur F. Deardorff Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557

16-1

STATUS AND UPDATE OF THE EPRI FATIGUEPRO FATIGUE MONITORING PROGRAM Stan T. Rosinski EPRI 1300 Harris Boulevard Charlotte, North Carolina 28262 Gary L. Stevens Arthur F. Deardorff Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557

ABSTRACT FatiguePro has been developed as an advanced tool for tracking cycles and fatigue usage at key reactor coolant pressure boundary components. Using data from the plant process computer, FatiguePro can logically determine the types of cycles occurring and can compute stress- time histories for fatigue-sensitive locations. Initially developed in 1986, many new features have recently been incorporated. Continued user input is collected through the ongoing FatiguePro Users Group to enhance the software and to ensure that it maintains its usability with changing computer technology. This paper summarizes the capabilities of FatiguePro, how it is being used at operating nuclear plants and the continued utility support to keep the software up to date. INTRODUCTION Fatigue usage accumulation resulting from plant transient operation contributes to aging of critical equipment in light water reactor (LWR) nuclear power plants. During the design of modern nuclear plants, the effects of stress and fatigue were evaluated and bounded using design rules contained in Section III of the ASME Boiler and Pressure Vessel Code [1] and/or USAS B31.7 [2]. End-of-life cumulative usage factors were determined, in accordance with these rules, showing that fatigue usage factors were less than unity. This fatigue assessment was based on an assumed conservative set of design transients to assure that components would not exceed the allowable cumulative usage factor of one throughout their lifetime (usually forty years). To assure that design safety margins remain adequate throughout the operating life of the plant, each plant's license generally included requirements to count cycles or otherwise demonstrate that actual operating experience remains bounded by that assumed in the original plant design. Typically, significant design transient operating cycles are logged and counted to assure that the design fatigue limits are not exceeded. In practice, however, many of the actual plant operating cycles are not well characterized by the design transients and, in fact, are often much less severe. Classification of individual plant events into one of the design transient categories is also a

16-3

difficult task, for which plant operators are given relatively little guidance. Further, many of the events included in the design analysis had only a minor contribution to fatigue usage. As a result, some operating plants have approached the limit for the number of analyzed design transients early in plant life. In other cases, plant operating cycles may have been classified incorrectly or inconsistently, resulting in a poor estimate of cumulative usage accumulation. Finally, there have been occurrences of loadings not considered in the original design basis (e.g., thermal stratification), that have a significant contribution to usage factors when considered in a design analysis. In most cases, the design transients very conservatively bound plant operation. In other cases, transients have occurred that exceeded the design basis. In the early years of nuclear plant operation, there was no practical means by which plant operators could easily evaluate the effects of the difference between the transients actually occurring and those considered in the plant design. In 1985, EPRI initiated a project to develop a prototype system for monitoring cumulative usage factors in nuclear power plant components. In this project, a methodology was developed for accessing plant instrument data and converting this directly into peak stress versus time at plant locations of interest. The key to this technique was a transfer function approach, using Greens Functions and transfer matrices to convert plant data to peak stress versus time based on the actual plant data [3]. Another innovation was implementation of the ordered overall range (OOR) cycle counting methodology to account for actual cycle sequence, overcoming the limitation in the ASME Code which considers the most conservative possible cycle sequence. The methodology was developed into a specialized software system called FatiguePro, with demonstrations at two plants [4,5]. Through continued plant applications and improvements to the system, many additional features have been included to more fully address plant cycle tracking requirements. To date, the following features have been incorporated into the FatiguePro software: Acquisition of plant operating data (from existing plant instrumentation) Calculation of peak stress versus time and cumulative usage factor for fatigue-critical plant locations Automated cycle counting Computation of cumulative usage factors at fatigue-critical components using a cycle-based fatigue approach Prediction of growth of actual or hypothetical cracks Real time analysis and detailed review capabilities of all input plant data and results Summary reporting capability for permanent documentation purposes

This paper provides an update on the FatiguePro software and describes how it can be used to fulfill cyclic tracking and/or other margin assessment requirements to demonstrate that fatigue is being properly managed. FATIGUEPRO LIFE ASSESSMENT APPROACH Typically, nuclear plants are required to track plant transients against cycle limits. These requirements are specified in the plant licensing bases and/or Technical Specifications. The

16-4

intent of these requirements is to ensure that actual plant operation remains within the envelope assumed in the design basis. When properly implemented, these requirements are consistent with the recommended compliance of Generic Safety Issue 78 [6] and other regulatory documents. In many cases, plants have implemented manual or computer-based transient monitoring systems to fulfill plant-specific requirements for tracking cycle accumulation. The intent of the FatiguePro software development was to provide an industry-approved tool that could be used by plant engineers to fulfill plant fatigue life tracking requirements. In addition, the software provides the capability to perform a life assessment based on actual plant data. This life assessment may be based on any combination of the three following approaches: Counting, categorizing and tracking plant transient events, and comparing the result to the number of cycles assumed in the design basis. Computing cumulative usage factors, based on either real-time stress time history prediction or on a cycle-based approach, and demonstrating that usage factors less than unity are maintained for all monitored locations. Using a Section XI, Appendix L flaw tolerance approach [7] based on real-time plant data to demonstrate that actual or postulated flaws remain within acceptable values.

All of these approaches are intended to demonstrate that Code margins for fatigue-sensitive components are maintained during actual plant operation. This approach generally demonstrates significantly more margin than is shown in component stress reports, since actual transients are typically less severe than those assumed in the original design. The applicability of EPRIs fatigue monitoring methodology was demonstrated through implementation and field testing of a prototype system at San Onofre Unit 2, a pressurized water reactor (PWR) [4]. A similar installation and field test was also performed at Quad Cities Unit 2, a boiling water reactor (BWR) [5]. DEVELOPMENTS FatiguePro was developed and first applied at nuclear plants in the late 1980s. At that time, personal computers were much less powerful than today, and the Windows operating system was in its early stages of development. Even with this lack of computing power, the methodology built into FatiguePro was quite efficient at utilizing and logically evaluating plant instrument data to predict local temperatures and loading conditions at fatigue-sensitive locations. As computer power has increased, newer capabilities have been added to the FatiguePro software shell and into plant unique applications. Some of the significant advances include the following: Modern Windows Interface. The software interface now allows the user to perform an evaluation of plant data to determine if there are data errors or missing data that might affect the results. The plant data is archived in special compressed data files that may be reevaluated at a later time. The current fatigue status is available in a set of standard reports. Automated Cycle Counting. Whereas the original FatiguePro software predicted fatigue usage only based on predicted stress time history, the current version also has an automated

16-5

cycle counting module. The software logically evaluates actual plant data and counts designbasis cycles, such as heatup, cooldown, plant trips, etc. The key characteristics of each event are saved in the records, and/or are used to predict a fatigue usage contribution. Thus, plant data can be evaluated to show that actual transients are less severe and contribute to a lower fatigue accumulation than predicted in the original stress analysis. Piping Transient Temperature Response Modeling. In many cases, the determination of temperatures in a piping system must be based on remote instrumentation. In addition, when there is no flow, the piping system will slowly cool to ambient conditions or will equalize with the temperature of a connected system. A standard model has been implemented in FatiguePro that will properly account for these situations. This more-exact modeling of fluid systems allows for better prediction of local thermal stresses that are a major contributor to fatigue usage accumulation [8]. Environmental Effects. Although not yet implemented into the standard FatiguePro software, a special version of the software was developed to assess the effects of reactor water on fatigue usage of selected components. This required that local temperature effects and strain rate be evaluated in the fatigue analysis [9,10,11].

FATIGUEPRO USER'S GROUP (FPUG) Following the initial FatiguePro development in 1986 and subsequent evolution to FatiguePro version 2.0 in 1997, there was a desire on the part of many utility users to form a group that would direct future enhancements of the software. This group could also provide a forum for sharing, solving and responding to industry fatigue-related issues. In particular, a way to disseminate all of the lessons learned by the large population of users was desired. To respond to these desires, EPRI formed the FatiguePro User's Group (FPUG) in 1998. The FPUG has the following objectives: Provide methods for management of fatigue margins at member utility nuclear plants, including the use of EPRI's FatiguePro software. Develop updated FatiguePro software that is compatible with evolving computer infrastructure and operating environments. Enhancing member utility fatigue management capabilities through networking, exchanging of ideas and experiences, and presentation of topics in the area of fatigue.

Membership is composed of one utility representative for each subscribing plant site. The FPUG is initially chartered for four years and includes member-directed FatiguePro software development and enhancement. Software development and enhancement will include the implementation of three major architectural changes to FatiguePro, thereby modernizing the software, and allowing for more efficient code maintenance over subsequent years. These include: Implementation of a 32-bit design which will operate under current computer operating systems and use the improved interface features standard to these operating systems. Implementation of a full-featured database engine for storing configuration data and program results using standard SQL format.

16-6

Implementation of integrated data review and graphics modules such that reliance on other external software will no longer be needed to perform these functions.

The culmination of this effort will be the next generation of the FatiguePro software (Version 3.0). Semi-annual meetings linked with annual fatigue seminars are planned for the duration of the FPUG term to enhance member participation. The intent of the FatiguePro software development is to provide an industry-approved tool that can be used by plant engineers to fulfill plant cyclic duty tracking requirements by using any combination of approaches. All of the approaches used by FatiguePro are intended to demonstrate, in an accurate, reliable, and retrievable fashion, that structural design margins for all critical components are maintained during actual plant operation. Therefore, FatiguePro provides the technical tools that may be used to assure that structural design margins are maintained. TYPICAL USE OF SOFTWARE Reactor coolant system components in nuclear plants are designed and analyzed in accordance with ASME Section III Class 1 code design rules or similar rules for older plants. The design process involves analyzing each applicable component for a set of design transients that bound the expected plant operation in terms of temperature and pressure profiles for the life of the plant. Each component is analyzed to meet all of the applicable design requirements, including limits on primary and secondary stresses and cyclic duty limits. During the license period of the plant, the plant operator is responsible to ensure that plant components remain within the licensing basis. The plant Safety Analysis Report (SAR) usually specifies the set of design transients that initially define the plant licensing basis. Requirements for tracking plant operations stem from the necessity for the plant owner to demonstrate that the plant operates within the licensing basis. Excursions outside the boundaries of the licensing basis require special analysis to demonstrate that structural design margins are maintained over the operating life of the plant, thus ensuring continued safe and reliable operation. Early in the life of most present-day operating plants, the most direct and easiest form of ensuring that operation remained within the licensing basis was considered to be counting and categorization of plant transient events. Such an approach was considered straightforward, and eliminated the need to re-perform costly, labor-intensive fatigue evaluations based on plantunique operating history. Later in plant life, however, many plants may experience certain plant events in quantities that either exceed the number assumed in the licensing basis, or accumulate at a rate that is projected to exceed the number assumed in the licensing basis prior to the end of the desired operating period. In addition, events or loads may be experienced that were not considered in the original design. In these cases, simple cycle counting is not sufficient to demonstrate acceptable design margins. In order to demonstrate acceptable design margins in these instances, the plant owner has several options. First, each component can be re-analyzed to a revised set of design transients that represent the observed operation of the plant in terms of number and severity of events. Alternatively, the plant owner can initiate a condition

16-7

assessment program that accounts for the actual operation of the plant, and addresses the effect of actual operation on the structural margin of the affected components. These approaches are consistent with those recommended in ASME Code, Section XI, Nonmandatory Appendix L [7], as depicted in Figure 1. A few selected fatigue-critical components can be used to bound the fatigue behavior of entire components or systems by virtue of the most severe loadings and/or highest stress concentrations. Thus, a detailed fatigue analysis of these bounding components using actual plant transient temperatures and pressures substitutes for a complete design basis analysis of an entire component or system. The FatiguePro software is intended to provide an industry-approved tool that can be used within an integrated management program to show that design safety margins are maintained. FatiguePro can be used to fulfill plant cyclic duty tracking requirements, perform component structural margin evaluations, and accommodate flaw tolerance evaluations of components with cumulative usage factor values that exceed allowable limits. FatiguePro can be implemented (i.e., installed) at any time during the operating lifetime of a power plant. Through a "baselining" procedure that can be performed as a part of installation, monitored component fatigue duty can be determined from the start of plant operation up to the point of installation. This duty can be reflected in FatiguePro, and "on-line" monitoring can be performed from that point forward. Trending can be used to estimate the remaining life expectancy of the monitored components. To accomplish condition assessment, FatiguePro incorporates technical capabilities that address the following three requirements: Plant Licensing Basis Cycle Counting. This requirement is accomplished by consistently and accurately counting, categorizing, and tracking plant transient events for comparison to the events assumed in the licensing basis. This activity provides a direct measure that plant cycles remain within cyclic limit requirements, and an indirect measure of structural design margin. Automatically recording the occurrence of plant cycles may eliminate the need for plant operating personnel to do so manually, and reduces the inaccuracies and unnecessary conservatism inherent in manual cycle counting. Determine Actual Fatigue Margins. This determination is accomplished by computing cumulative usage factors based on actual plant operation, and demonstrating that these values remain less than the design allowable for all monitored components. This method goes one step further than cycle counting alone, in that it provides a direct measure of fatigue design margin. In addition, assessing the fatigue status of the reactor coolant pressure boundary can be readily accomplished subsequent to an event that exceeds the number of allowable cycles specified in the plant design basis. Demonstrate Actual or Postulated Flaws Remain within Allowable Limits. Components in the plant that are expected to accumulate abnormally high cumulative usage factors may be closely monitored utilizing a fatigue crack growth fracture mechanics methodology. Plant owners can adjust plant operational procedures and inservice inspection programs accordingly to ensure that design structural margins are maintained.

16-8

All of these approaches are intended to demonstrate, in an accurate, reliable, and retrievable fashion, that structural design margins for all fatigue-critical components are maintained during actual plant operation. FatiguePro provides the technical tools that may be used to assure that structural design margins are maintained in accordance with the approach appropriately chosen by the plant owners. CONCLUSIONS FatiguePro addresses the need for assuring plant design safety margins by providing the following capabilities: Automatically and reliably records the occurrence of plant thermal cycles. Determines actual fatigue margins. Assesses the structural integrity of the reactor coolant pressure boundary after an event that exceeds the operating pressure and temperature limits or number of allowable cycles, as identified in the plant Technical Specifications. Closely monitors areas of the plant that are expected to accumulate abnormal cumulative usage factors, enabling reactor operators to adjust plant operational procedures and inservice inspection programs accordingly.

Pilot plant studies have been performed using FatiguePro for lead BWR and PWR plants [4, 5, 12]. Each of these pilot projects included plant-specific application of the FatiguePro software, as well as detailed evaluation of several years worth of plant data. The objective of these evaluations was to rigorously test the FatiguePro methodology and provide further confidence in the plant-specific application of the software and its ability to fulfill plant cyclic duty tracking requirements. In addition, fatigue duty extrapolation schemes were developed for generic use. These methodologies estimate cumulative usage factors in instances where plant data are not available (i.e., as in the case of establishing the cumulative usage factor at a time when data are not retrievable). These studies concluded that fatigue monitoring for a few selected fatiguecritical components is a technically acceptable alternative to the cycle counting requirements contained in most plant technical specifications, and can be used to fulfill the related requirements associated with the plant licensing basis. Since the original development of FatiguePro, there have been continuing projects at individual plants to upgrade fatigue monitoring programs. Table 1 reflects the plants with FatiguePro installed or under development. All of the approaches used by FatiguePro are intended to demonstrate, in an accurate, reliable, and retrievable fashion, that structural design margins for all critical components are maintained during actual plant operation. Therefore, FatiguePro provides the technical tools that may be used to assure that structural design margins are maintained.

16-9

REFERENCES 1. ASME Boiler & Pressure Vessel Code, Section III, Rules for Construction of Nuclear Power Plant Components, Division I, Subsection NB, Class 1 Components, 1989 Edition (or earlier editions). 2. USAS B31.7, Nuclear Power Piping, 1969. 3. A. Y. Kuo, S. S. Tang, and P. C. Riccardella, "An On-line Monitoring System for Power Plants: Part 1 Direct Calculation of Transient Peak Stress Through Transfer Matrices and Green's Functions," Proceedings, 1986 Pressure Vessels and Piping Conference and Exhibition, PVP-Vol. 112, pp. 25-32, ASME, Chicago, IL, July 1986. 4. EPRI Report NP-5835, FatiguePro: On-Line Fatigue Usage Transient Monitoring System, May 1988. 5. EPRI Report NP-6170-M, FatiguePro On-Line Fatigue Monitoring System: Demonstration at the Quad Cities BWR, January 1989. 6. NUREG-0933, A Prioritization of Generic Safety Issues, U. S. Nuclear Regulatory Commission, April 1999. 7. ASME Boiler & Pressure Vessel Code, Section XI, Nonmandatory Appendix L, Operating Plant Fatigue Assessment, 1995 Edition. 8. Stevens, G., Gerber, D. and Rosinski, S., Latest Advances in Fatigue Monitoring Technology Using EPRI's FatiguePro Software, in Proceedings, 15th International Conference on Structural Mechanics in Reactor Technology, August 15-20, 1999, Seoul, Korea.. 9. EPRI Report TR-110356, Evaluation of Environmental Thermal Fatigue on Selected Components in a BWR Plant, EPRI, Palo Alto, California, April 1998. 10. EPRI Report TR-110043, "Evaluation of Environmental Fatigue Effects for a Westinghouse Nuclear Power Plant, EPRI, Palo Alto, California, April 1998. 11. EPRI Report TR-107515, Evaluation of Thermal Fatigue Effects on Systems on Systems Requiring Aging Management Review for License Renewal for the Calvert Cliffs Nuclear Power Plant, EPRI, Palo Alto, California, December 1997. 12. EPRI Report TR-107448, FatiguePro, Version 2: Fatigue Monitoring Software, December 1997.

16-10

Table 1. FatiguePro Installations


PWR FATIGUEPRO PROJECTS NSSS Fatigue Cycle Plant Vendor Analysis Counting ANO-2 C-E X X Calvert Cliffs 1/2 C-E X X San Onofre 2/3 C-E X Waterford 3 C-E X Ft. Calhoun C-E X Millstone 3 W X Maanshan 1/2 (Taiwan) W X X Diablo Canyon1/2 W X X Sequoyah 1/2 W X X Prairie Island 1/2 W X Point Beach 1/2 W X X Vogtle 1/2 W X X Callaway W X X Wolf Creek W X X McGuire 1/2 W X Catawba 1/2 W X Vandellos (Spain) W X X Surry 1/2 * W North Anna 1/2 * W Kewaunee * W ANO-1 B&W X Oconee 1/2/3 B&W X Crystal River 3 B&W X *recommendation report completed or in progress BWR FATIGUEPRO PROJECTS Fatigue Cycle Plant Analysis Counting Quad Cities 2 (Demonstration) X Oyster Creek X Susquehanna 1/2 X X Kuosheng 1/2 (Taiwan) X X WNP-2 X X Brunswick 1/2 X Chin-Shan 1/2 (Taiwan) X X River Bend X X Grand Gulf X X Toshiba Plants (Japan) X Santa Maria de Garoa (Spain) X Cofrentes (Spain) X FitzPatrick * Clinton X X *recommendation report completed or in progress Fatigue Crack Growth X

X X

Fatigue Crack Growth X

16-11

Typical Remaining Life Assessment

Model Component

Define Loading / Collect Operating Data

Inspection Data

Stress Analysis FE or Closed Form

Material Properties

Repeat for New Operating Data

Damage Model

Crack Growth Model

Damage
vs. Time

Crack Growth
vs. Time

Remaining Life Prediction


95093r0

Figure 1. Typical Remaining Life Assessment

16-12

Status and Update of the EPRI FatiguePro Fatigue Monitoring Program

Stan T. Rosinski EPRI Gary L. Stevens Arthur F. Deardorff Structural Integrity Associates
International Conference on Fatigue of Reactor Components 31 July - 2 August 2000 Napa, California
16-13

Introduction
q

q q

Fatigue accumulation is an aging concern for operating nuclear plants Set of design cyclic conditions specified for plant life Plant duty tracking requirements typically specified in Technical Specifications or FSAR Tracking is difficult
s

Actual transients are often different from (and less severe than) assumed transients Classification performed inconsistently by different engineers

q q
16-14

Tracking does not account for implicit margin In response to these issues, EPRI initiated FatiguePro project in 1985

Introduction
q

FatiguePro development initiated in 1985


s

Methodology developed for accessing plant instrument data and converting directly into peak stress versus time x Transfer function approach Implementation of ordered overall range (OOR) cycle counting methodology to account for actual cycle sequences Specialized software system developed and demonstrated at San Onofre (1988) and Quad Cities (1989)

16-15

FatiguePro Features

Features incorporated into FatiguePro


s s

s s s s

Acquisition of plant data Stress peaks/valleys computed for stress-based usage factors Automated cycle counting for cycle-based usage factors Prediction of growth of actual or hypothetical cracks Real time analysis and detailed review Comprehensive documentation

16-16

Fatigue Monitoring Using FatiguePro


q

When properly implemented a fatigue monitoring program complies with requirements specified in various regulatory documents
s s s s s

Provides knowledge of fatigue duty Demonstrates margins for continued operation Provides valuable insight to guide plant operation Identifies previously unknown loadings Addresses license renewal concerns

FatiguePro developed to provide integrated assessment as part of overall fatigue management


s

Assessment based on actual plant data

16-17

FatiguePro Life Assessment Approach


Typical Remaining Life Assessment

FatiguePro can be integrated into component life assessment strategies


s

Model Component

Consistent with ASME Section XI, Appendix L Assure design margins are maintained

Define Loading / Collect Operating Data

Inspection Data

Stress Analysis FE or Closed Form

Material Properties

Extensive industry utilization of FatiguePro as part of overall fatigue management program

Repeat for New Operating Data

Damage Model

Crack Growth Model

Damage
vs. Time

Crack Growth
vs. Time

Remaining Life Prediction


95093r0

16-18

PWR FatiguePro Installations


PWR FATIGUEPRO PROJECTS NSSS Fatigue Cycle Plant Vendor Analysis Counting ANO-2 C-E X X Calvert Cliffs 1/2 C-E X X San Onofre 2/3 C-E X Waterford 3 C-E X Ft. Calhoun C-E X Millstone 3 W X Maanshan 1/2 (Taiwan) W X X Diablo Canyon1/2 W X X Sequoyah 1/2 W X X Prairie Island 1/2 W X Point Beach 1/2 W X X Vogtle 1/2 W X X Callaway W X X Wolf Creek W X X McGuire 1/2 W X Catawba 1/2 W X Vandellos (Spain) W X X Surry 1/2 * W North Anna 1/2 * W Kewaunee * W ANO-1 B&W X Oconee 1/2/3 B&W X Crystal River 3 B&W X *recommendation report completed or in progress Fatigue Crack Growth X

X X

16-19

BWR FatiguePro Installations

BWR FATIGUEPRO PROJECTS Fatigue Cycle Plant Analysis Counting Quad Cities 2 (Demonstration) X Oyster Creek X Susquehanna 1/2 X X Kuosheng 1/2 (Taiwan) X X WNP-2 X X Brunswick 1/2 X Chin-Shan 1/2 (Taiwan) X X River Bend X X Grand Gulf X X Toshiba Plants (Japan) X Santa Maria de Garoa (Spain) X Cofrentes (Spain) X FitzPatrick * Clinton X X *recommendation report completed or in progress

Fatigue Crack Growth X

16-20

FatiguePro Update

Significant recent advances to FatiguePro


s

s
16-21

Incorporation of sophisticated thermal-hydraulic models x Eliminate unrealistic steps associated with design transients x Better prediction of local thermal stresses Incorporation of fracture mechanics solutions for fatigue crack growth assessments Development and implementation of environmental fatigue algorithms x Not yet included into standard FatiguePro x Prototype developed for scoping studies Formation of a FatiguePro Users Group

FatiguePro Users Group

FPUG formed in 1998 following development of FatiguePro version 2.0 (1997) FPUG Objectives
s

Provide methods for management of fatigue margins at member utility nuclear plants, including the use of EPRI's FatiguePro software. Develop updated FatiguePro software that is compatible with evolving computer infrastructure and operating environments. Enhancing member utility fatigue management capabilities through networking, exchanging of ideas and experiences, and presentation of topics in the area of fatigue

16-22

FatiguePro Users Group

FPUG membership consists of 23 domestic/3 international plant sites FPUG directing enhancement of FatiguePro
s s

Implementation of a 32-bit design Implementation of a full-featured database engine for storing configuration data and program results using standard SQL format Implementation of integrated data review and graphics modules such that reliance on other external software will no longer be needed to perform these functions Extensive user-interface enhancements

q
16-23

Next generation FatiguePro (version 3.0) to be released December 2001

Conclusions
q

FatiguePro addresses the need for assuring plant design safety margins by providing the following capabilities
s

s s

Automatically and reliably records the occurrence of plant thermal cycles Determines actual fatigue margins Assesses structural integrity of the reactor coolant pressure boundary after an event that exceeds the operating pressure and temperature limits or number of allowable cycles, as identified in the plant Technical Specifications. Closely monitors areas of the plant that are expected to accumulate abnormal cumulative usage factors, enabling reactor operators to adjust plant operational procedures and inservice inspection programs accordingly

16-24

Conclusions

Approaches in FatiguePro are intended to demonstrate, in an accurate, reliable, and retrievable fashion, that structural design margins for all critical components are maintained FatiguePro is an effective tool that may be used to assure that structural design margins are maintained

16-25

17
REMARKS TO THE DIFFERENT FACTORS INFLUENCING FATIGUE ANALYSIS AND FATIGUE DESIGN CURVES

E. Roos K.-H. Herter S. Issler Staatliche Materialpruefungsanstalt (MPA) University of Stuttgart Pfaffenwaldring 32 70569 Stuttgart, Germany

17-1

REMARKS TO THE DIFFERENT FACTORS INFLUENCING FATIGUE ANALYSIS AND FATIGUE DESIGN CURVES

E. Roos K.-H. Herter S. Issler Staatliche Materialpruefungsanstalt (MPA) University of Stuttgart Pfaffenwaldring 32 70569 Stuttgart, Germany

Abstract Technical codes used for construction, design and operation of nuclear components and systems provide detailed stress analysis procedures, materials data and a design philosophy which guarantees a reliable behaviour of the structural components throughout the lifetime. Especially for c yclic stress evaluation the different codes provides fatigue analyses to be performed considering various loading histories and geometric complexities of the components. In order to fully unde rstand the background of fatigue analysis included in the codes as well as the fatigue design curves used as a limiting criteria (fatigue life usage factor), it is important to understand the history and the methodologies which are available for the design engineers. Using design by analysis in the codes a simplified elastic plastic fatigue analysis is recommended when the range of primary plus secondary stress intensity exceeds the 3 S m limit. For that case the fictitious alternating stress amplitude S a is calculated by multiplying the range of primary plus secondary plus peak stress intensity S n with the stress dependent plastification factor K e . In nuclear and non nuclear codes different methods calculating K e values are available. The safety margins of this simplified elastic plastic fatigue analysis have been studied by using experimental and numerical results. The design fatigue curves in the nuclear codes are based on uniaxial cycling failure data. A factor of 2 on stress and a factor of 20 on cycles, whichever is more conservative at each point, was a pplied to the best fit curve. Nowadays numerous experimental data is available to check the conservatism of the fatigue design curves concerning the influence of welds, environment, surface finish, loading, temperature, mean stress and size. The effects of the different factors influencing the fatigue analysis and the fatigue design curves will be discussed.

17-3

Introduction

Technical codes and standards like ASME-Code Section III [1], French RCC-M Code [2], British Standard BS 5500 [3] or German KTA Safety Standards [4] are the basis for construction, design and operation of nuclear components and systems. The general philosophy in the design of components and structures is to demonstrate that the function and the is guaranteed throughout the lifetime. It is important that the design concept accounts for most possible failure modes and provides rational margins of safety against each type of failure. Some of the potential failure modes which component and structure designers should take into account are for example: Excessive elastic deformation including elastic instability, Excessive plastic deformation, Brittle fracture, Fatigue, Corrosion. During design stage a complete picture of the state of stress within the component and structure obtained by calculation or measurement of both mechanical and thermal stresses during transient and steady state operation has to be created. It has to be demonstrated that all stresses (primary, secondary) as well as environmental loading are within the allowable stress limits, and the usage factor developed by a fatigue analysis (peak stresses) is well below the limiting value. It is possible to prevent failure modes caused by fatigue by imposing distinct limits on the peak stresses at the highest loaded regions of the component and structure since fatigue failure is r elated to and initiated by high local stresses or by reducing the load cycles. The design rules a ccording to the technical codes and standards [1,2,3,4] provides for explicit consideration of cyclic operation, using design fatigue curves of allowable alternating loads (allowable stress or strain amplitudes) vs. number of loading cycles (S/N-curves), specific rules for assessing the cumulative fatigue damage (cumulative fatigue life usage factor) caused by different specified or monitored load cycles. The influence of different factors like welds, environment, surface finish, temperature, mean stress and size must be taken into consideration. 2 Use of Design Fatigue Curves

Reviewing fatigue rules and codes for nuclear pressure vessels and piping it becomes apparent that the majority are similar to or identical with those in the ASME-Code Section III [1], like the German KTA Standards [4]. The ASME design fatigue curves for carbon and low alloy steels as well as austenitic stainless steels are based on stress amplitude and cycles to failure data which were obtained from small smooth-machined specimens tested under strain control loading, mainly in bending in room temperature and air environment [5,6,7]. The design curves were derived by introducing factors of 2 on stress and 20 on cycles, whichever gave the lowest curve and is meant to account for real effects (size, environment, surface finish, scatter of data) occuring during plant operation. The fatigue design curve in the British Standard BS 5500 [3] was derived from fatigue 17-4

test data obtained under axial load from welded specimens. The reason therefore was that the presence of a weld could reduce fatigue strength because of the inevitable presence of weld d efects. But all of the pressure vessel and piping fatigue design rules are based essentially on the same approach based on data from primarily low-cycle fatigue (LCF) tests carried out on machined specimens, mainly with plain unwelded specimens tested under strain control. Conservative S/N-curves are developed and used for the fatigue analysis in conjunction with stress concentration factors K t or fatigue strength reduction factors K f to take into account the structural di scontinuities in the components and structures including welds [8]. Different procedures exist in the German technical rules for pressure vessels AD-Merkblatt [9] and the European Standard EN 13445 [10] for unfired pressure vessels. The approach uses also S/N-curves with stress concentration factors like the ASME Code but much more advice is given about the use of the stress concentration factors to be adopted for weld details. Additional explicit factors in form of an equation or a curve are given to account for the influence of temperature, surface finish and weldment, size and mean stresses. Further German codes and rules used in m echanical engineering and machinery are the FKM-Guidelines [11] and the RKF-Guidelines [12] with detailed requirements for the determination of alternating stress amplitudes. Fatigue data are generally obtained from unwelded specimens at room temperature and are plotted in the form of nominal stress amplitude S a vs. number of cycles to failure. The total strain range t obtained from the tests is converted to nominal stress range by multiplying the strain range by the room temperature modulus of elasticity E Sa = E t 2 (1)

Most of the S/N-curves given in the codes and standards are to be applied for specific steels (e.g. distinguish between steels of different ultimate tensile strength R m ). Influence of Temperature The use of fatigue design curves is restricted in the nuclear codes to a specific maximum te mperature below the creep range. Using design fatigue curves it is necessary to adjust the allowable stresses if the modulus of elasticity E at operating temperature is different from the one of the d esign curves. The stress amplitude S a must be multiplied by the ratio of the modulus of elasticity given by the design fatigue curve to the value of the modulus of elasticity used in the analysis. Another approach is given in the German AD S2 rules [9], where the influence of temperature must be adjusted by a cycle depending factor f T .

17-5

Influence of Surface Finish and Welds Design curves in the nuclear codes include a factor of 2 on stress or 20 on cycles relative to the mean of the test data to account for differences between specimen test conditions and real vessels and piping. This includes effects of surface finish and welds. Furthermore there are in the nuclear codes specific requirements concerning the surface finish of components especially for welded r egions and for different vessel and piping products and different joints. Stress indices are available for use of the code equations determining the stress amplitudes. A special regard to the influence of surface finish depending upon peak-to-valley height R z and number of cycles is given in German AD S2 rule. The influence of the surface finishing is d escribed by the surface factor which is defined by fo =

a , f (R z < 6 m)

a, f (R z )

(2)

where a , f is the sustainable stress amplitude for different R z values. The requirements for determining f o according to AD S2 for a material with ultimate tensile strength of 500 MPa are shown in Fig. 1.
1
R z <6 m

0,9 S urface fa ctor f o

6 10

0,8
50

0,7

AD S2 R m = 500 M P a

100 200

0,6 10 00

10 000

10 000 0 C ycles N

10 000 00

10 000 000

Figure 1: Influence of surface finish according to AD S2 [9] Experimental values for surface factors f o derived by specimens with different R z values are 6 shown in Fig. 2. The experimental data has been evaluated for the endurance limit (N=210 ) of different materials.

17-6

1
ste e ls w ith U T S 5 70 M P a < R m <1 3 00 M P a

0,9

0,8 Surfa ce factor f o

0,7

0,6 safe 0,5 0,5 0,6 0,7 S urface factor f o 0,8

expe rim e nts by M PA S tuttgart

0,9

Figure 2: Surface finish factor f o for the endurance limit of stress (fatigue strength) Influence of Size Most of the material and failure behaviour has been determined using small laboratory specimens. However failure stress amplitudes are lower for components because of size effects caused by different stress gradients or statistical effects of material characteristics. Size effects are covered in the nuclear codes [1,4] by the factors 2 on stress or 20 on cycles. A different approach is included in German AD S2 rule, Fig. 3
1
< 2 5

30

0,9
50
d f r

0,8
1 00

0,7

A D S2 non w e ld ed com po nents

w all th ickn ess s=1 5 0 m m

100 0

100 00

100 000 C ycles N

100 0000

100 00000

Figure 3: Influence of size according to AD S2 [9]

17-7

Experimental data concerning the influence of size are available from [13], Fig. 4. It is evident that comparatively large scatter emerge, especially for the tests with larger specimens.
1 Exp. C rN iM o ste el (Kt=1 ) Exp. C rN iM o ste el (Kt=2 ) Exp. C rN iM o ste el (Kt=5 ) AD S2

0 ,9 S ize factor f d

0 ,8

0 ,7

0 ,6 0 20 40 60 E q uivalen t d ia m eter / m m 80 1 00

Figure 4: Size factor f d for the fatigue strength of round solid specimens under repeated (r eversed) bending stresses [13] depending upon the equivalent diameter (for pipes: equivalent d iameter = s/2) Influence of Mean Stress If a component is stressed by an alternating stress greater than the yield strength R p0,2 of the material, it makes no difference whether there is a present nominal mean stress m or not. In this stress state the true mean stress always will be zero. Therefore the fatigue design curves are adjusted to include the maximum effect of the mean stress only in the part of the fatigue curve lying below an alternating stress amplitude a = R p0,2 . In all the ASME-based codes the fatigue design curves are plotted in terms of stress amplitude independent of mean stress, the curves are showing already the full effect of maximum mean stress. The evaluation of the effect of mean stresses is accomplished by use of the modified Langer-Goodman Diagram, where mean stress is plotted as the abscissa and the amplitude of the a lternating stresses is plotted as the ordinate. Thus, for the adjusted fatigue curve there should not be any mean stress present which will cause fatigue failure in less than the given cycles. In non-nuclear codes the influence of mean stresses is taken into account individually. A simple equation was proposed by Wellinger and Dietmann [14]

a , f (R ) = a , f (R = 1) 1

m
Rm

(3)

with R as the stress ratio. The influence of mean stresses in the FKM-Guidelines [11] is described by the mean stress sensitivity M which was introduced by Schtz [15] as 17-8

M =

a, f (R = 1) a , f (R = 0) . a , f (R = 0)

(4)

If there is no experimental data available, the mean stress sensitivity for steels can approximately be determined by the equation M = 0.00035 R m 0.1 , with R m in dimension MPa. Equation (5) is illustrated in Fig. 5.
1 ,2 M e a n s tre ss se n sitiv ity M 1 ,0 0 ,8 0 ,6 0 ,4 0 ,2 0 0 5 00 1 00 0 1 50 0 2 00 0 U pp e r ten sile stres s R m / M P a
A lM g5 A lM gS i1

(5)

sc atte r a lu m inium allo ys tita n iu m a lloy s


3.1354.5 3.1254.7 3.4354.7 3.4364.7

ca ste d ste e ls
G S N iC oM o N iC oM o

ste els
A M 355

C k 45 S A E 4130 G S 25 C rM o4

N iC oM o geglht 41 C r4

Figure 5: Mean stress sensitivity M versus ultimate tensile strength [15] For the endurance limit of stress (fatigue strength) the mean stress effect on the alternating stress amplitude a , f can be adjusted by the Haigh's diagram. Fig. 6 shows the proposal by [11] and [14] compared with experimental data for a high strength low alloy rotor CrMoV-steel.
a,f / M P a
R= 500 R= 1 arctan M R =0 arctan M 3

St 37

St 52

P H 15-7 M o

1.6604.5

1.7704.6

R =0,5

FK M -G uidelines [11] W ellinge r/D ietm ann [14] E xp. M PA Stuttgart -500

yield lim it -100 100 500 M ean stress m / M P a R p0,2 R m

Figure 6: Influence of mean stress according to the fatigue strength diagram (Haigh's diagram) 17-9

Influence of Environment Despite of the factors 2 and 20 there have been relatively few corrosion fatigue failures in carbon or low-alloy steel components in LWR's and quite a lot of discussions are under way concerning the influence of environment to the fatigue design curves (crack initiation and crack growth under environmental conditions). Data from specimens testing indicated that fatigue life shorter than the fatigue design curve values are possible, if the tests are carried out under low frequency loading conditions in oxygenated water environment at elevated temperatures [16,17]. The investigations performed to determine corrosion-assisted crack growth rates for pressure boundary materials exhibit a big scatter [18]. 3 Calculation of Stress Intensity Range

The four equations used in Class 1 piping to calculate the stress intensity are the code equation 9 addressing primary stress margins B1 pD o D M + B 2 o i 1,5 S m 2t 2I (primary stress limit for design conditions), (6)

the code equation 10 addressing the shake down stress limit p D D M S n = C1 o o + C 2 o i + Thermal Stress Range (Tm ) 3 S m 2t 2I (limit for the primary + secondary stress intensity range) and the code equation 11 defining the peak stress range for fatigue analysis p D D M S p = K1C1 o o + K 2 C 2 o + Thermal Stress Range (Tm , T1 , T2 ) 2t 2I with the stress amplitude Sa = Sp 2 (9) (8) (7)

Research is still under way concerning the categorization of the stresses directly influencing the result of a fatigue analysis. Thermal Stresses Most of the fatigue relevant stresses in piping systems are caused by thermal loading. The difference between the density of the fluid caused by the temperature gradient from bottom to top of the pipe cross section (eg. pressurizer coolant and that of the somewhat cooler hot leg coolant) combined with low flow rates can result in thermal stratification in the horizontal portions of a

17-10

piping system. The hot and cold fluid levels of the stratified flow conditions are separated by a interface or mixing layer. On the other hand high flow rates can cause a temperature gradient in pipe longitudinal direction (jump of temperature) and result in a thermal shock loading on the inside pipe surface constant throughout the pipe cross section. To calculate thermal stresses in pipes the code equations 10 and 11 are available. Thermal stratification Thermal stratification in piping system causes an cirumferentially varying temperature distribution in the pipe wall resulting in local through wall axial stresses (through wall radial temperature gradient) and global bending stresses in the piping system (axial expansion forces and thermal m oments). Maximum local thermal stress is found when a thin interface layer occurs in the upper or lower parts of the pipe cross section. Maximum global thermal bending stress is found when a thin interface layer occurs in the middle of the pipe cross section. The ASME-Code Section III [1] or KTA Standards [4] rules calculating thermal stresses may not be completely applicable for the thermal stratification loading. Thermal shock The ASME-Code Section III [1] or KTA Standards [4] rules calculating thermal stresses are a pplicable for the thermal shock loading. Plastification Factor Ke For nuclear power plant components which are a subjected to cyclic loading a fatigue analysis in accordance to different safety rules [1-3] has to be performed. If elastic-plastic deformation is to be expected then generally costly non-linear FE-calculations have to be carried out. However under specific conditions a simplified elastic-plastic fatigue analysis may also be performed using plastification factors K e . This is considerably simpler since it is based on linear elastic material behaviour. The determination of K e values according to the German KTA safety rules, which have been largely adopted from the ASME code is shown in the following. Characteristic stresses and strains are represented by a hysteresis loop during a cycle, Fig. 7. The fictive elastic strains el are brought into line with the actual, elastic-plastic strain t by multiplying them by the plastification factor K e which is defined by Ke = t . el (10)

According to KTA 3201.2, Section 7.8.4 the plastification factor has to be calculated as follows:

17-11

Ke =

1 1+ 1 n S 1 n n 1 n (m 1) 3 S m

S for 0 n 3 Sm S for 3 < n < 3m Sm for Sn 3m Sm (11)

with S m as design stress intensity value for the material used.

a m in

a m m ax

m m ax

a m in
pl t

el

Figure 7: Hysteresis loop For example, for ferritic respectively martensitic steels the characteristic material value S m is calculated as R p0,2T R mT R mRT S m = Min ; ; . 3,0 2,7 1,5 (12)

S n is the fictive elastic equivalent stress range of primary and secondary stresses. The value of the material parameters m and n are to be taken from Tab. 1. Table 1: Material parameters Material low alloy carbon steel, martensitic stainless steel carbon steel, austenitic unalloyed stainless steel nickel-based alloy Group 1 2 3 m 2,0 3,0 1,7 n 0,2 0,2 0,3 Tmax / C 370 370 425

17-12

In Fig. 8 the dependence of K e on the load proportional reference magnitude S n / S m is shown for various material groups.
6 5 P la stifica tio n fa cto r K e 4 3 2 1 0 0 2 4 R a tio S n / S m 6 8 10

K TA 320 1.2

gro up 1 gro up 2 gro up 3

Figure 8: K e factors according to ASME [1] and KTA [4] An evaluation of experimental results from LCF tests with smooth specimens (K t = 1) for group 1 materials (ferritic and martensitic steels) is shown in Fig. 9. Its obvious that the scatter is a ccording to the available wide scale of different heat treatments and strengths relatively small. The calculation of K e values according to ASME/KTA is evidently very conservative.
8
K TA E xp . E xp . E xp . E xp . 3 2 0 1.2 (g ro u p 1 ) X 2 0 C rM o V 1 2 1 15 M nNi 6 3 1 3 C rM o 4 4 15 M o 3 E xp . E xp . E xp . E xp . 20 15 10 14 M nM oNi 5 5 N iC u M o N b 5 C rM o 9 1 0 M oV 6 3

P las tifica tion fa cto r K e

LC F tes ts at ro om tem pe rature (n =1 41)

0 0 5 10 15 20 25 R a tio S n / S m

Figure 9: K e factors from LCF tests compared with the values of KTA [4]

17-13

Conclusions

Fatigue is a potential failure mode for components and structures of nuclear power plants. For the explicite consideration of cyclic operational loads (mechanical, thermal) in different technical codes and standards specific fatigue analysis methods are available. As for the nuclear codes and standards the following conclusions can be drawn: - The influence of temperature is adequate addressed by considering the ratio of the modulus of ela sticity using the S/N-curves. - The design curves were derived by introducing factors of 2 on stress and 20 on cycles to account for real effects (size, environment, surface finish, scatter of data) occuring during plant operation. This implies that during manufacturing and design the specific requriements are met and during plant operation the conditions for environmental effects are monitored and controlled. - To account for thermal stresse within a fatigue analysis the code equations may not be completely applicable for the thermal stratification loading - Compared to experimental data the calculation of the plastification factor K e according to ASME/KTA is evidently very conservative. References 1. "Rules for Construction of Nuclear Power Plant Components". ASME Boiler and Pressure Vessel Code, Section III, The American Society of Mechanical Engineers, 1998 Edition "French Design and Construction Rules for Mechanical Components of PWR Nuclear Islands (RCC-M)". AFCEN - Association Franaise pour la Construction des Ensembles Nucl aires, Paris "Unfired Fusion Weld Pressure Vessels - BS 5500". British Standard Institution "Safety Standards of the Nuclear Safety Standards Commission (KTA)". KTA Rules 3201 and 3211, Carl Heymanns Verlag KG, Cologne, latest edition B. F. Langer, "Design of Pressure Vessels for Low-Cycle Fatigue", Journal of Basic Engineering, Vol. 84, No. 3, September 1962, pp. 389-402 C. E. Jaske, W. J. O'Donnell: "Fatigue Design Criteria for Pressure Vessel Alloys", Journal of Pressure Vessel Technology, November 1977, pp. 584-592 D. R. Diercks: "Development of Fatigue Design Curves for Pressure Vessel Alloys using a Modified Langer Equation", Journal of Pressure Vessel Technology, Vol. 101, November 1979, pp. 292-298 "Fatigue strength reduction and stress concentration factors for welds in pressure vessels and piping", Welding Research Council Bulletin, WRC 432, June 1998 17-14

2.

3. 4.

5.

6.

7.

8.

9.

German Technical Rules for Pressure Vessels (AD-Merkbltter), AD-S1 and AD-S2, Carl Heymanns Verlag KG, Cologne, latest edition

10. European Standard for "Unfired Pressure Vessels" EN 13445-3 (Part 3 - Design), CEN, 1999 11. "Rechnerischer Festigkeitsnachweis fr Maschinenbauteile (FKM)", Teil III - Ermdungsfestigkeitsnachweis, Heft 183-2, Forschungskuratorium Maschinenbau e.V., Frankfurt 1994 12. "Richtlinienkatalog Festigkeitsberechnung Behlter und Apparate (RKF)", Teil 5 und 6 - Ermdungsfestigkeit, Linde GmbH, Dresden, 1986 13. K.-H. Kloos, B. Fuchsbauer, W. Magin, D. Zankov: "bertragbarkeit von ProbestabSchwingfestigkeitseigenschaften auf Bauteile", VDI-Berichte, Nr. 354, 1979, pp. 59-72 14. K. Wellinger, H. Dietmann: "Festigkeitsberechnung - Grundlagen und technische Anwendung", Alfred Kroener Verlag, 1976 15. W. Schtz: " ber eine Beziehung zwischen der Lebensdauer bei konstanter und vernderlichen Beanspruchungsamplituden und ihre Anwendbarkeit auf die Bemessung von Flugzeugbauteilen", Zeitschrift fr Flugwissenschaften, Band 15, 1967 16. J. M. Keisler, O. K. Chopra, W. J. Shack: "Statistical models for estimating strain-life behaviour of pressure boundary materials in light water reactor environment", Nuclear Engineering and Design 167 (1996), pp. 129-154 17. O. K. Chopra, W. J. Shack: "Low-cycle fatigue of piping and pressure vessel steels in LWR environments", Nuclear Engineering and Design 184 (1998), pp. 49-76 18. K. Kussmaul, D. Blind, V. Lpple: "New observations on the crack growth rate of low alloy nuclear grade ferritic steels under constant active load in oxygenated high-temperature water", Nuclear Engineering and Design 168 (1997), pp. 53-75

17-15

18
THE RUSSIAN REGULATORY APPROACHES IN THE FATIGUE EVALUATION OF NPP CONSTRUCTION COMPONENTS

I. Kaliberda N. Karpunin Gosatomnadzor, Moscow, Russia

No presentation slides or technical paper available.

18-1

The Russian Regulatory Approaches in the Fatigue Evaluation of NPP Construction Components

I. Kaliberda N. Karpunin Gosatomnadzor, Moscow, Russia

No presentation slides or technical paper available.

18-3

19
THE EVALUATION SYSTEM OF THERMAL STRATIFICATION STRESS USING OUTER SURFACE TEMPERATURE

Itaru Muroya Kiminobu Hojo Takasago Research & Development Center Mitsubishi Heavy Industries, Ltd. Sigeki Suzuki Toshimitsu Umakoshi Kobe Shipyard and Machinery Works Mitsubishi Heavy Industries, Ltd.

19-1

19-3

19-4

19-5

19-6

19-7

19-8

19-9

19-10

19-11

19-12

19-13

19-14

NDE/NDT

20
MICROSTRUCTURAL CHANGES OF PRESSURE VESSEL STEEL DURING FATIGUE IN HIGH TEMPERATURE WATER ENVIRONMENT

Chie Fukuoka, Yukiya G. Nakagawa, Makoto Higuchi Ishikawajima Harima Heavy Industries, Co. Ltd. Tokyo, Japan Stan T. Rosinski EPRI Charlotte, North Carolina, USA

20-1

MICROSTRUCTURAL CHANGES OF PRESSURE VESSEL STEEL DURING FATIGUE IN HIGH TEMPERATURE WATER ENVIRONMENT Chie Fukuoka, Yukiya G. Nakagawa, Makoto Higuchi Ishikawajima Harima Heavy Industries, Co. Ltd. Tokyo, Japan Stan T. Rosinski EPRI Charlotte, North Carolina, USA Abstract An effective method for measuring the fatigue damage accumulation state in a structural material was applied to samples removed from an A533 Grade B, Class 1 plate material fatigue cycled in a reactor water environment. The method, Selected Area Diffraction (SAD), is a microstructural examination technique that is used for identifying small cell-to-cell angular misorientations in the crystal lattice. Samples were fatigue cycled as a function of strain amplitude and strain rate in a reactor water environment and angular misorientation measurements taken utilizing the SAD technique. SAD measurements were then correlated with total fatigue damage accumulation. The angular misorientation was found to increase as the fatigue deformation increased, rapidly increasing in the early stages of cyclic loading followed by a more gradual increase during subsequent cycling. This profile is consistent with previous SAD measurements taken on samples fatigue tested in air. For samples fatigued to failure in a reactor water environment a similar value of angular misorientation was measured, independent of the testing strain rate. This suggests that measurement of a threshold angular misorientation value is feasible for component life assessment of materials exposed to reactor water environments. Introduction Reduction in forced outages and avoidance of expensive repairs would be facilitated by an inspection technique that could predict when fatigue macro crack initiation will occur in the metallic structures, systems, and components of power plants. As nuclear power plants age, evaluating pre-fatigue damage becomes increasingly important. A microstructural technique for measuring the early stages of fatigue in reactor vessel materials has been developed 1. The method, Selected Area Diffraction (SAD) is based on microstructural examination by electron diffraction for identifying dislocation cell to cell angular misorientation in the crystal lattice of quenched and tempered high strength low alloy steels. It was observed that this misorientation increases as the fatigue deformation accumulates, which is a prerequisite for fatigue crack initiation 2.

Y.G. Nakagawa, H. Yoshizawa, M.E. Lapides, Metall. Trans. A, 1990, vol.21A, pp.1769-73 C. Fukuoka, H. Yoshizawa, Y.G. Nakagawa, M.E. Lapides, Metall. Trans. A, 1993, vol.24A, pp.2209-16
2

20-3

It has been demonstrated that the fatigue lifetime of steels in a pressurized high temperature water environment is shorter than in air 3,4,5. These studies also showed that the fatigue life was strongly influenced by strain rate and dissolved oxygen (DO) content. A formula for the relationship between fatigue lifetime and strain rate has been suggested4: N25w = A(dT / dt)P high temperature water and dT / dt is the tensile strain rate. (1)

where N25w is defined as the fatigue life, measured as a 25% reduction in applied peak load, in

The objective of this work was to investigate the effect of reactor water environment on the microstructural changes during fatigue. The SAD was used to correlate the impact of the environment on the component fatigue life by measuring microstructural changes in the materials exposed to reactor water environment. Additional details of this work have already been published elsewhere6 Experimental Procedure Fatigue Testing in High Temperature Water Fatigue test specimens were machined from an A533 Grade B Class 1 (A533B-1) reactor pressure vessel plate material. This material had been previously prepared to specifically simulate Japanese early vintage reactor material and contains a higher sulfur content and generally exhibits a lower toughness. The chemical composition and mechanical properties of this material are summarized in Table 1. Figure 1 (a) and (b) provide dimensions of the fatigue test specimens that were utilized during testing in air and in LWR water environments, respectively. Fatigue damage was induced in the test specimens by axial strain controlled cycling at two different strain rates, 0.004%/s and 0.4%/s, and at different strain amplitudes. For fatigue testing at the slow strain rate (0.004%/s) a slow-fast saw-tooth shaped cycling pattern was utilized in order to reduce the overall fatigue testing time. The fast portion of the saw-tooth shape was utilized during the compressive portion of each fatigue cycle. It has been previously suggested that only the tensile strain portion of the loading had significant influence on the fatigue

M. Higuchi and H. Sakamoto, Low Cycle Fatigue Of Carbon Steel In High Temperature Pure Water Environment, Tetsu to Hagane (Journal of Iron and Steel Institute of Japan), vol.71, no.8, 1985, pp.101-107 4 M. Higuchi and K. Iida, Fatigue Strength Correction Factors For Carbon And Low-Alloy Steels In Oxygen-Containing High-Temperature Water, Nuclear Engineering and Design, vol.129, 1991, pp.293-306 5 M. Higuchi and K. Iida, Reduction In Low-Cycle Fatigue Life Of Austenitic Stainless Steels In High-Temperature Water, ASME PVP-vol.353, 1997, pp.79-85 6 Measuring Fatigue Damage in Materials Phase 2, EPRI, Palo Alto, CA: 1999, TR-110251. 20-4

lifetime7. Fatigue testing in air was conducted either at room temperature (RT) or at a nominal LWR operating temperature, 289C. Fatigue tests in simulated LWR water were performed in an autoclave with a closed water circulating loop. The DO level in the water was established by controlling the rate of bubbling argon gas and oxygen into deionized water. Table 2 summarizes the environmental conditions applied during testing. During fatigue testing both in air and simulated LWR water, sample failure (represented by the number of cycles to failure, N25) was taken to be the point beyond which the applied peak stress fell 25% under the same strain conditions. SAD Measurements The SAD method used in this study is schematically illustrated in Figure 2 and has been described elsewhere1. This method uses transmission electron microscopy to measure the average cell to cell misorientation in grains oriented around the <111> zone axis. An example of the statistical analysis performed on data obtained from the SAD technique is shown in Figure 3. The normal distribution of the maximum angular deviation, , is illustrated for an as-received SA508 sample and for samples previously fatigued at a total strain range of 0.78 % to N/N f = 50, 100 % (Nf =2,100). The fraction of life, N/Nf, expresses the state of fatigue damage, where N is the number of the cycles applied to the sample and N f is the number of cycles to failure. The mean value of , equivalent to 50% probability in each plot, is considered to represent an average angular deviation of the cells from the reference direction, <111>. Figure 4 illustrates typical SAD measurements performed on SA508 samples of different heat origins and fatigued at different strain ranges. As shown in Figure 4, the values at N/Nf = 100% are similar, 4 to 5 degrees, for all tests, suggesting that a critical orientation value may exist for this material. Those samples whose fatigue tests were interrupted for microstructural examination were first visually inspected for cracking. The gage section was examined with an optical microscope (x10 magnification). All test bars were cut perpendicular to the stress axis within the gage section. Small disks (3 mm diameter x 0.1 mm thickness) were fabricated and electropolished to prepare TEM and SAD samples. Microstructural damage was evaluated by the SAD method and correlated to the fatigue testing conditions. Fatigue lifetime changes due to the different types of testing conditions were correlated to the microstructural observations by the SAD method and surface observations.

Higuchi, M., Iida, K., and Asada, Y., Effect Of Strain Rate Change On Fatigue Life Of Carbon Steel In High-Temperature Water, ASTM STP 1298, American Society of Testing and Materials, 1997 20-5

Results The results of fatigue testing performed in air on the A533B material, including applicable angular misorientation values measured via the SAD technique, are provided in Table 3. Results of fatigue testing performed in a simulated LWR environment, including applicable angular misorientation values measured via the SAD technique, are provided in Table 4. Fatigue Test Results Strain amplitude versus fatigue lifetime curves of SA508-2 in air and A533B-1 in air and in water are plotted in Figure 5. Open and solid symbols represent fatigue test results in water and in air, respectively. A fatigue curve and its formula are also shown in this figure, which was obtained as the mean data curve for Japanese low alloy steels in room temperature air 8. Closed circles represent fatigue test results for SA508-2 in air, triangle markers represent testing results at normal strain rate (0.4%/s) for A533B-1, and square markers represent testing results at slow strain rate (0.004%/s) for A533B-1. It is evident that fatigue lifetime is significantly reduced in the LWR environment, especially at slow strain rate (0.004%/s). Figure 6 shows the relationship between strain rate and fatigue life. The slope of the curves shown in the figure is parameter P in equation (1). The P value (0.476 for strain amplitude of 0.6%, 0.747 for that of 0.31%, and 0.746 for that of 0.24%) measured by this study was much larger than previous observations for low alloy steels (reported P values for low alloy steels typically ranged from 0.31 to 0.44)9. This may be due to the relatively high sulfur content of the plate used in this experiment. It has been demonstrated that sulfur content of steels has a significant effect on fatigue crack growth rate10. Higuchi and Iida also proposed the P value of carbon and low alloy steels can be roughly estimated by a following empirical formula 8: PC = 0.864 + 14.6 * S (%) - 0.00092 * UTS (MPa) (2)

where PC is the P value at 289C and 1-8 ppm oxygen content. This formula implies that the strain rate more significantly impacts fatigue life with an increase in the sulfur content. The P C value of the A533B used for this experiment is 0.588. SAD and Microstructural Examination Results Figure 7 shows representative fracture surfaces of fatigue specimens from this study. Fracture Higuchi, M., Fatigue Curves And Fatigue Design Criteria For Carbon And Low Alloy Steels In High-Temperature Water, ASME PVP Vol. 386, 1999, pp.161-169. 9 Higuchi, M. and Iida, K., Effects of Strength and Sulfur Content on Fatigue Strength of Carbon Steel Weldments in Oxygenated High Temperature Water, International Conference on Pressure Vessel Technology Vol.1, ASME, 1996 10 Kitagawa H., Nakajima H., Nagata N., Sakaguchi Y., and Iwadate T., The Japanese Domestic Round Robin Tests On Cyclic Crack Growth Of A533B-1 Steels In Simulated LWR Primary Coolants, Proceedings of the second IAEA specialists meeting on subcritical crack growth, held in Sendai, May 15-17, 1985, NUREG/CP-0067, MEA-2090, Vol.1, 1986 20-6
8

surfaces are relatively flat and smooth for the specimens fatigued in air (a) or at normal strain rate (0.4%/s) in LWR environment (b). Those specimens fatigued at a lower strain rate in LWR environment always exhibit a rough fracture surface (c). Multiple cracks initiate for the specimen tested in the LWR water environment when strain rate is slow, many of which subsequently propagate simultaneously and ultimately link. Multiple hair cracks and corrosion pits were seen on the surface of the specimens fatigued in the LWR water environment at a strain rate of 0.004%/s as seen in Figure 8 (a). This condition was observed for most of the specimens fatigued at the slow strain rate (0.004%/s) in water. However, only a small number of corrosion pits were observed on the surface of specimens fatigued at the higher strain rate (0.4%/s) even in LWR water environment, as shown in Figure 8(b). Figure 8 (c) shows the surface of the specimens fatigued in air, where no corrosion pits or hair cracks were observed. It has been reported that crack initiation occurs in the early stages of fatigue life when tested in air, but the number of crack initiation sites are few, and usually very difficult to observe at the early stage of the fatigue life 11. Mean misorientation change of the cells during fatigue life is shown in Figure 9. Closed circle markers represent the SA508 specimens fatigued in air at normal strain rate (0.4%/s), and closed square markers represent the A533B-1 specimen fatigued in air at slow strain rate (0.004%/s). The mean misorientation is slightly lower for A533B-1 for both as-received and fatigued samples, and this is in good agreement with the fact that the cell structure is less developed for A533B-1 as compared to SA508. Open markers represent the specimens fatigued in the LWR environment. All of the misorientation (SAD) values of the failed specimens are smaller than the value of the specimens fatigued in air, but the misorientation (SAD value) also increases during the cycling in LWR environment. The observation that the SAD increases quickly in the early stage of the fatigue life followed by slower increase later in the life, is quite similar to the SAD changes of the specimens fatigued in air. Discussion The results show that the SAD method can effectively be utilized to monitor in-service components in contact with the reactor coolant and to determine if and when the components are in a fatigue damaged state even without detection of cracks. Life assessment is also possible for the components provided that the change of the SAD as a function of N/N f and the threshold of the SAD value in a water environment are determined. It is important to note that the microstructural changes during fatigue tests in water, even at the very slow strain rate, occur at the same order of magnitude as the changes observed during fatigue tests in air. This suggests that the mechanical damage is essential in determining component fatigue life, regardless of environment. The chemical (corrosion) factor in the water environment is secondary but considered to increase crack velocity during cycling. Thus, we can apply the microstructural conditioning concept for fatigue crack initiation proposed in the

Miller, K.J., Initiation And Growth Rates Of Short Fatigue Cracks, Fundamentals of Deformation and Fracture, Eshelby Memorial Symposium, Sheffield, 2-5 April 1984, pp.477499 20-7

11

analyses of the relationship between the SAD and N/N f in air to testing in water. Since the SAD at N/Nf=1 in water (3-3.5 degree) is slightly smaller than the SAD values observed in air (4-4.5 degrees), the chemical factor contributes to reduce the critical or threshold value for material conditioning. The mechanism of the physical damage process during fatigue is yet to be fully understood. It has been commonly accepted and has been confirmed in the past studies of this project that micro-cracks exist from the first several cycles of fatigue tests, initiating at small inclusions on the surface of the samples fatigued in air 12. The crack sizes found on the specimen surface were mostly the same (about 0.05 mm), i.e. crack initiation occurs in the very early stage, but the cracks did not grow until near the end of the lifetime, Nf. At this point small cracks started to grow abruptly leading to sample failure. One of the characteristic differences observed for the sample surfaces tested in water at the slow strain rate (0.004%/s) was that the density of the small cracks was much higher than those of fatigue samples tested in air. This is likely due to the high density of specimen surface corrosion pits as seen in Figure 8 (a) that provided additional crack initiation sites. It was also observed for the samples tested in water that the small initiated cracks started to grow from the beginning to the end of the fatigue life, through the continuous process of the growth and combination of the micro-cracks to develop one or two major cracks in the order of 5-10 mm length, resulting in an order of magnitude reduction in the fatigue life. The strain rate difference of the fatigue tests in water seems to influence the pitting density, and thus the micro crack growth behavior. For the normal strain rate fatigue tests, 0.4%/s, the number of surface pits and micro cracks were observed to be significantly less than fatigue tests at the low strain rate, 0.004%/s. The crack growth behavior at the normal strain rate was closer to that for fatigue tests in air. This may explain the difference in the crack surface morphology of the samples after tests. Specifically the samples tested in water at the slow strain rate exhibited a rough fracture surface while the fracture surfaces of the samples tested at the normal strain rate and in air were relatively flat as shown in Figure 7. Figure 10 schematically summarizes the results of the present study, where the crack growth behavior and the mean angular misorientation in the microstructure are plotted as a function of fatigue cycles for both tests in air and in water (at slow strain rate). Fatigue crack initiation and the micro crack growth during fatigue is surface sensitive and strongly depends on the environment and strain rate. The microstructural change measured by the SAD technique represents change in the bulk property. Thus, the SAD value is independent of the testing environment and only depends on the mechanical strains induced in the sample. The critical SAD value is given by the value at N=N f in any testing environment. Conclusions 1. The SAD value of the specimens fatigue tested in the water environment (DO = 1ppm) increases from about 1.5 (as received) to above 3.0 (N/Nf = 1) degrees, irrespective of the strain rate. The change is less than the SAD changes observed for specimens tested in air but
12

Measuring Fatigue Damage in Materials-Phase I; EPRI, Palo Alto, CA: 1998. TR-110250 20-8

well above the level of the distinction between the virgin and fatigue-failed state. 2. The observation in the specimens fatigue tested to N/N f < 1 that the SAD values increase quickly in the early stage of the fatigue life followed by a slower increase in later life, is quite similar to the SAD changes of the specimens fatigued in air . 3. At the very slow strain rate (0.004%/s) marked reductions in fatigue life are observed in the samples fatigue tested in the LWR water environment, namely the number of cycles to failure are an order of magnitude smaller than those tested in air or tested at the strain rate of 0.4%/s.

20-9

Table 1 Chemical Composition and Major Mechanical Properties of A533B-1 (a) Chemical Composition (weight percent) C 0.2 Si 0.3 Mn 1.5 P 0.02 S 0.020 Ni 0.5 Cr 0.1 Mo 0.5 Cu 0.1

(b) Mechanical Properties Yield Strength Tensile Strength (MPa) (MPa) 481 617

Elongation (%) 26

Reduction of Area (%) 61

Table 2 Environmental Fatigue Test Conditions Environmental Conditions Medium Temperature Pressure Electrical Conductivity pH Dissolved Oxygen Content Water Flow Rate Deionized Water 289 8.0 MPa <0.2 S / cm 5-7 1 ppm 60 liter/h Fatigue Test Conditions Control Mode Wave Shape Strain Rate Rising Phase Falling Phase Strain Ratio Axial Strain Triangle/Saw Tooth 0.4-0.004 %/s 0.4 %/s -1

Table 3 Fatigue Test and SAD Results for A533 in Air Specimen # Temperature (C) as-received SQV2A-A1 SQV2A-A2 SQV2A-A3 SQV2A-A4 SQV2A-A5 RT RT 289 289 289 a (%) 0.6 0.31 0.31 0.31 0.31 dT/dt N25 cycles N/Nf (%/s) 0.4 0.4 0.4 0.004 0.004 1680 6740 7380 5026 N=2500 0 1 1 1 1 0.5 SAD (deg) 1.42

3.86

R=-1, wave shape:triangle

20-10

Table 4 Fatigue Test and SAD Results for A533 in LWR Water Environment Specimen # Temperature a (C) (%) as-received SQV2A-W3 SQV2A-W4 SQV2A-W5 SQV2A-W6 SQV2A-W7 SQV2A-W8 SQV2A-W9 SQV2A-W10 SQV2A-W11 SQV2A-W12 289 289 289 289 289 289 289 289 289 289 0.6 0.6 0.31 0.31 0.31 0.31 0.31 0.31 0.24 0.24 dT/dt (%/s) 0.4 0.004 0.4 0.004 0.004 0.004 0.004 0.004 0.4 0.004 N25 cycles 845 94 4800 154 1 15 38 77 10895 350 N/Nf 0 1 1 1 1 0.01 0.1 0.25 0.5 1 1 SAD (deg) 1.42 2.8 3.05 3.51 2.26 2.14 2.33

R=-1, wave shape: triangle & saw tooth, DO=1 ppm

20-11

Figure 1 Fatigue Test Specimens Dimensions (unit: mm)

Figure 2 Schematic Illustration of the SAD Procedure 20-12

Figure 3 Normal distribution of the cell orientation for 0, 50, and 100% of fatigue life for samples cycled at total strain range of 0.78%

Figure 4 Mean Misorientation Difference Between Cells Measured by SAD Versus N/Nf

20-13

Figure 5 Strain-life Curve for A533

Figure 6 Relationship Between Strain Rate and Fatigue Life 20-14

Figure 7 Fractography of Fatigue Samples, Tested (a) in Air, (b) in Water at Normal Strain Rate, and (c) in Water at Slow Strain Rate

Figure 8 Surface Observation of Fatigued Samples by Optical Microscopy, Tested (a) in Water at Slow Strain Rate, (b) in Water at Normal Strain Rate, and (c) in Air

Figure 9 Mean Misorientation Change During Fatigue Life 20-15

Figure 10 Schematic Illustration of Relationship Between Fatigue Crack Growth and Misorientation Change During Fatigue Life

20-16

Microstructural Changes Of Pressure Vessel Steel During Fatigue In High Temperature Water Environment
C. Fukuoka, Y. Nakagawa, M. Higuchi Ishikawajima Harima Heavy Industries, Inc.- Japan S. Rosinski EPRI - USA
International Conference on Fatigue of Reactor Components 31 July - 2 August 2000 Napa, California
20-17

Objectives

Develop a technique to measure microstructural changes due to accumulated fatigue damage Correlate microstructural changes with measured changes in fatigue life
s

Various loading/cycling histories - above and below fatigue limit Consider reactor water environment

Assess potential for application to improved life assessment capabilities

20-18

Introduction

q q

Joint program between EPRI and IHI Previous activities under joint program
s

Fatigue testing performed on SA508 reactor pressure vessel forging material Influence of loading history on fatigue life investigated combination of cycling above and below fatigue limit Microstructural changes correlated with measured changes in fatigue life Results published in EPRI TR-110250, Measuring Fatigue Damage in Materials - Phase 1, March 1998

20-19

Program Scope

Scope of Phase 2 program


s

Investigate the effect of reactor water environment on microstructural changes that occur during fatigue Correlate microstructural changes with fatigue damage accumulation for A533B material exposed to reactor water environment Samples were fatigue cycled as a function of strain amplitude and strain rate Microstructural changes measured using Selected Area Diffraction (SAD)

20-20

Selected Area Diffraction (SAD)


q

Microstructural examination technique s Identify small angular misorientations in crystal lattice Angular misorientations observed through TEM s Misorientation measured between grains oriented in the <111> zone axis

20-21

Experimental Setup

All dimensions in mm

Environmental Conditions Medium Temperature Pressure Electrical Conductivity Deionized Water 289 8.0 MPa <0.2 S / cm

Fatigue Test Conditions Control Mode Wave Shape Strain Rate Rising Phase Falling Phase Strain Ratio Axial Strain Triangle/Saw Tooth 0.4-0.004 %/s 0.4 %/s -1

pH 5-7 Dissolved Oxygen Content 1 ppm


20-22

Water Flow Rate

60 liter/h

Results
Specimen # Temperature (C) a (%) 0.6 0.31 0.31 0.31 0.31 dT/dt N25 cycles N/Nf (%/s) 0.4 0.4 0.4 0.004 0.004 1680 6740 7380 5026 N=2500 0 1 1 1 1 0.5 SAD (deg) 1.42

Fatigue test and SAD results for A533B tested in air

as-received SQV2A-A1 SQV2A-A2 SQV2A-A3 SQV2A-A4 SQV2A-A5

RT RT 289 289 289

3.86

R=-1, wave shape:triangle Specimen # Temperature a (C) (%) as-received SQV2A-W3 SQV2A-W4 SQV2A-W5 SQV2A-W6 SQV2A-W7 SQV2A-W8 SQV2A-W9 SQV2A-W10 SQV2A-W11 SQV2A-W12 289 289 289 289 289 289 289 289 289 289 0.6 0.6 0.31 0.31 0.31 0.31 0.31 0.31 0.24 0.24 dT/dt (%/s) 0.4 0.004 0.4 0.004 0.004 0.004 0.004 0.004 0.4 0.004 N25 cycles 845 94 4800 154 1 15 38 77 10895 350 N/Nf 0 1 1 1 1 0.01 0.1 0.25 0.5 1 1 SAD (deg) 1.42 2.8 3.05 3.51 2.26 2.14 2.33

Fatigue test and SAD results for A533B tested in reactor water environment

20-23

R=-1, wave shape: triangle & saw tooth, DO=1 ppm

Results

Strain amplitude versus fatigue lifetime of SA508-2 in air and A533B-1 in air and in water

20-24

Results

Relationship between strain rate and fatigue life

20-25

Selected Area Diffraction Results


SAD values increase quickly early in life followed by more gradual increase SAD values of A508/A533 samples are lower when tested in reactor water environment

20-26

Discussion

Previous studies suggested threshold SAD value (N/Nf=1) for testing in air
s s

= 1.5 degrees as-received; 4-5 agrees at N/Nf=1 Independent of strain amplitude, strain rate

vs N/Nf for A508-2 as a function of strain amplitude

20-27

Discussion

Similar threshold SAD value suggested for A533B samples tested in air, ~4 degrees For testing in reactor water environment, lower threshold SAD values (N/Nf=1) measured, ~ 3 degrees Lower SAD values for both materials when tested in reactor water environment
s s

Independent of strain rate Suggests that chemical process contributes to reduce critical (threshold) SAD value

20-28

Conclusions
q

Selected Area Diffraction successfully used to measure microstructural changes in A533B material exposed to reactor water environment and correlate with fatigue damage accumulation SAD values increase with fatigue accumulation for samples tested in air and reactor water environment
s s

Independent of applied strain rate Increase in SAD value is smaller for specimens tested in reactor water environment than in air Change is above level of distinction between as-received and fatigue-failed (N/Nf=1) state; threshold value suggested

SAD technique appears to be feasible to determine component fatigue status, independent of environment

20-29

Recommended Activities

For development and application as an engineering tool, following activities recommended:


s

Determine SAD threshold values for wider spectrum of fatigue test experimental variables x Dissolved oxygen x Strain rate x Strain amplitude Apply SAD to materials extracted from operating plants

20-30

21
MICROSTRUCTURAL INVESTIGATIONS AND MONITORING OF DEGRADATION OF LCF DAMAGE IN AUSTENITIC STEEL X6CRNITI 18-10

D. Kalkhof, M. Grosse, M. Niffenegger Paul Scherrer Institute Nuclear Energy and Safety Villigen, Switzerland CH- 5232 D. Stegemann, W. Weber University of Hannover Institute of Nuclear Engineering and NDT Elbestrasse 38A Hannover, Germany D- 30419

21-1

EPRI / USNRC / OECD NEA CSNI INTERNATIONAL CONFERENCE ON FATIGUE OF REACTOR COMPONENTS JULY 31 - AUGUST 2, 2000 NAPA, CALIFORNIA, USA

MICROSTRUCTURAL INVESTIGATIONS AND MONITORING OF DEGRADATION OF LCF DAMAGE IN AUSTENITIC STEEL X6CRNITI 18-10
D. Kalkhof, M. Grosse, M. Niffenegger Paul Scherrer Institute Nuclear Energy and Safety Villigen, Switzerland CH- 5232 D. Stegemann, W. Weber University of Hannover Institute of Nuclear Engineering and NDT Elbestrasse 38A Hannover, Germany D- 30419

Abstract The microstructural changes in the pre-crack stage of low-cycle fatigue damage (LCF) in austenitic piping steels are characterised by neutron and X-ray diffraction. The LCF damage evolution in the metastable austenitic steel causes a deformation-induced phase transformation from austenite to martensite. Thresholds exist for the formation of martensite as a function of both, the load amplitude and the number of LCF cycles. Magnetic stray field and eddy current measurements were chosen to transfer the results of material characterisation to NDT methods. The density and distribution of martensite obtained from neutron diffraction experiments were used to adjust the NDT signals. Both NDT techniques were able to detect the very low amount of martensite in the different aged specimens (0.5 - 3.1 vol.% martensite, usage factors from 0 up to 1.0).

21-3

INTRODUCTION

The development of advanced diagnostic systems that are able to identify and detect material degradation in the microstructures of steels is a new challenge for non-destructive testing (NDT). Especially in nuclear power plants, early detection of material degradation can contribute to improve safety and reliability of the primary circuit boundary as well as plant life time management of certain components. Lifetime extension of nuclear power plants (NPP) has become an important topic and resulted in the requirement for advanced safety management tools. Therefore systems for early detection of material degradation of pressurised primary loop components are demanded. With regard to leaks in pressurised water pipes of nuclear power plants [e.g., 1-3], which were caused by thermal fatigue damage, deformation-induced changes in the microstructure of metastable austenitic steels were investigated. It is well known that varying temperatures and thermal stratification in piping of power plants might cause fatigue damage by thermal strains. Such strains promote the phase transition from a cubic face centred (fcc) austenite to cubic body centred (bcc) martensite [4]. Our structure investigations using metallography as well as neutron and X-ray diffraction are primarily intended to identify and quantify martensite formation in the pre-crack stage of low-cycle fatigue damage (LCF) in austenitic piping steels and its influencing parameters. Martensite is a ferromagnetic phase distributed in the austenitic matrix which itself is paramagnetic. This opens the possibility to detect the martensitic content by means of magnetometers. To monitor the pre-crack damage state, the application of magnetic NDT methods based on magnetic stray field and eddy current measurements is investigated. 2 MATERIALS AND TEST CONDITIONS

The investigations were performed on hour-glass specimens (according to ASTM E606, Fig.1). The material under investigation was the titanium stabilised austenitic metastable steel X6CrNiTi18-10, which is widely used for vessels and piping. The chemical composition (in wt.-%) is given in Tab. 1. In order to homogenise the microstructure, the specimens were annealed at 1040 C (1 h) and afterwards quenched in oil. The fatigue tests were performed with alternating loading with a test frequency of 2 s-1 at room temperature and were controlled by the total strain amplitude. A first test with a strain amplitude of 2.55 mm/m did not result in crack initiation. Therefore a strain amplitude of 2.95 mm/m was applied referring to the position of the smallest cross section. Different usage factors were realised by well defined cycles of strain loads in series of hour-glass specimens. Tab.2 gives the cycle number for some investigated specimens and the corresponding usage factor D. The usage factor indicates how many of the life time is passed. D is defined to be equal to 1 for specimens, where crack initiation has already occurred. The LCF tests were stopped when a 5% force drop was reached. Therefore the corresponding cycle numbers vary for specimens with D=1 due to differences in the damage progression. For specimens without a crack, D is defined as the ratio of the applied cycle number to an averaged cycle number representing crack initiation. This averaged value for crack initiation is obtained in pre-testing on 5 specimens. For the metallographic investigations, the specimens were cut lengthwise. After mechanical grinding and polishing, the specimens were electrochemically etched using Beraha-II solution [5]. Fig.2 shows the microstructure of specimen 1.6 at three axial positions. At middle positions dark streaks were found. With increasing distances from the middle of the specimen the width of the streaks decreases down to zero. At positions far from the middle, the microstructure is homogeneous. The microstructure of the dark streak regions consists of martensite needles in an austenitic matrix (Fig.3a). Outside these streaks only austenite grains can be found (Fig.3b). 21-4

Table 1. Chemical composition in wt.-% of the X6CrNiTi18-10 specimens C 0.05 Mn 1.08 P 0.039 S 0.019 Cr 17.55 Ni 9.86 Ti 0.39

d=18mm 32mm 75mm 220mm


Figure 1. Hour-glass specimens (ASTM E606)

-30mm from the middle of the sample

middle of the sample

+30mm from the middle of the sample

Figure 2. Microstructure of specimen 1.6 at different axial positions

21-5

a) inside the streaks

b) outside the streaks

Figure 3. Microstructure of specimen 1.6 in- and outside the dark streaks 3 STRUCTURE INVESTIGATIONS BY MEANS OF DIFFRACTION EXPERIMENTS

The martensite content formed by phase changes was measured by means of neutron and X-ray diffraction experiments. Austenite and martensite differ in their microstructure. Austenite has a face centered cubic lattice cell whereas the lattice cell of martensite is body centered cubic. This makes it possible to determine the phase composition in the steel by neutron and X-ray diffraction experiments. Neglecting textures, the volume fraction of a phase v1 in a two phase system (austenite and martensite) is given by [6]: v1=[(I{h,k,l}2 / I{h,k,l}1) K1,2 +1]-1 (1)

I{h,k,l} is the intensity scattered at the lattice plane {h,k,l}. K1,2 is a factor which considers the different structure type of martensite and austenite and the different angular positions of the reflexions. K1,2 factors are tabulated for X-rays in [6] and can be calculated for neutrons [7]. 3.1 Neutron diffraction (ND) The neutron diffraction experiments were performed at the powder diffractometer DMC at SINQ (Swiss Spallation Neutron Source, PSI Villigen) [8]. The instrument is optimised for high intensity. It allows the determination of the phase content down to values below 1 %. With a neutron beam wave length of = 0.38 nm, a range of the scattering angle 2 of 68 2 147 was analysed. The measurements on the LCF specimens were performed with a beam cross section of 40 mm (width) x 10 mm (height). The measuring position was in the middle of the LCF specimens (5 mm in height). Results of martensite content obtained by these measurements are summarised in Tab.2. At four specimens also the axial distribution of the martensite was measured. For this purpose a reduced beam height of 2 mm was used and the specimens were scanned in axial direction.

21-6

Fig.4 shows the specimen cross section variation, the corresponding strain amplitude and the measured martensite content in dependence on the axial coordinate of the specimen. It can be concluded that the martensite formation is limited to the area of the smallest cross section. In Fig.5, the axial distribution of the martensite content measured with the reduced beam height of 2 mm is given for four specimens. Three of them, 1.6, 2.3 and 2.5 have D = 1. The maxima of the martensite content are detected at the crack position. It seems that the crack formation results in an additional martensitic transformation. In the specimen 1.7 (without crack) the maximum of the martensite content was found at the smallest cross section. The martensite distribution in axial direction is unsymmetric for this specimen. This is also proved by the eddy current measurements (see 4.2). In Fig.6, the measured martensite content is plotted as a function of the strain amplitude. Below a threshold strain amplitude (about 2.2 mm/m for specimens 1.7, 2.3 and 2.5 and 2.5 mm/m for specimen 1.6), no martensite was detectable by neutron diffraction (detection limit of these experiments is about 0.5 vol.-%). For strains higher than the threshold strain amplitude, a nearly linear dependence of the martensite content on the strain amplitude was found if the data measured at the crack locations were not included. The slope of the curves depends on the cycle number. The dependence of the martensite content on the cycle number N is given in Fig.7. Also for the cycle number a threshold value (Nc = 13'000) exists, below which no martensitic transformation can be detected. At higher cycle numbers the martensite content depends nearly linearly on the cycle number.

21-7

Table 2. Martensite content obtained by neutron diffraction Sample No. 1.2 2.4 1.8 1.4 2.2 1.3 1.7 2.5 2.3 1.6 Cycle number 0 0 11200 15850 15850 22400 22400 26000 32000 63200 Usage factor D 0 0 0.4 0.6 0.6 0.8 0.8 1.0 1.0 1.0 Crack position 5 mm away from the middle in the middle in the middle Martensite (vol.-%) < 0.50 < 0.50 < 0.50 0.61 0.54 0.95 0.93 1.49 1.49 3.13

sample cross section


5.0

martensite content in vol.-%, cross section, strain amplitude

martensite content in %

sample 1.6: D=1.0, N=63000, crack at 0 mm

4.0

sample 2.3: D=1.0, N=32000, crack at 0 mm sample 1.7: D=0.8, N=22400, no crack sample 2.5: D=1.0, N=26000, crack at + 5 mm

strain amplitude

3.0

2.0

content of martensite
0 -40 -20 0 20 40

1.0

0.0 -30

-25

-20

-15

-10

-5

10

15

20

25

30

axial sample position in mm

distance from the middle of the sample in mm

Figure 4. Axial distribution of the martensite content for specimen 1.6 together with the axial dependence of the specimen cross section and the local strain amplitude

Figure 5. Axial distribution of the martensite content for specimens 1.6, 1.7, 2.3 and 2.5

21-8

5 sample 1.6: D=1.0, N=63000, crack at 0 mm

martensite content in %

4 sample 2.3: D=1.0, N=32000, crack at 0 mm 3 sample 2.5: D=1.0, N=26000, crack at +5 mm

martensite content in %

0 2.0 2.2 2.4 2.6 2.8 3.0

0 0 20000 40000 60000 80000

strain amplitude in mm/m

cycle number N

Figure 6. Martensite content in dependence on the strain amplitude for the specimens 1.6, 2.3, 2.5

Figure 7. Martensite content in dependence on the cycle number

3.2 X-ray diffraction (XRD) X-ray diffraction experiments using synchrotron light sources enable to analyse martensite distribution in two dimensions. The method guarantees a higher local resolution than neutron diffraction in shorter measurement periods, but is limited to the surface area. X-ray diffraction results give additional information on the radial distribution of martensite. The X-ray diffraction experiments were performed at the ROBL beamline at ESRF Grenoble (France) [9]. In order to use the K1,2 values from [6], the wavelength of the Mo-K radiation ( = 0.07107 nm) was applied. The investigated specimen 1.6 was cut into two halves parallel to the sample axis. A mapping of the axial and radial martensite distribution was gained by scanning the specimen through the fixed beam. With the applied wave length and the beam cross section of 0.2 mm x 0.5 mm, the analysed volume was about 1 mm (axial) x 0.5 mm (radial) x 0.005 mm (in depth). For the determination of the martensite content, the intensity relations between the {111}-austenite and the {110}-martensite peak measured in XRD experiments were used. Due to the small beam size and the very small beam divergence in the XRD experiments and the relative large grain size of the austenite (diameter about 20 - 30 m), the statistic condition for a polycrystal (infinite number of grains NG seen by the beam, in practice NG > 1000) is not completely fulfilled. A wide scatter of the intensity values of the {111}-austenite reflexion between the several specimen locations is the consequence. The grain size of the martensite is small enough to meet this condition. As the content of martensite is very small compared with the content of austenite, the mean value of the austenite {111}intensity can be used for the phase analysis. Using this mean value, the relative error of the martensite content (xM) is smaller than 6% (xM = 0.06 x xM).

21-9

Fig.8 shows the axial and radial mapping of the martensite content of specimen 1.6 determined from the Xray diffraction data. Some typical scattering patterns are included. In agreement with the ND measurements, in the axial direction, the martensite is found at the smallest cross section of the specimen. In the radial direction, the martensite content is concentrated near the surface (y= -6 and 7mm) and, as a broad area, in the center of the specimen (x,y = 0mm).

1 E +4

1 E +4

1 E +3

1 E +3

1 E +2 9 .6 9 .7 9 .8 9 .9 10 1 0 .1 1 0 .2 1 0 .3

1 E +2 9 .6 9 .8 10 1 0 .2 1 0 .4

8 6 4 2 0

A x is

3 3 2 2 1 1 0

.5 .0 .5 .0 .5 .0 .5

0 0 0 0 0 0 0

0 0 0 0 0 0 0 0

--------

4 3 3 2 2 1

.0 0 .5 0 .0 0 .5 0 .0 0 .5 0 1 .0

0 0 0 0 0 0 0 0

1 E +4
-2 -4

1 E +3
-6 -8 -1 8 -1 6 -1 4 -1 2 -1 0 -8 -6 -4 -2 0 2

1 E +2 9 .6 9 .8 10 1 0 .2 1 0 .4

A x is

1 E + 4

1 E + 4

1 E + 3

1 E + 3

1 E + 2 9 .6 9 .8 10 1 0 .2 1 0 .4

1 E + 2 9 .6 9 .7 9 .8 9 .9 10 1 0.1 1 0.2 1 0.3

crack tip position Figure 8. Mapping of the radial and axial distribution of the martensite content and typical XRD patterns

21-10

TESTING FOR LCF DEGRADATION BY MEANS OF MAGNETIC NDT METHODS

By means of neutron diffraction experiments it was demonstrated that the martensite content depends on the usage factor for the realised loading conditions at room temperature. Based on these results, it appears challenging to develop methods for the use in screening tests to identify areas with an increased amount of martensite. In this way, an early detection of LCF damaged zones could be managed. In case of LCF damage in metastable austenitic steels, the application of magnetic methods is possible as the deformation-induced martensite is of ferromagnetic and the austenitic matrix of paramagnetic nature. The problem is to measure very low amounts of martensite between 0.5 and 3 vol.-%. Therefore sensitive magnetic sensors must be used to identify local differences in magnetic properties. For an advanced NDT technique it is also expected that the local distribution of the martensite can be visualised. In principle it is possible to apply active and passive magnetic techniques. Concerning the martensite content, results of the neutron diffraction experiments are used to adjust the NDT signals. Note, that for the usage factor D = 1, the LCF damage is different with regard to the cycle numbers. 4.1 Measurements of magnetic stray field strength Due to the ferromagnetism of martensite it is possible to apply a passive magnetic technique. After magnetising of the specimens, the remanent field strength can be measured. The magnetisation of the specimens was carried out in a field with a strength of about 11 kA/m. Afterwards all specimens were examined using high sensitive SQUID and fluxgate sensors. Measurements of the axial and normal components of the magnetic stray field were performed in a distance of 20 mm from the bar axis. The fluxgate sensor was sensitive enough to detect martensitic areas and the results did not show any differences compared with the SQUID measurements. The fluxgate output (axial component) is plotted as a function of the axial scanning position in Fig.9. This output shows a pronounced positive peak when the sensor is directly over the position of the smallest cross section of the LCF samples. In the case of measurements on samples with cracks (D = 1), the amplitude of field strength depends on the cycle number. For the pre-crack stages the amplitude of the fluxgate output increased with the usage factor or with the corresponding cycle numbers. The LCF damage in the specimens was detected down to the usage factor of D=0.6. The fluxgate output for the usage factors 0.6, 0.4 and 0.0 do not differ significantly.

21-11

3,0
T

2,5 2,0 2 3

Strength (axial) of magnetic stray field, 10

1,5 1,0 0,5 0,0 -0,5 -1,0 -1,5 -200

4 5 6 7

1 : D = 1 .0 ( N = 6 3 0 0 0 ) 2 : D = 1 .0 ( N = 3 2 0 0 0 ) 3 : D = 1 .0 ( N = 2 6 0 0 0 ) 4 : D = 0 .8 ( N = 2 2 4 0 0 ) 5 : D = 0 .8 ( N = 2 2 4 0 0 ) 6 : D = 0 .6 ( N = 1 5 8 5 0 ) 7 : D= 0

-6

-150

-100

-50

50

100

150

A x i a l p o s i t i o n , mm

Figure 9. Stray field measurements (fluxgate output) at LCF specimens of different usage factors, cycle numbers and martensite content

21-12

4.2 Measurements by means of eddy current techniques Eddy current techniques seem to be very promising for detecting LCF degradation. If the eddy current field is locally disturbed by changes of permeability, the eddy current signal is changed. The induced voltage amplitude referred to a defined phase angle of the receiving coil was recorded. It is favourable to apply miniaturised sensors. The distance between the measuring sensor and the surface should be small. The applied manipulator system was able to follow the hour-glass shape of the specimens, thus the distance between sensor and surface was always 1 mm. An essential aspect for eddy current measurements is the choice of the exciting frequency. For martensite visualisation a frequency of 50 kHz was used which provides a reasonable balance between high local resolution and depth of the eddy current field in the steel. The lower and upper bound of 0.5 and 3.15 vol.-% martensite content define the adjustment of the eddy current sensitivity. For eddy current measurements, the end pieces of the LCF specimens were cut off and only the middle sections 75 mm in length were used. Additionally, the specimens were cut lengthwise into two halves. Thus it is possible to determine the martensite distribution also in the radial specimen direction and obtain information about the progress of the martensite formation. Some examples of eddy current output obtained on the circumferencial surface and on the axial cross section are plotted in the Figs.10 and 11. The plots visualise the martensite distribution on the circumferential surface (Fig.10) and on the axial cross section (Fig.11). The axial specimen direction corresponds to the horizontal coordinate axis, the circumferential and radial specimen directions correspond to the vertical coordinate axis. Plots are shown for different usage factors (specimen numbers 1.4, 1.7 and 2.3 with the corresponding usage factors 0.6, 0.8 and 1.0). In Fig.10 the eddy current signal amplitudes at the circumferential surface are plotted for an angle range of 180. The martensite content of 0.61 vol.-% for specimen 1.4 was clearly detected. The martensite is homogeneously distributed in the circumferential direction. The plot shows the beginning of the martensite formation during LCF loading. The changes of the martensite density in the axial direction show that the formation of martensite was really deformation-induced. In the following damage stages D = 0.8 and D = 1.0, the martensitic pattern is the darker as the usage factor increases. For specimen 1.7, the martensite is not as homogeneously distributed as for specimen 1.4. It is possible that the axial loading was accompained by a small amount of bending. For specimen 2.3, loaded up to crack initiation, a single large crack was detected at the position of the smallest cross section. Furthermore, some local areas with a very high martensite concentration were found. These are probably new sources for macroscopic cracks. The martensite is homogeneously distributed. At the crack position an additional formation of martensite was observed. In Fig.11, the eddy current signal amplitudes at the axial cross section surface are plotted. The sensor output for specimen 1.4 shows, that the formation of martensite is limited at the surface in the position of the smallest cross section. No martensite was detected in depth. For further damage progression (specimen 1.7), the martensite tends to spread in depth. The distribution of martensite in the axial direction is wider at the surface than in the bulk. A macroscopic crack initiation will be effected by stress relaxation around this location. This results in a decrease of the martensite spreading in this area which is shown for specimen 2.3. The shape of the martensitic pattern is broader at the opposite side of the crack.

21-13

Figure 10. Eddy current measurements on the circumferencial surface of LCF specimens 1.4 (D=0.6, xM=0.61 vol.-%) 1.7 (D=0.8, xM=0.93 vol.-%) 2.3 (D=1.0, xM=1.49 vol.-%)

Figure 11. Eddy current measurements on the axial cross section of LCF specimens 1.4 (D=0.6, xM=0.61 vol.-%) 1.7 (D=0.8, xM=0.93 vol.-%) 2.3 (D=1.0, xM=1.49 vol.-%)

21-14

CONCLUSION

All applied methods (neutron and X-ray diffraction, magnetic stray field measurements with SQUID and fluxgate sensors, eddy current measurement) allowed the detection of martensite. While the amount of martensite could be quantified with the diffraction methods, the magnetic techniques yield only relative results; that means, they have to be calibrated somehow, i.e. with results from neutron diffraction experiments. Neutron and X-ray diffraction are powerful tools to determine the deformation-induced martensite content and its distribution in hour-glass specimens. Neutron diffraction is the only method to obtain the absolute value of martensite content in a specimen section covering the whole thickness. It was shown that the content of martensite depends on the usage factor. The existence of thresholds for the cycle number and the total strain amplitude to form deformation-induced martensite could be concluded (total strain amplitude: 2.2 .. 2.5 mm/m, cycle number: 13'000). These thresholds are useful when examining industrial components under service conditions. It should be clarified whether thresholds also exist for other conditions of loading and temperature. Both applied magnetic techniques, stray field and eddy current measurements, were able to detect material degradation due to low-cycle fatigue. For our testing conditions the eddy current technique was able to detect LCF damage down to D=0.6, the stray field measurements down to D=0.8. No significant differences of measuring signals for both techniques were found for D < 0.6, which corresponds to the results obtained by neutron diffraction experiments. For practical application, eddy current measurements in combination with neutron diffraction investigations are recommended. The correlation between the usage factor and the NDT signal is clearly demonstrated which is rather promising for further investigations. However, further investigations namely for higher temperature ranges are needed before the application of the method for real components can be tested. 6 ACKNOWLEDGMENTS

This work was supported by the Swiss Federal Nuclear Safety Inspectorate, HSK. REFERENCES 1. 2. 3. 4. 5. Experience with thermal fatigue in LWR piping caused by thermal mixing and stratification. Paris: OECD Nuclear Energy Agency, 1998. OECD NEA CSNI R(98)8 Leak on the Residual Heat Removal System at Civaux 1. Vienna: International Atomic Energy Agency, 1999. IAEA TC 485.27 R. Rantala, Thermal Fatigue Experiences and Countermeasures in Finland. Paris: Meeting of the OECD Nuclear Energy Agency CSNI Principal Working Group 3, 1999 R.W.K. Honeycombe, Steels, Microstructure and Properties. Edward Arnold Publishers Ltd., 1981 Metallographische Anleitung zum Farbtzen nach dem Tauchverfahren, Band2: Farbtzung nach Beraha und ihre Anwendungen. Dsseldorf: Deutscher Verlag fr Schweisstechnik, Herausgeber, E.Weck, E.Leistner, 1983

21-15

6. 7. 8. 9.

G.Fanninger, U.Hartmann, Physikalische Grundlagen der Quantitativen Rntgenographischen Phasenanalyse (RPA). Hrterei Technische Mitteilungen HTM 27, 1972, p. 233 G.E.Bacon, Neutron Diffraction, Oxford: Clarendon Press, 1975. http://www1.psi.ch/www_sinq_hn/SINQ/instr/ DMC.html W.Matz et al., ROBL- a CRG Beamline for Radiochemistry and Materials Research at the ESRF. Journal Synchrotron Radiation 6, 1999, p. 1076

21-16

22
NDE TECHNOLOGY FOR DETECTION OF THERMAL FATIGUE DAMAGE IN PIPING
Stan Walker Pedro Lara EPRI NDE Center 1300 WT Harris Blvd. Charlotte, NC 28262

22-1

NDE Technology for Detection of Thermal Fatigue Damage in Piping

Stan Walker Pedro Lara


EPRI NDE Center 1300 WT Harris Blvd. Charlotte, NC 28262 704/547-6081 swalker@epri.com

ABSTRACT Recent piping failures related to thermal fatigue have prompted a review of examination practices currently used for detection of thermal fatigue damage in piping, particularly in small-diameter piping (less than 4-inches in diameter). Because volumetric examination techniques are difficult and most examinations of small-diameter piping are currently limited to surface examinations, the industry has expressed a desire to develop a plan for both near- and long-term solutions. A program is being developed that includes criteria and methods to improve the reliability of small-diameter piping examinations. In response to these concerns, an Issues Task Group (ITG) was put together under EPRIs Pressurized Water Reactor Materials Reliability Project. The inspection-related deliverables of the ITG, Thermal Fatigue Inspection Guidelines, are the subject of this paper. Specifically, this paper describes the results of the evaluation of candidate nondestructive evaluation (NDE) technologies for thermal fatigue crack detection and recommends guidance for NDE examiner qualifications.
INTRODUCTION

Leaks of thermal fatigue origin in piping connected to the reactor coolant system have occurred at several power plants in the US, Europe, and Japan, and the Nuclear Regulatory Commission (NRC) has expressed concern with the occurrences. Because these failures have included small diameter piping where volumetric inspection techniques were considered to be unreliable, the NRC has expressed their desire for the electric industry to develop a plan for both near- and long-term solutions. This activity was aimed at helping the industry reach consensus on the optimal NDE methodology and its capability to detect thermal fatigue cracking in small diameter piping. To accomplish this objective, the relevant experiences and currently-accepted practices were reviewed. Evaluations were performed on mockups containing thermal fatigue cracks to document the capabilities of various NDE techniques. From these evaluations, guidance about the qualification of NDE examiners was established.

22-3

THERMAL FATIGUE CRACKING Thermal fatigue damage typically occurs as a result of a component malfunction, which causes an unplanned mixing of hot and cold fluids. In addition to a faulty component, conditions for thermal fatigue require the formation of a temperature stratified fluid region, with the stratification varying periodically. The temperature difference between the mixing fluids must be sufficiently large so that the associated principal stress exceeds the fatigue endurance limit value. Although the crack distribution is somewhat random, some common patterns can be identified from the reported events. The thermal fatigue cracks originate at the inner surface and grow to the outside surface. The cracks initiate as a shallow network called craze. These craze cracks are transgranular in nature and exhibit extensive branching. Although the direction of these cracks is somewhat random, a checker-board appearance has been observed in many situations. These craze cracks are reported as shallow, however, actual quantification of their depth is scarcely documented. The deeper cracks (responsible for the component failure) generally occur in the midst of the craze cracking, although some exceptions to this pattern have been reported. The resulting crack system typically includes one or more deep cracks surrounded by an extensive network of shallower penetrating defects. Cracks that originate away from welds tend to have an axial orientation. These axially-oriented cracks are skewed, with the skewness angle varying somewhat randomly. In the vicinity of welds, the crack orientation is likely to be circumferential.1

MOCKUPS Six specimens with implanted thermal fatigue cracks were manufactured, with the purpose of providing targets against which the effectiveness of the candidate NDE techniques could be gauged, and to be used as part of the examiner qualification procedure. The mockups were made to test the ability of the NDE technique to Perform a volumetric examination on high curvature piping and elbow surfaces; Size the length of axial, circumferential, and skewed cracks located in the midst of a craze network; Size the extent of the craze crack network.

All specimens are made of austenitic stainless steel type 304L.

22-4

The crazing and crack penetration depths were selected to be at the low end of the measurements reported at power plants which experienced failures: from 0.025 to 0.147 inch. Accordingly, the specimens were made with craze depths of 0.02 or 0.04 inch, and cracks were produced with depths twice the depth of the surrounding craze.

RESULTS From the evaluations performed to-date, the following trends and observations have been derived Manual ultrasonic techniques were capable of detecting and measuring the length of craze and thermal fatigue cracks in each mockup. The time-of-flight diffraction technique detected, length sized and depth sized the craze and thermal fatigue cracks in 5 of the 6 mockup samples tested. The craze was not detected in one sample. Conventional radiography detected the thermal fatigue cracks and some of the craze cracks in the mockup specimens. Craze detection was sufficient to outline the craze affected area. Ultrasonic spectroscopy detected the craze cracks in 4 of the 6 mockups evaluated. The areal extension of the craze was accurately sized. The thermal fatigue cracks were not detected. Pulsed Eddy Current detected the craze cracks with depths greater than 10 percent. The technique did not detect the thermal fatigue cracks. Vibro-Modulation detected the presence of craze cracks in 2 of the 3 mockup specimens tested. Areal extension of the craze could not be derived from the signals. Conventional eddy current was judged not to be suitable for this application because of the limited penetration of the current field.

From these results the following conclusions were obtained: Manual ultrasonic examination techniques were found to perform best among the technologies considered in this evaluation. This technique is viable for detecting and length-sizing cracks of thermal fatigue origin in small bore piping when applied in accordance with the EPRI procedure. Time-of-flight diffraction is a viable detection technique for scanning large areas semi-automatically to be supplemented with manual ultrasonic techniques when more precise length-sizing evaluations can lead to a better repair schedule. Conventional radiography, with fine-grain film, is a viable technique for detecting thermal fatigue cracks deeper than 10 percent of the wall thickness. This technique can be supplemented with manual ultrasonic techniques when the examination objective includes detecting craze damage in its early stages.

22-5

Ultrasonic spectroscopy, pulsed eddy current, and vibro-modulation require further development before they can be implemented in a power plant environment for thermal fatigue crack inspection.

Accordingly, a generic procedure for inspection using manual ultrasonic techniques was developed. The generic procedure was further tested on a field-extracted, 1-inch diameter elbow that exhibited thermal fatigue damage. Finally, it is recommended that examiners receive approximately 4 hours of indoctrination prior to performing examination in power plants. This indoctrination should be administered to examiners who have previously demonstrated proficiency in ultrasonic examination of piping welds through some industry standard. 1

QUALIFICATION OF ULTRASONIC EXAMINERS Examiner qualification is an important aspect of NDE. Beginning in 1983, it has been mandatory that all persons performing ultrasonic examination at any domestic boiling water reactor (BWR) facility for detection of intergranular stress corrosion cracking (IGSCC) satisfactorily complete training and performance demonstration. This was mandated because of the history of poor performance, and/or the occurrence of leakage due to IGSCC within a short time after examination. Currently, the majority of NDE personnel performing examinations at nuclear power stations has received training for ultrasonic detection of IGSCC and has demonstrated satisfactory performance on field-removed specimens. Because thermal fatigue cracking has characteristics entirely different from IGSCC, and occurs from different causal factors, it is recommended that all personnel performing ultrasonic examination for thermal fatigue cracking receive additional indoctrination (or training). The indoctrination recommendations are based on knowledge of crack morphology, laboratory experimentation, and field experience. It is recommended that this indoctrination be administered to examiners who have previously demonstrated proficiency in ultrasonic examination of piping welds through some industry standard. The industry standard could be a formal qualification such as the American Society of Mechanical Engineers (ASME) Section XI Appendix VIII qualifications as administered by the Performance Demonstration Initiative (PDI) or some other industry-recognized standard. In addition to the completion of some existing piping qualification program (which may have been conducted on samples with IGSCC or other cracking), it is recommended that examiners for thermal fatigue damage receive indoctrination of approximately 4 hours prior to performing examinations at power plants. This indoctrination should include information on formation of thermal fatigue cracking, the appearance of craze cracking, thermal fatigue crack morphology as contrasted to IGSCC, ultrasonic interaction with thermal fatigue cracking, and the use of specialized ultrasonic probes for small diameter pipes. This indoctrination should also include a visual evaluation of thermal fatigue cracks and a hands-on ultrasonic examination of known thermal fatigue cracking. A course is currently being developed to provide this
22-6

recommended indoctrination. It is not necessary for examiners to complete a practical demonstration of proficiency in addition to the one completed for the aforementioned industry standard.1

REFERENCE 1. Draft 100015, NDE Technology for Detection of Thermal Fatigue Damage in Piping (MRP-TF-06), EPRI, Palo Alto, CA and MRP TF-ITG, May 2000

22-7

23
NDE METHOD FOR DETECTION OF FATIGUE DAMAGE
G. L. Kessel Dominion Generation Innsbrook Technical Center 500 Dominion Boulevard Glen Allen, Virginia 23060 J. K. Na Energy Service Group International, Inc. 8979 Pocahontas Trail Williamsburg, Virginia 23185-6243 W. T. Yost and J. H. Cantrell NASA Langley Research Center Hampton, Virginia 23681-0001 Y. L. Hinton U. S. Army Research Laboratory Vehicle Technology Directorate MS 231 Hampton, Virginia 23681-2219

23-1

NDE METHOD FOR DETECTION OF FATIGUE DAMAGE

G. L. Kessel Dominion Generation Innsbrook Technical Center 500 Dominion Boulevard Glen Allen, Virginia 23060 J. K. Na Energy Service Group International, Inc. 8979 Pocahontas Trail Williamsburg, Virginia 23185-6243 W. T. Yost and J. H. Cantrell NASA Langley Research Center Hampton, Virginia 23681-0001 Y. L. Hinton U. S. Army Research Laboratory Vehicle Technology Directorate MS 231 Hampton, Virginia 23681-2219

23-3

Abstract An NDE tool measuring fatigue levels of steam turbine blades has been developed by NASA and Virginia Power. The background technology is based on the nonlinear acoustics that has been used to study fatigue damage in aluminum alloys at NASA. As an initial effort for the joint development, a base line study was performed with 410Cb stainless steel by using the state of the art facility in NASA. In-situ data collection on steam turbine blades at the Virginia Powers power stations became possible when a portable fatigue measurement system was designed and built. The results show that the nonlinearity parameter, , is useful to predict the fatigue life. Introduction Microstructural damages caused by fatigue in metallic materials are usually related to the accumulation of various types of dislocation structures. Depending on the materials crystal structures, somewhat different dislocation substructures are formed during the fatigue process. However, in general, dislocation dipole densities increase as the fatigue process continues until the initiation of a crack. A pictorial diagram in Figure 1 shows the relation between the acoustic nonlinearity parameter and the fatigue cycles as the fatigue process continues until the formation of a crack in wavy-slip materials. It has been found theoretically and experimentally that the values of increase as the fatigue damage accumulates through a dynamical process involving dislocation dipoles [1, 2].

Stress Crack Singularity Formatio Vacancy Generation Region


PS B
tion Disloca Dipole s

Intersection of PSB with Material Discontinuity

Frozen PSB

Pre PSB (vein) like structures

tions Disloca
Log Cycles Figure 1. Relation between and number of fatigue cycles for typical wavy-slip alloys.

23-4

NASA Efforts in Nonlinear Acoustics and Early Work in Fatigue For the last decade or so, NASA has been developing the nonlinear acoustics to use as a nondestructive evaluation technique for assessment of fatigue damages in aluminum alloys used in aircraft. The initial work in fatigued aluminum was performed in aluminum 2024-T4 and the results have been published [3, 4, 5]. The nonlinear acoustics technique referred in this paper relies on the detection of the second harmonic signals which generated from wave distortion caused in part by the microstructure of materials. In order to generate harmonic signals, a transmitting ultrasonic transducer is mounted on one side of the test specimen as shown in Figure 2(a). Initially, it launches a pure sinusoidal ultrasonic tone burst signal with a fixed frequency into the material. The medium distorts the wave during propagation and the distorted waveform reaches the receiving transducer on the other side of the specimen. The signal is frequency analysed during which the amplitude of fundamental and second harmonic components are determined. See Figure 2(b). The acoustic nonlinearity parameter, , is then determined by measuring the amplitudes of both fundamental and second harmonic signals and using the following expression: = 8/(k2a) (A2/A12) where k is the wave number (2/), a is the thickness of the test specimen, A2 and A1 are the amplitudes of second harmonic and fundamental signals, respectively. For calibration of reference samples, we use a capacitive detector (microphone) as the receiving transducer placed on a carefully prepared sample. The use of the capacitive detector permits absolute determination of the acoustic wave amplitudes. The capacitive detection system has been used to study the fatigue damage in aluminum 2024-T4 induced by loading a cyclic tensile stress at 85% of the yield strength on an MTS machine. The results of the test show that the value of increases as the fatigue cycles increases. More details can be found in the references [6, 7]. (1)

Current Work between NASA and Virginia Power NASA and Virginia Power have developed an NDE tool which can help to determine the fatigue damage in steam turbine blades. This joint effort has been exercised since 1993. As the first step of the investigation we performed a basic study with 410Cb stainless steel as we did with the aluminum 2024-T4 alloy. The dog bone shaped fatigue specimens were examined by sectioning out the gauge area after they were fatigued on an MTS frame. The results showed the same trend in as the aluminum alloys as shown in Figure 3. It is clear that the value if increases as the fatigue damage increases. After 10 million cycles of fatigue, the increase in value was found to be

23-5

Specimen

Transmitting Transducer

Receiving Transducer

(a) Acoustic signal distorts as it propagates through the materials under test.

Fundamental
Amplitude (dB)

Second Harmonic Third Harmonic Frequency (MHz)

(b) Frequency analyzed of the received signal after transmission through the material under test.

Figure 2. Generation of harmonic signals due to wave distortion due to micro structures. about 90%. However, the value measured on a section cut out from the lower vane area of an actual turbine blade, which was retired for cause, was approximately 500% higher than the value of virgin material. This result indicates that the fatigue damage on the turbine blades are much higher than the simple tensile stress fatigue mechanism. In turbine blades, the fatigue mechanisms may be influenced by environmental factors not present during the fatigue of the 410 stainless steel used in the fatigue specimens. The measurement system used for the basic studies uses generic apparatus for each function and consequently takes almost two full racks of equipment. Since a bulky system is impractical for use in power plant environment, we identified a commercial manufacturer (Ritec, Inc) of a system adoptable for field use. A block diagram for this measurement system is shown in Figure 4. A 5 MHz pure sinusoidal tone burst signal is generated by a gated amplifier and filtered by a high power low pass filter before it is sent to the transmitting transducer. The signal received by the receiving transducer is split into two band pass filters. One is for the fundamental component (5 MHz) and the

23-6

other is for the second harmonic component (10 MHz). Once the amplitudes of these two components are measured by a digital oscilloscope, the value can be determined by a reference technique developed for these measurements.

80 70 60 50 40 30 20 10 0 10 1 10 2 10 3 10 4 10 5 10 6 10 7 y = 1.2 Log(x) + 13.0 nonlinearity parameter of an actual turbine blade

Figure 3. Nonlinearity parameter as a function of fatigue cycles for 410Cb stainless steel.

Tone Burst Genrator/Receiver Gated Receiver Signal Amplifier Monitor No.1 No.2 Output

50 Terminator

5 MHz Low Pass Filter

0.48 cm Receiving Transducer 5 MHz Fundamental Band Pass Filter Splitter 10 MHz Band Pass Second Harmonic Filter Digital Oscilloscope (LeCroy) 1.27 cm Transmitting Transducer (5 MHz)

Sample

Figure 4. Block diagram for the fatigue damage measurement system.

23-7

The reference sample measurement technique [8] uses a standard calibration sample to determine . Field measurements on the turbine blade are directly compared with equivalent measurements on the reference sample. Under these conditions, the field measurements are converted directly to the nonlinearity parameter values, . The advantage of this reference sample method is that the voltage levels measured for the fundamental and second harmonic signals are directly used without making a complicated calibration procedure normally required with the capacitive detection method. This is a significant improvement in reducing the measurement time in the field because measurements on more than one hundred blades need to be tested in a limited access time for non-destructive inspection during the outage. Other factors which affect these field measurements include non-parallel surfaces and surface roughness. Blades are normally sand blasted with 220 grit aluminum oxide beads for an inspection. In addition to this, the cross section of the blades is airfoil shaped with cant angle variable across the surface from 5 to 15 degrees. We have developed correction formulas to compensate for theses effects. The experimental results are shown in Table 1 [9]. The general trends can be seen from this table are as follows: the rougher the surface, the lower the value while the higher the cant angle, the higher the value.

Results of Turbine Blade Study The nonlinearity parameter was measured on 403 stainless steel turbine blades used for 17 years on Unit 5 at the Virginia Powers Chesterfield Power Station during a maintenance outage. Fourteen blades were selected randomly from the L-0 stage blades. The measurements were taken on the boss areas around the two tie wire holes and a region midway between the tie wire holes. The test results are shown in Figure 5. Five or more measurements at each location on each blade were taken. The standard deviations are shown by the error bars. In most cases, the standard deviations are small which demonstrates the reliability of the measurement. At a glance, we notice that the values of the outer tie wire regions are slightly higher than the other two regions. Although each blade may have experienced different fatigue loading over the entire service period, the overall trend seems to be that the outer tie wire area is subject to more fatigue damage. For comparison of the values of the three test regions, average values for each region of all tested blades are calculated and plotted in Figure 6(a). The average value for the outer tie wire is 41.3, an increase of nearly 500% over of the virgin material which ranges between 7.2 and 8. Near the inner tie wire, the average is 27.3, while midway between the two tie wires the value is 32.2 (increases of about 300% and 400% over the virgin value). This may indicate that the region around the outer tie wire experiences more fatigue induced damage than the other two regions. The tie wires

23-8

Table 1. values of canted specimens at various surface finishes Roughness RMS 330 RMS 280 RMS 200 RMS 125 RMS 63 RMS 2 Optical 0o 4.2 4.8 5.3 5.2 5.8 6.5 7.3 5o 7.7 7.2 7.5 3.4 4.0 4.5 5.1 10o 12.4 12.2 11.6 11.5 10.7 9.9 9.3 15o 26.3 23.6 20.4 18.7 17.2 15.4 13.6

reduce the amount of flexural bending in the turbine blades during normal operation. The region around the inner tie wire would therefore experience less bending due to constraints by the rest of the blade resulting in a lower value compared to the outer wire region. In addition, the boss area around the tie wire is of greater thickness, which increases the flexural stiffness and further reduces bending. The area between the two tie wires is slightly less restrained and able to bend more, and hence is subject to more fatigue damage accumulation than the inner tire wire area. More nonlinearity parameter measurements were made on the lower vane section of randomly selected turbine blades of L-1 stage on Unit 3 at the Mount Storm Power Station of Virginia Power and the results are shown in Figure 7. The values of these 10 year old blades are higher than the virgin material, with one blade showing a value of over 60. This is relatively high for a 10 year old blade. However, considering the fact that some blades have failed within 20 years of service life in the past, this result may indicate that the particular blade may be experiencing a higher fatigue loading compared to the other blades on the same rotor. Recently, Virginia Power provided a 26 inch long L-1 stage steam turbine blade (410Cb stainless steel) retired from service from the Chesterfield Power Station. The blade was designed and manufactured by General Electric and has one tie wire hole and a single boss on the pressure side. The blade served for 18 years and was removed from the rotor because of a crack that was initiated from the tip of the tie wire hole on the suction side and has grown toward the trailing edge approximately a half inch long. The crack is visible with naked eyes from both sides of the blade. An X-ray image showing the crack is shown in Figure 8 with a picture of the corresponding region.

23-9

The nonlinearity parameter was measured on the vane section of the cracked blade as shown in the Figure 9. The rectangular image was constructed with the values measured at the step of one eighth of an inch over the entire area. An enlarged image is shown in Figure 10. Higher values were measured on the boss area compared to other

Beta at Outer Tie Wire


70 60 50

40 30 20 10 0

Beta Between Tie Wires


70 60 50

40 30 20 10 0

Beta at Inner Tie Wire


70 60 50

40 30 20 10 0

Individual Blades

Figure 5. measured at three different locations on each of fourteen blades. Error bars indicate standard deviations. 23-10

60 50 40 Material: 403 Stainless Steel Age: 17 Years

30 20 10 0 Outer Tie Wire Midspan Inner Tie Wire

Figure 6. Average values at the three different locations on 17-year old Chesterfield blades. 70 60 50 Material: 403 Stainless Steel Age: 10 years

40 30 20 10 0 Individual Blades

Figure 7. values of 10-year old Mount Storm blades.

23-11

regions on the vane. From this result, it is clear that the boss area is subject to a higher levels of fatigue damage. Some spots on the boss are even higher than 150 for the value. The direction of crack propagation appears to intersect regions of maximum in the boss area. If one compares the X-ray image in Figure 8 and image in Figure 10, the crack had propagated toward the two high value spots indicated by the two hollow arrows. The values on these spots are over 150.

Tie Wire Hole (suction side)

Tie Wire Hole (pressure side)

Fatigue Crack

Tip

Root

(a) X-ray image

(b) Photographic image

Figure 8. An X-ray image and its corresponding photographic image of the crack at the tie wire hole of a retired blade. On the photographic image, the invisible suction side tie wire hole and the crack is drawn in for a better comparison.

Anticipated Benefits The fatigue sensor results will be used by turbine owners to determine the optimum time to replace components. The present method for blade replacement is to repair cracked components and then replace these blades as materials become available. The procurement time for some steam turbine blades can be over one year. The fatigue sensor results will allow blades to be replaced prior to crack initiation, and will allow the equipment to operate for longer periods between maintenance inspection intervals.

23-12

Boss Area 9" Pressure Side

1.25" Tie Wire Hole

23-13
5 0 . 0 1 0 0 . 0 1 5 0 . 0

Boss Area

23-14

The down time due to a steam turbine blade failure can be weeks or even months in length, depending on the availability of parts. By being able to measure the fatigue of a component, the life of a component can be predicted, and the part can be replaced prior to failure. The ability to replace blades prior to crack initiation will increase the reliability of the turbine, and will reduce the unexpected down time caused by blade failures. This ability to optimise blade replacements will reduce the turbine operating and maintenance costs

Acknowledgment This work is supported by Virginia Power.

References 1. Hikata, A., Chick, B. B., Dislocation contribution to the second harmonic generation of ultrasonic waves, Journal of Applied Physics, Vol 36, 1965, p. 229. 2. Cantrell, J. H., Yost, W. T., Acoustic harmonic generation from fatigue-induced dislocation dipoles, Philosophical magazine A, Vol. 69, No 2, 1994, pp. 315-326. 3. William T. Yost and John H. Cantrell Materials characterisation using acoustic nonlinearity parameter and harmonic generation: engineering materials, Review of Progress in Quantitative Nondestructive Evaluation, Vol. 9, 1990, p. 1669. 4. William T. Yost and John H. Cantrell Nonlinear acoustical assessment of precipitate nucleation and growth in aluminum alloy 2024, Review of Progress in Quantitative Nondestructive Evaluation, Vol. 19, 2000, p. 1375. 5. William T. Yost and John H. Cantrell Fatigue cycle induced variation of the acoustic nonlinearity parameter in aluminum alloy 2024, Review of Progress in Quantitative Nondestructive Evaluation, Vol. 18, 1999, p. 2345. 6. Jacob Philips and M. A. Breazeale, Physical Acoustics, Vol. XVII (Academic Press, New York, 1984), Ch. 1. 7. Jeong K. Na, John H. Cantrell and William T. Yost Linear and nonlinear ultrasonic properties of fatigued 410Cb stainless steel, Review of Progress in Quantitative Nondestructive Evaluation, Vol. 15, 1996, p. 1347. 8. William T. Yost and John H. Cantrell Calibration techniques for electronic-based systems used in measurement of nonlinearity parameters, Review of Progress in Quantitative Nondestructive Evaluation, Vol. 18, 1999, p. 2345.

23-15

9. Jeong K. Na, John H. Cantrell and William T. Yost Effects of surface roughness and nonparallelism on the measurement of the acoustic nonlinearity parameter in steam turbine blades, Review of Progress in Quantitative Nondestructive Evaluation, Vol. 19, 2000, p. 1417.

23-16

ENVIRONMENTAL FATIGUE I

24
ENVIRONMENTAL EFFECTS ON FATIGUE CRACK INITIATION IN PIPING AND PRESSURE VESSEL STEELS

Omesh K. Chopra Argonne National Laboratory, Argonne, Illinois 60439

24-1

International Conference on Fatigue of Reactor Components July 31 - August 2, 2000 Napa California, USA

ENVIRONMENTAL EFFECTS ON FATIGUE CRACK INITIATION IN PIPING AND PRESSURE VESSEL STEELS
Omesh K. Chopra Argonne National Laboratory, Argonne, Illinois 60439 Phone 630-252-5117

KEYWORDS Fatigue Crack Initiation, Strain vs. Life (SN) Curve, LWR Environment, Carbon Steel, LowAlloy Steel, Austenitic Stainless Steel

ABSTRACT

The ASME Boiler and Pressure Vessel Code provides rules for the construction of nuclear power plant components. Appendix I to Section III of the Code specifies fatigue design curves for structural materials. However, the effects of light water reactor (LWR) coolant environments are not explicitly addressed by the Code design curves. Test data illustrate potentially significant effects of LWR environments on the fatigue resistance of carbon and lowalloy steels and austenitic stainless steels. This paper summarizes the work performed at Argonne National Laboratory on the fatigue of piping and pressure vessel steels in LWR coolant environments. The existing fatigue SN data have been evaluated to establish the effects of various material and loading variables, such as steel type, strain range, strain rate, temperature, and dissolved oxygen level in water, on the fatigue lives of these steels. Statistical models are presented for estimating the fatigue SN curves for carbon and lowalloy steels and austenitic stainless steels as a function of material, loading, and environmental variables. Methods for incorporating environmental effects into the ASME Code fatigue evaluations are discussed. Differences between the methods and their impact on the design fatigue curves are also discussed.

INTRODUCTION

Cyclic loadings on a structural component occur because of changes in mechanical and thermal loadings as the system goes from one load set (e.g., pressure, temperature, moment, and force loading) to any other load set. For each load set, an individual fatigue usage factor is determined by the ratio of the number of cycles anticipated during the lifetime of the component to the allowable cycles. Figures I9.1 through I 9.6 of Appendix I to Section III of the ASME Boiler and Pressure Vessel Code specify design fatigue curves that define the allowable number of cycles as a function of applied stress
24-3

amplitude. The cumulative usage factor (CUF) is the sum of the individual usage factors, and the ASME Code Section III requires that the CUF at each location must not exceed 1. The ASME Code fatigue design curves, given in Appendix I of Section III, are based on straincontrolled tests of small polished specimens at room temperature in air. The fatigue design curves were developed from the bestfit curves of the experimental data by first adjusting for the effects of mean stress on fatigue life and then reducing the fatigue life at each point on the adjusted curve by a factor of 2 on strain or 20 on cycles, whichever was more conservative. As described in the Section III criteria document, these factors were intended to account for data scatter (heattoheat variability), effects of mean stress or loading history, and differences in surface condition and size between the test specimens and actual components. The factors of 2 and 20 are not safety margins but rather conversion factors that must be applied to the experimental data to obtain reasonable estimates of the lives of actual reactor components. However, because the mean fatigue curve used to develop the current Code design curve for austenitic stainless steels (SSs) does not accurately represent the available experimental data (Jaske and ODonnell, 1977; Chopra, 1999), the current Code design curve for SSs includes a reduction of only  DQG  IURP WKH PHDQ FXUYH IRU WKH SS data, not the 2 and 20 originally intended. As explicitly noted in Subsection NB3121 of Section III of the Code, the data on which the design fatigue curves (Figs. I9.1 through I9.6) are based did not include tests in the presence of corrosive environments that might accelerate fatigue failure. Article B2131 in Appendix B to Section III states that the owner's design specifications should provide information about any reduction to design fatigue curves that has been necessitated by environmental conditions. Existing fatigue strainvs.life (SN) data illustrate potentially significant effects of light water reactor (LWR) coolant environments on the fatigue resistance of carbon steels (CSs)

and lowalloy steels (LASs) (Ranganath et al., 1982; Higuchi and Iida, 1991; Nagata et al., 1991; Van Der Sluys, 1993; Kanasaki et al., 1995; Nakao et al., 1995; Higuchi et al., 1997; Chopra and Shack, 1997, 1998a, b, c, 1999) and of austenitic SSs (Fujiwara et al., 1986; Mimaki et al., 1996; Higuchi and Iida, 1997; Kanasaki et al., 1997a, b; Hayashi, 1998; Hayashi et al., 1998; Chopra and Gavenda, 1997, 1998; Chopra and Smith, 1998; Chopra, 1999) (Fig. 1). Under certain environmental and loading conditions, fatigue lives of CSs can be a factor of 70 lower in the environment than in air (Higuchi and Iida, 1991; Chopra and Shack, 1998b). Therefore, the margins in the ASME Code may be less conservative than originally intended. Two approaches have been proposed for incorporating the effects of LWR environments into ASME Section III fatigue evaluations: develop new design fatigue curves for LWR applications, and use a fatigue life correction factor to account for environmental effects. Both approaches are based on the existing fatigue SN data for LWR environments, i.e., the bestfit curves to the experimental fatigue SN data in LWR environments are used to obtain the design curves or fatigue life correction factor. As and when more data became available, the bestfit curves have been modified and updated to include the effects of various material, loading, and environmental parameters on fatigue life. Interim design fatigue curves that address environmental effects on fatigue life of carbon and lowalloy steels and austenitic SSs were first proposed by Majumdar et al. (1993). Design fatigue curves based on a rigorous statistical analysis of the fatigue SN data in LWR environments were developed by Keisler et al. (1995, 1996). Results of the statistical analysis have also been used to estimate the probability of fatigue cracking in reactor components. The Idaho National Engineering Laboratory assessed the significance of the interim fatigue design curves by evaluating samples of components in the reactor coolant pressure boundary (Ware et al., 1995). Six locations were evaluated from facilities designed by each of the four U.S. nuclear steam supply system (NSSS) vendors. Selected components from older vintage plants, designed according to the B31.1 Code, were also included in the evaluation. The design curves and statistical models for estimating fatigue lives in LWR environments have recently been updated for carbon and lowalloy steels (Chopra and Shack, 1998b, c, 1999) and austenitic SSs (Chopra and Smith, 1998; Chopra, 1999). The alternative approach, proposed initially by Higuchi and Iida (1991), considers the effects of reactor coolant environments on fatigue life in terms of a fatigue life correction factor Fen, which is the ratio of the life in air to that in water. To incorporate environmental effects into the ASME Code fatigue evaluations, a fatigue usage for a specific load set, based on the current Code design curves, is multiplied by the correction factor. Specific expressions for Fen, based on the statistical models (Chopra and Shack, 1998b, c, 1999; Chopra, 1999; Mehta and Gosselin, 1996, 1998) and on the correlations developed by the Environmental Fatigue Data
24-4

Committee of the Thermal and Nuclear Power Engineering Society of Japan (Higuchi, 1996), have been proposed. This paper summarizes the data that are available on the effects of various material, loading, and environmental parameters on the fatigue lives of carbon and lowalloy steels and austenitic SSs. The two methods for incorporating the effects of LWR coolant environments into the ASME Code fatigue evaluations are presented. Although estimates of fatigue lives based on the two methods may differ because of differences between the ASME mean curves that were used to develop the current design curves and the bestfit curves to the existing data that were used to develop the environmentally adjusted curves, either method provides an acceptable approach to account for environmental effects.
10.0

Carbon Steel Strain Amplitude, 0a (%)


Temp. (C) DO (ppm) Rate (%/s) S (wt.%) : <150 : 0.05 : 0.4 : 0.006 150250 0.050.2 0.010.4 0.006 >250 >0.2 <0.01 0.006

1.0

Mean Curve RT Air

0.1

ASME Design Curve

10 1

10 2

10 3

10 4

10 5

10 6

Fatigue Life (Cycles)


10.0

0 (%) a

Austenitic Stainless Steels

Temp. (C) : 100200 250325 260325 DO (ppm) : -0.005 -0.005 0.2 Rate (%/s) : -0.01 0.01 0.4

1.0

Mean Curve RT Air

0.1

ASME Design Curve 10 2 10 3 10 4 10 5 10 6

10 1

Fatigue Life (Cycles)

Fig. 1. Fatigue SN data for carbon steels and austenitic stainless steels in water; RT = room temperature

FATIGUE SN DATA IN LWR ENVIRONMENTS Carbon and LowAlloy Steels The fatigue lives of both CSs and LASs are decreased in LWR environments; the reduction depends on temperature, strain rate, dissolved oxygen (DO) level in water, and S content of the steel. Fatigue life is decreased significantly when four conditions are satisfied simultaneously, viz., strain amplitude, temperature, and DO in water are above a minimum level, and strain rate is below a threshold value. The S content in the steel is also important; its effect on life depends on the DO level in water. Although the microstructures and cyclichardening behavior of CSs and LASs differ significantly, environmental degradation of fatigue

life of these steels is very similar. For both steels, only moderate decrease in life (by a factor of <2) is observed when any one of the threshold conditions is not satisfied. The effects of the critical parameters on fatigue life and their threshold values are summarized below. (a) Strain: A minimum threshold strain is required for environmentally assisted decrease in fatigue lives of CSs and LASs (Chopra and Shack, 1998b, c, 1999). The threshold value most likely corresponds to the rupture strain of the surface oxide film. Limited data suggest that the threshold value is limit for the steel. (b) Strain Rate: Environmental effects on fatigue life occur primarily during the tensileloading cycle, and at strain levels greater than the threshold value required to rupture the surface oxide film. When any one of the threshold conditions is not satisfied, e.g., DO <0.05 ppm or temperature <150C, the effects of strain rate are consistent with those in air, i.e., only the heats that are sensitive to strain rate in air show a decrease in life in water. When all other threshold conditions are satisfied, fatigue life decreases logarithmically with decreasing strain rate below 1%/s (Higuchi and Iida, 1991; Katada et al., 1993; Nakao et al., 1995); the effect of environment on life saturates at and Shack, 1998b, c, 1999). The dependence of fatigue life on strain rate for A106Gr B CS and A533Gr B LAS is shown in Fig. 2. For A533Gr B steel, the fatigue life at a strain rate of 0.0004%/s in highDO water ( ppm DO) is more than a factor of 40 lower than that in air. (c) Temperature: When other threshold conditions are satisfied, fatigue life decreases linearly with temperature above 150C and up to 320C (Higuchi and Iida, 1991; Nagata et al., 1991; Nakao et al., 1995). Fatigue life is insensitive to temperatures below 150C or when any other threshold condition is not satisfied. (d) Dissolved Oxygen in Water: When other threshold conditions are satisfied, fatigue life decreases logarithmically with DO above 0.05 ppm; the effect

saturates at ppm DO (Nagata et al., 1991; Nakao et al., 1995). Fatigue life is insensitive to DO level below 0.05 ppm or when any other threshold condition is not satisfied.
A106Gr B Carbon Steel



 KLJKHU WKDQ WKH IDWLJXH

Fatigue Life (Cycles)

10 4

288C, 0 a -0.75%

10 3

10 2 10 -5 10 -4 10 -3 10 -2 Strain Rate (%/s)

Air Simulated PWR -0.7 ppm DO

10 -1

10 0

A533Gr B LowAlloy Steel

Fatigue Life (Cycles)

10 4

288C, 0 a -0.75%

10 3

V &KRSUD

10 2 10 -5 10 -4 10 -3 10 -2 Strain Rate (%/s)

Air Simulated PWR -0.7 ppm DO

10 -1

10 0



Fig. 2. Dependence of fatigue life of carbon and lowalloy steels on strain rate (e) S Content of Steel: The effect of S content on fatigue life depends on the DO content of the water. When the threshold conditions are satisfied and for DO content @1.0 ppm, the fatigue life decreases with increasing S content. Limited data suggest that environmental effects on life saturate at a S content of wt.% (Chopra



Table 1. Fatigue test results for Type 304 austenitic SS at 288C


Pre Dis. Dis. soak xygena Hydrogen Li Boron Test (days) (cc/kg) (ppm) (ppm) No. (ppb) 1805 1808 4 23 2 1000 1 1821 2 23 2 1000 1 1859 2 23 2 1000 1 1861 1 23 1 1862 2 23 5 1863 1 5 aDO and ECPs measured in effluent. bConductivity of water measured in feedwater supply tank. Conduc -tivityb pH at RT (2S/cm) 6.4 18.87 6.5 22.22 6.5 18.69 6.2 0.06 6.2 0.06 6.3 0.06 ECPa teel mV (SHE) 686 693 692 610 603 520 Ten. Rate (%/s) 4.0E-3 4.0E-3 4.0E-3 4.0E-3 4.0E-3 4.0E-3 4.0E-3 Stress Range (MPa) 467.9 468.3 474.3 471.7 463.0 466.1 476.5 Strain Range (%) 0.76 0.77 0.76 0.77 0.79 0.78 0.77 Life N25 (Cycles) 14,410 2,850 2,420 2,420 2,620 2,450 2,250

24-5

and Shack, 1998b). At high DO levels, e.g., >1.0 ppm, fatigue life seems to be insensitive to S content in the range of 0.0020.015 wt.% (Higuchi, 1995). When any one of the threshold conditions is not satisfied, environmental effects on life are minimal and relatively insensitive to changes in S content.

life saturates at SSs (Chopra and Smith, 1998; Kanasaki et al., 1997b). Existing data are too sparse to define the saturation strain rate for cast austenitic SSs.
Type 304 SS 288C, 0 -0.75% Strain Rate 0.004/0.4 %/s DO -0.84 ppm

V IRU ZURXJKW

Austenitic Stainless Steels The fatigue lives of austenitic SSs are decreased in LWR environments; the reduction depends on strain rate, level of DO in water, and temperature (Chopra and Gavenda, 1997, 1998; Chopra and Smith, 1998; Kanasaki et al., 1997a). The effects of LWR environments on fatigue life of wrought materials are comparable for Types 304, 316, and 316NG SS. Although the fatigue lives of cast SSs are relatively insensitive to changes in ferrite content in the range of 12 to 28% (Kanasaki et al., 1997a), the effects of loading and environmental parameters on the fatigue life of cast SSs differ somewhat. The significant results and threshold values of critical parameters are summarized below. (a) Strain: A minimum threshold strain is required for environmentally assisted decrease in fatigue life of austenitic SSs. Limited data suggest that the threshold strain range is 0.32 to 0.36% (Chopra and Smith, 1998; Kanasaki et al., 1997b). (b) Dissolved Oxygen in Water: For wrought austenitic SSs, environmental effects on fatigue life are more pronounced in lowDO, i.e., <0.01 ppm DO, than in highDO, i.e., J0.1 ppm DO, water (Chopra and Smith, 1998; Kanasaki et al., 1997a). In highDO water, environmental effects are moderate (less than a factor of 2 decrease in life) when conductivity is maintained at <0.1 2S/cm and electrochemical potential (ECP) of the steel has reached a stable value (Fig. 3). For fatigue tests in highDO water, the SS specimens must be soaked for 5 to 6 days for the ECP of the steel to reach a stable value. Figure 3 shows that although fatigue life is decreased by a factor of is increased from 2S/cm, the period for presoaking appears to have a larger effect on life than the conductivity of water. In lowDO water, the additions of Li and B, or low conductivity, or preexposing the specimen for dissolved H, have no effect on fatigue life of Type 304 SS (Table 1). Also, for cast austenitic SSs, the effect of DO content is somewhat different; the fatigue lives are approximately the same in both high or lowDO water and are comparable to those observed for wrought SSs in lowDO water (Chopra and Smith, 1998). (c) In highDO water (conductivity Strain Rate: <0.1 2S/cm and stable ECP of the steel), fatigue life is insensitive to changes in strain rate. In lowDO water, fatigue life decreases logarithmically with decreasing strain rate below ; the effect of environment on

Fatigue Life (Cycles)

10 4

ECP Steel Electrode mV(SHE) Open Symbols: 145165 (-120 h soak) Closed Symbols: 30145 (-20 h soak)

10 3 10 -2

10 -1 Conductivity of Water (2 S/cm)

10 0

Fig. 3. Effects of conductivity of water and soak period on fatigue life of Type 304 SS in highDO water (d) Temperature: Existing data are too sparse to establish the effects of temperature on fatigue life over the entire range from room temperature to reactor operating temperatures. Limited data indicate that environmental effects on fatigue life are minimal below 200C and significant at temperatures above 250C (Kanasaki et al., 1997b); life appears to be relatively insensitive to changes in temperature in the range of 250330C. The pressure vessel research council (PVRC) steering committee for cyclic life and environmental effects (CLEE) has proposed a ramp function to describe temperature effects on the fatigue lives of austenitic SSs; environmental effects are moderate at temperatures below 180C, significant above 220C, and increase linearly from 180 to 220C (Yukawa, 1999).

 ZKHQ FRQGXFWLYLW\ RI ZDWHU  WR 

 GD\V EHIRUH WKH WHVW RU

V

OPERATING EXPERIENCE IN NUCLEAR POWER INDUSTRY Experience with operating nuclear power plants worldwide reveals that many failures may be attributed to fatigue; examples include piping components, nozzles, valves, and pumps (Kussmaul et al., 1983; Iida, 1992). In most cases, these failures have been associated with thermal loading due to thermal stratification and striping, or mechanical loading due to vibratory loading. Significant thermal loadings due to flow stratification were not included in the original design basis analysis. The effect of these loadings may also have been aggravated by corrosion effects due to a hightemperature aqueous environment. A review of significant occurrences of corrosion fatigue damage and failures in various nuclear power plant systems has been presented in an EPRI report (Dooley and Pathania, 1997); the results are summarized below. Cracking in Feedwater Nozzle and Piping Fatigue cracks have been observed in feedwater piping and nozzles of the pressure vessel in boiling water reactors (BWRs) and steam generators in pressurized water reactors
24-6

(PWRs) (Kussmaul et al., 1984; NRC 1979, 1993). The mechanism of cracking has been attributed to corrosion fatigue (Watanabe, 1980; Gordon et al., 1987) or straininduced corrosion cracking (SICC) (Lenz et al., 1983). Case histories and identification of conditions that lead to SICC of LASs in LWR systems have been summarized by Hickling and Blind (1986). In BWR nozzle cracking, initiation has been attributed to highcycle fatigue caused by the leakage of cold water around the thermal sleeve junction area, and crack propagation has been attributed to lowcycle fatigue due to plant transients such as startup/shutdowns and any feedwater on/off transients. The frequency of the highcycle fatigue phenomenon due to leakage around the sleeve is  Hz and therefore is not expected to be influenced by the reactor coolant environment. Estimates of strain range and strain rates for typical transients associated with lowcycle fatigue are given in Table 2 (Ford et al., 1993). Under these loading and environmental conditions, significant reduction in fatigue life has been observed for carbon and lowalloy steels (Chopra and Shack, 1998b, 1999). Table 2. Typical chemical and cyclic strain transients
Component FW Nozzle FW Piping FW Piping FW Piping FW Piping FW Piping FW Piping Operation Startup Startup Startup Turbine Roll Hot Standby Cool Down Stratification DO (ppb) 20/200 20/200 20/200 <200 <200 <20 200 Temp. Strain Strain (C) Range (%) Rate (%/s) 216/38 0.2-0.4 102 216/38 0.2-0.5 103102 288/38 0.07-0.1 48x106 288/80 288/90 288/RT 250/50 0.4 0.26 0.2 0.2-0.7 36x103 4x104 6x104 104103

example, the deepest crack was straight, nonbranching, transgranular through both the ferrite and pearlite regions without any preference, and showed considerable oxidation and some pitting at the crack origin. In fatigue test specimens, nearsurface cracks grow entirely as tensile cracks normal to the stress and across both the soft ferrite and hard pearlite regions, whereas in air, cracks grow at an angle of 45 to the stress axis and only along the ferrite regions (see Fig. 4 of a companion paper at this conference by Chopra and Park). The identical crack morphologies indicate that environment played a dominant role in crack initiation. Similar characteristics of transgranular crack propagation through both weld and base metal, without regard to microstructural features, have also been identified in German reactors (Hickling and Blind, 1986). Components tests have also been conducted to validate the calculation procedures and the applicability of the test results from specimen to actual reactor component. Tests on pipes, plates, and nozzles, under cyclic thermal loading in aqueous environment (Kussmaul et al., 1983) indicate that crack initiation in simulated LWR environments may occur earlier than the values of the ASME Section III fatigue design curve; environmental effects are more pronounced in the ferritic steel than in the austenitic cladding. Tests at the HDRfacility (Katzenmeier et al., 1990) have also shown good agreement between the fatigue lives applicable to specimens and components, e.g., first incipient crack on pipes appeared in 1200 cycles, compared with 1400 cycles for a test specimen made of the same material and tested under comparable conditions (8 ppm DO). Safety Injection System and Pressurizer Surge Line Significant cracking has also occurred in unisolable pipe sections in the safety injection system piping connected to the PWR coolant system (NRC 1988a, b). This phenomenon, which is similar to the nozzle cracking discussed above, is caused by thermal stratification. Also, regulatory evaluation has indicated that thermal stratification can occur in all PWR surge lines (NRC 1988c). In PWRs, the pressurizer water is heated to & )  7KH KRW ZDWHU IORZLQJ DW D YHU\ slow rate from the pressurizer through the surge line to the hotleg piping, rides on a cooler water layer. The thermal gradients between the upper and lower parts of the pipe can be as high as 149C (300F). Fullscale mock-up tests to generate thermal stratification in a pipe in a laboratory have confirmed the applicability of laboratory data to component behavior (Lenz et al., 1990). The material, loading, and environmental conditions were simulated on a 1:1 scale, using only thermohydraulic effects. Under the loading conditions, i.e., strain rate and strain range typical of thermal stratification in these piping systems, the coolant environment is known to have a significant effect on fatigue crack initiation (Chopra, 1999; Kanasaki et al., 1997a, b).

In PWR feedwater pipe cracking, cracking has been attributed to a combination of thermal stratification and thermal striping (Dooley and Pathania, 1997). Environmental factors, such as high DO in the feedwater, are believed to also have played a significant role in crack initiation. The thermal stratification is caused by the injection of lowflow, relatively cold feedwater during plant startup, hot standby, and variations below 20% of full power, whereas thermal striping is caused by rapid, localized fluctuations of the interface between hot and cold feedwater. Lenz et al. (1983) showed that in feedwater lines, the strain rates are 103105%/s due to thermal stratification and 101%/s due to thermal shock and that thermal stratification is the primary cause of crack initiation due to SICC. Also, the results from smallsize specimens, mediumsize components (model vessels), and fullsize thermalshock experiments suggest an influence of oxygen content in pressurized water on crack initiation behavior (Kussmaul et al., 1984). A detailed examination of cracking in a CS elbow adjacent to the steam generator nozzle weld (Enrietto et al., 1981) indicates crack morphologies that are identical to those observed in smooth specimens tested in highDO water. For
24-7

Steam Generator Girth Weld cracking Another instance of thermalfatigueinduced cracking where environmental effects are believed to have played a role in crack initiation has been observed at the weld joint between the two shells of a steam generator (Foley et al., 1991). The feedwater temperature in this region is nominally 204227C (440440F), compared with the steam generator temperature of 288C (550C). The primary mechanism of cracking has been considered corrosion fatigue with possible slow crack growth due to stress corrosion cracking. A detailed analysis of girthweld cracking indicates that crack initiation was dominated by environmental influences, particularly under relatively highDO content and/or oxidizing potential (Bamford et al., 1991).

effects are not significant. The design fatigue curves are then obtained by lowering the adjusted bestfit curve by a factor of 2 on stress or 20 on cycles, whichever is more conservative, to account for differences and uncertainties in fatigue life that are associated with material and loading conditions. Statistical models based on the existing fatigue SN data have been developed for estimating the fatigue lives of pressure vessel and piping steels in air and LWR environments (Chopra and Shack, 1998b, 1999; Chopra, 1999; Chopra and Smith, 1998). In air at room temperature, the fatigue life N of CSs is represented by ln(N) = 6.564 1.975 ln(0a 0.113), and of LASs by ln(N) = 6.627 1.808 ln(0a 0.151), (2b) where 0a is applied strain amplitude (%). In LWR environments, the fatigue life of CSs is represented by
ln(N) = 6.010 1.975 ln(0a 0.113) + 0.101 S* T* O* H *,(3a)

(2a)

INCORPORATING ENVIRONMENTAL EFFECTS INTO ASME FATIGUE EVALUATIONS Two procedures are currently being proposed for incorporating effects of LWR coolant environments into the ASME Section III fatigue evaluations; develop a new set of environmentally adjusted design fatigue curves (Chopra and Shack, 1998b, 1999; Chopra, 1999; Chopra and Smith 1998) or use fatigue life correction factors Fen to adjust the current ASME Code fatigue usage values for environmental effects (Chopra and Shack, 1999; Chopra, 1999; Mehta and Gosselin, 1996, 1998). For both approaches, the range and bounding values must be defined for key service parameters that influence fatigue life. It has been demonstrated that estimates of fatigue lives based on the two methods may differ because of differences between the ASME mean curves used to develop the current design curves and the bestfit curves to the existing data used to develop the environmentally adjusted curves. However, either of these methods provides an acceptable approach to account for environmental effects. Design Fatigue Curves A set of environmentally adjusted design fatigue curves can be developed from the bestfit stressvs.life curves to the experimental data in LWR environments by using the same procedure that has been used to develop the current ASME Code design fatigue curves. The stressvs.life curves are obtained from the strainvs.life curves, e.g., stress amplitude is the product of strain amplitude and elastic modulus. The bestfit experimental curves are first adjusted for the effect of mean stress by using the modified Goodman relationships V u V y for Sa < 8 y , (1a) Sc Sa a V u  Sa and Sc = Sa a for Sa > 8 y , (1b) where Sc is the adjusted value of stress amplitude Sa, and 8 y a and V u are yield and ultimate strengths of the material, respectively. Equations 1a and 1b assume the maximum possible mean stress and typically yield a conservative adjustment for mean stress, at least when environmental
24-8

and of LASs by
ln(N) = 5.729 1.808 ln(0a 0.151) + 0.101 S* T* O* H *,(3b) where S*, T*, O*, and H * are transformed S, temperature, DO, and strain rate, respectively, defined as follows:

(DO > 1.0 ppm) S* = 0.015 (DO @ 1.0 ppm & 0 < S @ 0.015 wt.%) S* = S (DO @ 1.0 ppm & S > 0.015 wt.%) (4a) S* = 0.015 *=0 (T < 150C) T (T = 150350C) (4b) T* = T 150 *=0 (DO < 0.05 ppm) O O* = ln(DO/0.04) (0.05 ppm @ DO @ 0.5 ppm) (DO > 0.5 ppm) (4c) O* = ln(12.5) * = 0 > 1%/s) (H H (0.001 @ H @ 1%/s) H * = ln( H ) * = ln(0.001) < 0.001%/s). (H (4d) H The discontinuity in the value of O* at 0.05 ppm DO is due to an approximation and does not represent a physical phenomenon. In air at room temperature, the fatigue data for Types 304 and 316 SS are best represented by Eq. 5a ln(N) = 6.703 2.030 ln(0a 0.126), and for Type 316NG, by Eq. 5b (5a)

ln(N) = 7.422 1.671 ln(0a 0.126).


Carbon Steel RoomTemp. Air

(5b)

0'

8u = 551.6 MPa 8y = 275.8 MPa

10 3

E = 206.84 GPa
2

10

Design Curve Based on Statistical Model ASME Code Curve

10 1

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

a (MPa)

LowAlloy Steel RoomTemp. Air

8u = 689.5 MPa 8y = 482.6 MPa


Stress Amplitude, S

10 3

E = 206.84 GPa Design Curve Based on Statistical Model ASME Code Curve

10

10 1

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

(7b) O' = 0.260 (DO < 0.05 ppm) O' = 0 (DO J 0.05 ppm). (7c) The models are recommended for predicted fatigue lives of @106 cycles. The design fatigue curves were obtained from the bestfit curves, represented by Eqs. 2a3b for CSs and LASs, and by Eqs. 5a and 6a for austenitic SSs. To be consistent with the current Code design curves, the meanstressadjusted bestfit curves were decreased by the same margins on stress and cycles that are imposed in the current Code curves, e.g., the adjusted bestfit curves were decreased by a factor of 2 on stress for CSs and LASs and by a factor of 1.5 for austenitic SSs. A factor of 20 on life was used for all of the curves, although the actual margin on life is 1016 for SSs because of the differences between the ASME mean curve and the bestfit curve to existing fatigue data. The new design fatigue curves for CSs and LASs and austenitic SS in air are shown in Fig. 4, whereas those in LWR coolant environments are shown in Figs. 57; only the portions of the environmentally adjusted curves that fall below the current ASME Code curve are shown in Figs. 57. Because the fatigue life of Type 316NG is superior to that of Types 304 or
Carbon Steel Water
When any one of the following conditions is true: Temp. <150C DO <0.05 ppm Strain Rate 1%/s

=0 ( 0 > 0.4%/s) ' = ln( 0 /0.4) (0.0004 @ 0 @ 0.4%/s) 0 ' = ln(0.0004/0.4) ( 0 < 0.0004%/s) 0

Stress Amplitude, S (MPa) a

Stress Amplitude, S (MPa) a

Austenitic Stainless Steel Room Temp. Air

8u = 648.1 MPa 8y = 303.4 MPa

10 3

(MPa)

10 3 Stress Amplitude, S

E = 195.1 GPa Statistical Model ASME Code Curve

10 2 10 1

10 2

Statistical Model ASME Code Curve

10 2

10 3 10 4 10 5 Number of Cycles, N

10 6

10 7

10 1

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

Stress Amplitude, S a (MPa)

Fig. 4. Design fatigue curve developed from statistical model in air at room temperature In LWR environments, the fatigue data for Types 304 and 316 SS are best represented by
ln(N) = 5.768 2.030 ln(0a 0.126) + T' 0 ' O',

LowAlloy Steel Water


When any one of the following conditions is true: Temp. <150C DO <0.05 ppm Strain Rate 1%/s

10 3

(6a)

and for Type 316NG, by


ln(N) = 6.913 1.671 ln(0a 0.126) + T' 0 ' O',

(6b)

10 2

Statistical Model ASME Code Curve

where T', 0 ', and O' are transformed temperature, strain rate, and DO, respectively, defined as follows:

10 1

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

T' = 0 T' = (T 180)/40 T' = 1

(T < 180C) (180 @ T < 220C) (T J 220C)

(7a)

Fig. 5. Design fatigue curves developed from statistical model under service conditions where one or more critical threshold values are not satisfied
24-9

Stress Amplitude, S (MPa) a

Stress Amplitude, Sa (MPa)

Carbon Steel Water


3

LowAlloy Steel Water


3

10

Temp. 200C DO 0.2 ppm Sulfur 0.015 wt.%

10

Temp. 200C DO 0.2 ppm Sulfur 0.015 wt.%

10 2 10 1

Strain Rate (%/s) 0.1 0.01 0.001 ASME Code Curve

Strain Rate (%/s) 0.1 0.01 0.001 ASME Code Curve

10 2 10 1

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

Stress Amplitude, S (MPa) a

10

Temp. 250C DO 0.2 ppm Sulfur 0.015 wt.%

Stress Amplitude, Sa (MPa)

Carbon Steel Water

LowAlloy Steel Water


Temp. 250C DO 0.2 ppm Sulfur 0.015 wt.%

10 3

10 2 10 1

Strain Rate (%/s) 0.1 0.01 0.001 ASME Code Curve

Strain Rate (%/s) 0.1 0.01 0.001 ASME Code Curve

10 10 5 10 6

10 2

10 3 10 4 Number of Cycles, N

10 1

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

Stress Amplitude, Sa (MPa)

a (MPa)

Carbon Steel Water


Temp. 288C DO 0.2 ppm Sulfur 0.015 wt.%

LowAlloy Steel Water


Temp. 288C DO 0.2 ppm Sulfur 0.015 wt.%

10 3 Stress Amplitude, S

10 3

Strain Rate (%/s) 0.1 0.01 0.001 ASME Code Curve

10 2

10 2
10 5 10 6

Strain Rate (%/s) 0.1 0.01 0.001 ASME Code Curve

10 1

10 2

10 3 10 4 Number of Cycles, N

10 1

10 2

10 3 10 4 Number of Cycles, N

10 5

10 6

Fig. 6. Design fatigue curves developed from statistical model for carbon and lowalloy steels under service conditions where all critical threshold values are satisfied
DO <0.05 ppm Stress Amplitude Sa (MPa)
<180C, All Strain Rates or 220C, 0.4%/s

DO 0.05 ppm Stress Amplitude Sa (MPa)

All Temperatures & Strain Rates ASME Code Design Curve

10 3
ASME Code Design Curve 200C Strain Rate (%/s) 0.04 0.004 0.0004

10 3

10 2 10 1

10 2 10 6 10 7 10 1 10 2 10 3 10 4 10 5 Number of Cycles N 10 6 10 7

10 2

10 3 10 4 10 5 Number of Cycles N

Fig. 7. Design fatigue curves developed from statistical models for Types 304 and 316 SS in water with <0.05 and J 0.05 ppm DO

24-10

316 SS, the design curves in Figs. 4 and 7 will be somewhat conservative for Type 316NG SS. For carbon and lowalloy steels, a set of design curves similar to those shown in Fig. 6 can be developed for lowS steels, i.e., steels with @ 0.007 wt.% S. The results indicate that in room temperature air, the current ASME Code design curves for CSs and LASs are conservative with respect to the curves based on the statistical models, and those for austenitic SSs are nonconservative at stress levels above 300 MPa. For environmentally adjusted design fatigue curves (Figs. 57), we define a minimum threshold strain, below which environmental effects are modest. The threshold strain for CSs and LASs appears to be limit of the steel. This translates into strain amplitudes of 0.140 and 0.185%, respectively, for CSs and LASs. These values must be adjusted for mean stress effects and variability due to material and experimental scatter. The threshold strain amplitudes are decreased by CSs and by LASs to account for the effects of mean stress, and by a factor of 1.7 on strain to provide 90% confidence for the variations in fatigue life that are associated with material variability and experimental scatter (Keisler et al., 1995). Thus, a threshold strain amplitude of 0.07% (or a stress amplitude of 145 MPa) is obtained for both CSs and LASs. The existing fatigue data indicate a threshold strain range of SSs. This value is decreased by effects and by a factor of 1.5 to account for uncertainties in fatigue life that are associated with material and loading variability. Thus, a threshold strain amplitude of 0.097% (stress amplitude of 189 MPa) is obtained for austenitic SSs. The PVRC steering committee for CLEE (Yukawa, 1999) has proposed a ramp for the threshold strain; a lower strain amplitude below which environmental effects are insignificant, a slightly higher strain amplitude above which environmental effects decrease fatigue life, and a ramp between the two values. The two strain amplitudes are 0.07 and 0.08% for carbon and lowalloy steels, and 0.10 and 0.11% for austenitic SSs (both wrought and cast SS). These threshold values have been used to develop Figs. 6 and 7.

Fen = exp(0.898 0.101 S* T* O* 0 *),

(9b) (9c)

and for austenitic SSs, by


Fen = exp(0.935 T' 0 ' O'),

 KLJKHU WKDQ WKH IDWLJXH

 IRU

 IRU

 IRU DXVWHQLWLF  WR DFFRXQW IRU PHDQ VWUHVV

where the constants and are defined in Eqs. 4a 4d, and T', 0 ' and O' are defined in Eqs. 7a7c. A strain threshold is also defined, below which environmental effects are modest. The strain threshold is represented by a ramp, i.e., a lower strain amplitude below which environmental effects are insignificant, a slightly higher strain amplitude above which environmental effects are significant, and a ramp between the two values. Thus, the negative terms in Eqs. 9a 9c are scaled from zero to their actual value between the two strain thresholds. The two strain amplitudes are 0.07 and 0.08% for CSs and LASs, respectively, and 0.10 and 0.11% for austenitic SSs (both wrought and cast SS). To incorporate environmental effects into the Section III fatigue evaluation, a fatigue usage for a specific stress cycle, based on the current Code design fatigue curve, is multiplied by the correction factor. The experimental data adjusted for environmental effects, i.e., the product of experimentally observed fatigue life in LWR environments and Fen, are presented with the bestfit SN curve in roomtemperature air in Fig. 8. A similar approach has been proposed by Mehta and Gosselin (1996, 1998); however, they defined Fen as the ratio of the life in air to that in water, both at service temperature. The Fen approach, also known as the EPRI/GE approach, has recently been updated to include the revised statistical models and the PVRC discussions on environmental fatigue evaluations (Mehta, 1999). An effective fatigue life correction factor, expressed as Fen,eff = Fen/Z, is defined where Z is a factor that constitutes the perceived conservatism in the ASME Code design curves. The Fen,eff approach presumes that all uncertainties have been anticipated and accounted for.

S*,

T* , 0 *

O*

Fatigue Life Correction Factor The effects of reactor coolant environments on fatigue life have also been expressed in terms of a fatigue life correction factor Fen, which is the ratio of life in air at room temperature to that in water at the service temperature (Higuchi and Iida, 1991). A fatigue life correction factor Fen can be obtained from the statistical model (Eqs. 27), where ln(Fen) = ln(NRTair) ln(Nwater). The fatigue life correction factor for CSs is given by
Fen = exp(0.554 0.101 S* T* O* 0 *),

(8) (9a)

CONCLUSIONS The work performed at Argonne National Laboratory on fatigue of carbon and lowalloy steels in LWR environments is summarized. The existing fatigue SN data have been evaluated to establish the effects of various material and loading variables such as steel type, strain range, strain rate, temperature, sulfur content in steel, orientation, and DO level in water on the fatigue life of these steels. Statistical models are presented for estimating the fatigue SN curves as a function of material, loading, and environmental variables. Case studies of fatigue failures in nuclear power plants are presented and the contribution of environmental effects on crack initiation is discussed.

for LASs, by
24-11

Carbon Steels
Strain Amplitude, 0a (%) Statistical Model Room Temp. Air 1.0

0.1

102

103 104 105 106 Adjusted Fatigue Life, Fen x N25 (Cycles)

107

LowAlloy Steels
0a(%)

Statistical Model Room Temp. Air 1.0 Strain Amplitude,

0.1

10 2

10 3 10 4 10 5 106 Adjusted Fatigue Life, Fen x N 25 (Cycles)

107

The environmentally adjusted design fatigue curves provide allowable cycles for fatigue crack initiation in LWR coolant environments. The new design curves maintain the margin of 20 on life. However, to be consistent with the current ASME Code curves, the margin on stress is 2 for carbon and lowalloy steels and 1.5 for austenitic SSs. In the Fen method, environmental effects on life are estimated from the statistical models but the correction is applied to fatigue lives estimated from the current Code design curves. Therefore, estimates of fatigue lives that are based on the two methods may differ because of differences in the ASME mean curve and the bestfit curve to existing fatigue data. The current Code design curve for CSs is comparable to the statisticalmodel curve for LASs, whereas it is somewhat conservative at stress levels of <500 MPa when compared with the statisticalmodel curve for CSs. Consequently, usage factors based on the Fen method would be comparable to those based on the environmentally adjusted design fatigue curves for LASs and would be somewhat higher for CSs. Figure 4 indicates that for austenitic SSs, the current Code design fatigue curve is nonconservative when compared with the statisticalmodel curve, i.e., it predicts longer fatigue lives than the bestfit curve to the existing SN data. Therefore, usage factors that are based on the Fen method would be lower than those determined from the environmentally corrected design fatigue curves.

Austenitic Stainless Steels


Strain Amplitude, 0a (%) Statistical Model Room Temp. Air 1.0

ACKNOWLEDGMENTS This work was sponsored by the Office of Nuclear Regulatory Research, U.S. Nuclear Regulatory Commission, Job Code W6610. REFERENCES Bamford, W. H., Rao, G. V., and Houtman, J. L., 1992, Investigation of Service Induced Degradation of Steam Generator Shell, Proc. 5th Intl. Symp. on Environmental Degradation of Materials in Nuclear Power Systems Water Reactors, American Nuclear Society, La Grange Park, IL. Chopra, O. K., 1999, Effects of LWR Coolant Environments on Fatigue Design Curves of Austenitic Stainless Steels, NUREG/CR5704, ANL98/31. Chopra, O. K., and Gavenda, D. J., 1997, Effects of LWR Coolant Environments on Fatigue Lives of Austenitic Stainless Steels, Pressure Vessel and Piping Codes and Standards, PVP Vol. 353, D. P. Jones, B. R. Newton, W. J. O'Donnell, R. Vecchio, G. A. Antaki, D. Bhavani, N. G. Cofie, and G. L. Hollinger, eds., American Society of Mechanical Engineers, New York, pp. 8797. Chopra, O. K., and Gavenda, D., J., 1998, Effects of LWR Coolant Environments on Fatigue Lives of Austenitic Stainless Steels, J. Pressure Vessel Technol. 120, pp. 116121.

0.1

102

103 104 105 106 Adjusted Fatigue Life, Fen x N25 (Cycles)

107

Fig. 8. Comparison of experimental data adjusted for environmental effects with bestfit fatigue SN curve in roomtemperature air The current two methods for incorporating the effects of LWR coolant environments into the ASME Code fatigue evaluations, i.e., the design fatigue curve method and the fatigue life correction factor method, are presented. Both methods are based on the statistical models for estimating fatigue lives of carbon and lowalloy steels and austenitic SSs in LWR environments. Although estimates of fatigue lives based on the two methods may differ because of differences between the ASME mean curves used to develop the current design curves and the bestfit curves to the existing data used to develop the environmentally adjusted curves, either of these methods provides an acceptable approach to account for environmental effects.
24-12

Chopra, O. K., and Shack, W. J., 1997, Evaluation of Effects of LWR Coolant Environments on Fatigue Life of Carbon and LowAlloy Steels, Effects of the Environment on the Initiation of Crack Growth, ASTM STP 1298, W. A. Van Der Sluys, R. S. Piascik, and R. Zawierucha, eds., American Society for Testing and Materials, Philadelphia, pp. 247266. Chopra, O. K., and Shack, W. J., 1998a, LowCycle Fatigue of Piping and Pressure Vessel Steels in LWR Environments, Nucl. Eng. Des. 184, pp. 4976. Chopra, O. K., and Shack, W. J., 1998b, Effects of LWR Coolant Environments on Fatigue Design Curves of Carbon and LowAlloy Steels, NUREG/CR6583, ANL 97/18. Chopra, O. K., and Shack, W. J., 1998c, Fatigue Crack Initiation in Carbon and LowAlloy Steels in Light Water Reactor Environments Mechanism and Prediction, Fatigue, Environmental Factors, and New Materials, PVP Vol. 374, H. S. Mehta, R. W. Swindeman, J. A. Todd, S. Yukawa, M. Zako, W. H. Bamford, M. Higuchi, E. Jones, H. Nickel, and S. Rahman, eds., American Society of Mechanical Engineers, New York, pp. 155168. Chopra, O. K., and Shack, W. J., 1999, Overview of Fatigue Crack Initiation in Carbon and LowAlloy Steels in Light Water Reactor Environments, J. Pressure Vessel Technol., in press. Chopra, O. K., and Smith, J., L., 1998, Estimation of Fatigue StrainLife Curves for Austenitic Stainless Steels in Light Water Reactor Environments, Fatigue, Environmental Factors, and New Materials, PVP Vol. 374, H. S. Mehta, R. W. Swindeman, J. A. Todd, S. Yukawa, M. Zako, W. H. Bamford, M. Higuchi, E. Jones, H. Nickel, and S. Rahman, eds., American Society of Mechanical Engineers, New York, pp. 249259. Dooley, R. B., and Pathania, R. S., 1997, Corrosion Fatigue of Water Touched Pressure Retaining Components in Power Plants, EPRI TR106696, Electric Power Research Institute, Palo Alto, CA. Enrietto, J. F., Bamford, W. H., and White, D. F., 1981, Preliminary Investigation of PWR Feedwater Nozzle Cracking, Intl. J. Pressure Vessels and Piping, 9, pp. 421443. Foley, W. J., Dean, R. S., and Hennick, A., 1991, Closeout of IE Bulletin 7913: Cracking in Feedwater System Piping, NUREG/CR5258, US Nuclear Regulatory Commission, Washington, DC. Ford, F. P., 1986, Overview of Collaborative Research into the Mechanisms of Environmentally Controlled Cracking in the Low Alloy Pressure Vessel Steel/Water System, Proc. 2nd Intl. Atomic Energy Agency Specialists' Meeting on Subcritical Crack Growth, NUREG/CP0067, MEA2090, Vol. 2, pp. 371. Ford, F. P., Ranganath, S., and Weinstein, D., 1993, Environmentally Assisted Fatigue Crack Initiation in LowAlloy Steels A Review of the Literature and the
24-13

ASME Code Requirements, EPRI TR102765,. Electric Power Research Institute, Palo Alto, CA. Fujiwara, M., Endo, T., and Kanasaki, H., 1986, Strain Rate Effects on the Low Cycle Fatigue Strength of 304 Stainless Steel in High Temperature Water Environment, Fatigue Life: Analysis and Prediction, Proc. of the Intl. Conf. and Exposition on Fatigue, Corrosion Cracking, Fracture Mechanics, and Failure Analysis, ASM, Metals Park, OH, pp. 309313. Gavenda, D. J., Luebbers, P. R., and Chopra, O. K., 1997, Crack Initiation and Crack Growth Behavior of Carbon and LowAlloy Steels, Fatigue and Fracture 1, Vol. 350, S. Rahman, K. K. Yoon, S. Bhandari, R. Warke, and J. M. Bloom, eds., American Society of Mechanical Engineers, New York, pp. 243255. Gordon, B. M., Delwiche, D. E., and Gordon, G. M., 1987, Service Experience of BWR Pressure Vessels, Performance and Evaluation of Light Water Reactor Pressure Vessels, PVP Vol.119, American Society of Mechanical Engineers, New York, pp. 917. Hayashi, M., 1998, Thermal Fatigue Strength of Type 304 Stainless Steel in Simulated BWR Environment, Nucl. Eng. Des. 184, pp. 135144. Hayashi, M., Enomoto, K., Saito, T., and Miyagawa, T., 1998, Development of Thermal Fatigue Testing with BWR Water Environment and Thermal Fatigue Strength of Austenitic Stainless Steels, Nucl. Eng. Des. 184, pp. 113122. Hickling, J. and Blind, D., 1986, StrainInduced Corrosion Cracking of LowAlloy Steels in LWR Systems Case Histories and Identification of Conditions Leading to Susceptibility, Nucl. Eng. Des. 91, pp. 305330. Higuchi, M., 1995, presented at Working Group Meeting on SN Data Analysis, the Pressure Vessel Research Council, June, Milwaukee. Higuchi, M., 1996, presented at Working Group Meeting on SN Data Analysis, the Pressure Vessel Research Council, April, Orlando, FL. Higuchi, M., and Iida, K., 1991, Fatigue Strength Correction Factors for Carbon and LowAlloy Steels in Oxygen Containing HighTemperature Water, Nucl. Eng. Des. 129, pp. 293306. Higuchi, M., and Iida, K., 1997, Reduction in LowCycle Fatigue Life of Austenitic Stainless Steels in High Temperature Water, Pressure Vessel and Piping Codes and Standards, PVP Vol. 353, D. P. Jones, B. R. Newton, W. J. O'Donnell, R. Vecchio, G. A. Antaki, D. Bhavani, N. G. Cofie, and G. L. Hollinger, eds., American Society of Mechanical Engineers, New York, pp. 7985. Higuchi, M., Iida, K., and Asada, Y., 1997, Effects of Strain Rate Change on Fatigue Life of Carbon Steel in High Temperature Water, Effects of the Environment on the Initiation of Crack Growth, ASTM STP 1298, W. A. Van Der Sluys, R. S. Piascik, and R. Zawierucha, eds.,

American Society for Testing and Materials, Philadelphia, pp. 216231. Iida, K., 1992, A Review of Fatigue Failures in LWR Plants in Japan, Nucl. Eng. Des. 138, pp. 297312. Jaske, C. E., and ODonnell, W., J., 1977, Fatigue Design Criteria for Pressure Vessel Alloys, Trans. ASME J. Pressure Vessel Technol. 99, pp. 584592. Kanasaki, H., Hayashi, M., Iida, K., and Asada, Y., 1995, Effects of Temperature Change on Fatigue Life of Carbon Steel in High Temperature Water, Fatigue and Crack Growth: Environmental Effects, Modeling Studies, and Design Considerations, PVP Vol. 306, S. Yukawa, ed., American Society of Mechanical Engineers, New York, pp. 117122. Kanasaki, H., Umehara, R., Mizuta, H., and Suyama, T., 1997a, Fatigue Lives of Stainless Steels in PWR Primary Water, Trans. 14th Intl. Conf. on Structural Mechanics in Reactor Technology (SMiRT 14), Lyon, France, pp. 473483. Kanasaki, H., Umehara, R., Mizuta, H., and Suyama, T., 1997b, Effects of Strain Rate and Temperature Change on the Fatigue Life of Stainless Steel in PWR Primary Water, Trans. 14th Intl. Conf. on Structural Mechanics in Reactor Technology (SMiRT 14), Lyon, France, pp. 485493. Katada, Y., Nagata, N., and Sato, S., 1993, Effect of Dissolved Oxygen Concentration on Fatigue Crack Growth Behavior of A533 B Steel in HighTemperature Water, ISIJ Intl. 33 (8), pp. 877883. Katzenmeier, G., Kussmaul, K., Roos, E., and Diem, H., 1990, Component Testing at the HDR-Facility for Validating the Calculation Procedures and the Transferability of the Test Results from Specimen to Component, Nucl. Eng. Des. 119, pp. 317-327. Keisler, J., Chopra, O. K., and Shack, W. J., 1995, Fatigue StrainLife Behavior of Carbon and LowAlloy Steels, Austenitic Stainless Steels, and Alloy 600 in LWR Environments, NUREG/CR6335, ANL95/15. Keisler, J., Chopra, O. K., and Shack, W. J., 1996, Fatigue StrainLife Behavior of Carbon and LowAlloy Steels, Austenitic Stainless Steels, and Alloy 600 in LWR Environments, Nucl. Eng. Des. 167, pp. 129154. Kussmaul, K., Blind, D., and Jansky, J., 1984, Formation and Growth of Cracking in Feed Water Pipes and RPV Nozzles, Nucl. Eng. Des. 81, pp. 105119. Kussmaul, K., Rintamaa, R., Jansky, J., Kemppainen, M., and Trrnen, K., 1983, On the Mechanism of Environmental Cracking Introduced by Cyclic Thermal Loading, IAEA Specialists Meeting Corrosion and Stress Corrosion of Steel Pressure Boundary Components and Steam Turbines, VTT Symp. 43, Espoo, Finland, pp. 195 243. Lenz, E., Liebert, A., and Wieling, N., 1990, Thermal Stratification Tests to Confirm the Applicability of Laboratory Data on Strain Induced Corrosion Cracking to
24-14

Component Behavior, 3rd IAEA Specialists Meeting on SubCritical Crack Growth, Moscow, pp. 6791. Lenz, E., Stellwag, B., and Wieling, N., 1983, The Influence of StrainInduced Corrosion Cracking on the Crack Initiation in LowAlloy Steels in HTWater A Relation Between Monotonic and Cyclic Crack Initiation Behavior, IAEA Specialists Meeting Corrosion and Stress Corrosion of Steel Pressure Boundary Components and Steam Turbines, VTT Symp. 43, Espoo, Finland, pp. 243267. Majumdar, S., Chopra, O. K., and Shack, W. J., 1993, Interim Fatigue Design Curves for Carbon, LowAlloy, and Austenitic Stainless Steels in LWR Environments, NUREG/CR5999, ANL93/3. Mehta, H. S., 1999, An Update on the EPRI/GE Environmental Fatigue Evaluation Methodology and its Applications, Probabilistic and Environmental Aspects of Fracture and Fatigues, PVP Vol. 386, S. Rahman, ed., American Society of Mechanical Engineers, New York, pp. 183193. Mehta, H. S., and Gosselin, S. R., 1996, An Environmental Factor Approach to Account for Reactor Water Effects in Light Water Reactor Pressure Vessel and Piping Fatigue Evaluations, Fatigue and Fracture Volume 1, PVP Vol. 323, H. S. Mehta, ed., American Society of Mechanical Engineers, New York, pp. 171185. Mehta, H. S., and Gosselin, S. R., 1998, Environmental Factor Approach to Account for Water Effects in Pressure Vessel and Piping Fatigue Evaluations, Nucl. Eng. Des. 181, pp. 175197. Miller, K. J., 1985, Initiation and Growth Rates of Short Fatigue Cracks, Fundamentals of Deformation and Fracture, Eshelby Memorial Symposium, Cambridge University Press, Cambridge, U.K., pp. 477500. Miller, K. J., 1995, Damage in Fatigue: A New Outlook, International Pressure Vessels and Piping Codes and Standards: Volume 1 Current Applications, PVP Vol. 3131, K. R. Rao and Y. Asada, eds., American Society of Mechanical Engineers, New York, pp. 191192. Mimaki, H., Kanasaki, H., Suzuki, I., Koyama, M., Akiyama, M., Okubo, T., and Mishima, Y., 1996, Material Aging Research Program for PWR Plants, Aging Management Through Maintenance Management, PVP Vol. 332, I. T. Kisisel, ed., American Society of Mechanical Engineers, New York, pp. 97105. NRC, 1979, Cracking in Feedwater System Piping, IE Bulletin No. 7913, U.S. Nuclear Regulatory Commission, Washington, DC. NRC, 1988a, Safety Injection Pipe Failure, NRC Information Notice 8801, U.S. Nuclear Regulatory Commission, Washington, DC. NRC, 1988b, Thermal Stresses in Piping Connected to Reactor Coolant Systems, NRC Bulletin No. 8808, U.S. Nuclear Regulatory Commission, Washington, DC.

NRC, 1988c, Pressurizer Surge Line Thermal Stratification, NRC Bulletin No. 8811, U.S. Nuclear Regulatory Commission, Washington, DC. NRC, 1993, Thermal Fatigue Cracking of Feedwater Piping to Steam Generators, NRC Information Notice 9320, U.S. Nuclear Regulatory Commission, Washington, DC. Nagata, N., Sato, S., and Katada, Y., 1991, LowCycle Fatigue Behavior of Pressure Vessel Steels in High Temperature Pressurized Water, ISIJ Intl. 31 (1), pp. 106114. Nakao, G., Kanasaki, H., Higuchi, M., Iida, K., and Asada, Y., 1995, Effects of Temperature and Dissolved Oxygen Content on Fatigue Life of Carbon and LowAlloy Steels in LWR Water Environment, Fatigue and Crack Growth: Environmental Effects, Modeling Studies, and Design Considerations, PVP Vol. 306, S. Yukawa, ed., American Society of Mechanical Engineers, New York, pp. 123128. Ranganath, S., Kass, J. N., and Heald, J. D., 1982, Fatigue Behavior of Carbon Steel Components in High Temperature Water Environments, BWR Environmental Cracking Margins for Carbon Steel Piping, EPRI NP 2406, Electric Power Research Institute, Palo Alto, CA, Appendix 3. Van Der Sluys, W. A., 1993, Evaluation of the Available Data on the Effect of the Environment on the Low Cycle Fatigue Properties in Light Water Reactor Environments, Proc. 6th Intl. Symp. on Environmental Degradation of Materials in Nuclear Power Systems Water Reactors, R. E. Gold and E. P. Simonen, eds., The Metallurgical Society, Warrendale, PA, pp. 14. Ware, A. G., Morton, D. K., and Nitzel, M. E., 1995, Application of NUREG/CR5999 Interim Design Curves to Selected Nuclear Power Plant Components, NUREG/CR6260, INEL95/0045. Watanabe, H., 1980, Boiling Water Reactor Feedwater Nozzle/Sparger, Final Report, NEDO21821A, General Electric Co., San Jose, CA. Yukawa, S., 1999, Meeting of the Steering Committee for Cyclic Life and Environmental Effects (CLEE), the Pressure Vessel Research Council, June, Columbus, OH.

24-15

25
MECHANISM OF FATIGUE CRACK INITIATION IN LIGHT WATER REACTOR COOLANT ENVIRONMENTS
Omesh K. Chopra Argonne National Laboratory Argonne, IL 60439 USA HeungBae Park Korea Power Engineering Company, Inc., 3609 Mabukri, Kusongmyon, Yonginshi, Kyunggido, Korea

25-1

International Conference on Fatigue of Reactor Components July 31 - August 2, 2000 Napa California, USA

MECHANISM OF FATIGUE CRACK INITIATION IN LIGHT WATER REACTOR COOLANT ENVIRONMENTS


Omesh K. Chopra1 and HeungBae Park2 National Laboratory, Argonne, IL 60439 USA 2Korea Power Engineering Company, Inc., 3609 Mabukri, Kusongmyon, Yonginshi, Kyunggido, Korea
1Argonne

KEYWORDS Fatigue Crack Initiation, Strainvs.Life (SN) Curve, LWR Environment, Crack Growth Rate, Slip Oxidation/Dissolution
ABSTRACT

The effects of a light water reactor (LWR) coolant environment on the fatigue resistance of materials is not explicitly addressed in the ASME Code design fatigue curves. Existing fatigue strainvs.life (SN) data illustrate potentially significant effects of LWR coolant environments on the fatigue life of piping and pressure vessel steels. In this paper, we discuss the influence of reactor environments on the mechanism of fatigue crack initiation. Decreased fatigue lives of carbon and lowalloy steels in water with high dissolved oxygen are caused primarily by the effects of environment on the growth of short cracks that are <100 2m deep. In LWR environments, the growth of these small cracks occurs by a slip oxidation/dissolution process. A fracture mechanics approach has been used to evaluate the effects of environment on fatigue crack initiation. The fatigue life, defined as the number of cycles required to form an engineeringsize crack, i.e., a 3 mmdeep crack, is considered to be composed of the growth of microstructurally and mechanically small cracks. The growth of the latter is characterized in terms of J and crackgrowth rate (da/dN) data in air and LWR environments. The growth of microstructurally small cracks is expressed by a modified Hobson relationship in air and by the slip oxidation/dissolution model in water. The estimated fatigue S-N curves agree well with the experimental data for carbon and lowalloy steels in air and water environments.
INTRODUCTION

The formation of surface cracks and their growth as shear and tensile cracks to an engineering size (i.e., 3 mm deep) constitute the fatigue life of a material, which is represented by stress or strain amplitudevs.fatigue life (SN) curves. These curves define, for a given stress or strain amplitude, the number of cycles needed to form an engineeringsize crack.

Cyclic loadings on a structural component occur because of changes in mechanical and thermal loadings as the system passes from one load set (e.g., pressure, temperature, moment, and force loading) to any other load set. For each load set, an individual fatigue usage factor is determined by the ratio of the number of cycles anticipated during the lifetime of the component to the allowable cycles. Figures I9.1 through I 9.6 of Appendix I to Section III of the ASME Boiler and Pressure Vessel Code specify fatigue design curves that define the allowable number of cycles as a function of appliedstress amplitude. The cumulative usage factor (CUF) is the sum of the individual usage factors, and the ASME Code Section III requires that the CUF at each location must not exceed 1. The current ASME Code design fatigue curves are based on straincontrolled fatigue tests of small polished specimens in air at room temperature. The design fatigue curves have been obtained by first adjusting the best-fit curves to the experimental data for mean stress effects and then decreasing the adjusted curves by a factor of 2 on stress or 20 on cycles, whichever was more conservative, at each point on the curve. These factors were intended to account for the differences and uncertainties in relating fatigue lives of laboratory test specimens to those of actual reactor components. The factors of 2 and 20 are not safety margins but rather conversion factors that must be applied to the experimental data to obtain reasonable estimates of the lives of actual reactor components. The effects of light water reactor (LWR) coolant environments on fatigue resistance of a material are not explicitly addressed in the Code design fatigue curves. Existing fatigue SN data illustrate potentially significant effects of LWR coolant environments on the fatigue resistance of carbon steels (CSs), lowalloy steels (LASs) (Chopra and Shack, 1997, 1998a, 1998b, 1999; Chopra and Muscara, 2000), and austenitic stainless steels (Chopra and Gavenda, 1998; Chopra and Smith, 1998; Chopra, 1999, Chopra and

25-3

Muscara, 2000). The key parameters that influence fatigue life in LWR environments are temperature, dissolved oxygen (DO) level in the water, loading or strain rate, and strain (or stress) amplitude; for carbon and lowalloy steels, the sulfur content of the steel is also important. Under certain environmental and loading conditions, the environmental effects alone substantially exceed the factor of 20 on life that is used to account for the differences between specimen tests and component behavior. The objective of this paper is to use crack growth data and fracture mechanics analysis to examine the fatigue S-N behavior of carbon and low-alloy steels in air and LWR environments. The influence of reactor environments on the mechanism of fatigue crack initiation is discussed. Fatigue life is considered to be composed of the growth of microstructurally small cracks (MSCs) and mechanically small cracks. The growth of the latter has been characterized in terms of the Jintegral range J and crackgrowthrate (CGR) data in air and LWR environments. The growth of microstructurally small cracks in air is expressed by a modified version of the relationship presented by Hobson (1982) and by the slip dissolution/oxidation process (Ford et al., 1993) in water. MECHANISM OF FATIGUE CRACK INITIATION The formation of surface cracks and their growth as shear (Stage I) and tensile (Stage II) cracks to an engineering size (3 mm deep) constitute the fatigue life of a material, which is represented by the fatigue SN curves. The curves specify, for a given stress or strain amplitude, the number of cycles needed to form an engineering crack. During fatigue loading of smooth test specimens, surface cracks 10 2m or longer form quite early in life (i.e., <10% of life) at surface irregularities or discontinuities either already in existence or produced by slip bands, grain boundaries, secondphase particles, etc. (Miller, 1985; Tokaji et al., 1988; Gavenda et al., 1997; Obrtlik et al., 1997; Chopra and Shack, 1998a). Consequently, fatigue life may be considered to be composed entirely of crack propagation (Miller, 1995). Growth of these surface cracks may be divided into two regimes; an initial period, which involves growth of MSCs, that is very sensitive to microstructure and is characterized by decelerating crack growth (Region AB in Fig. 1), and a propagation period, that involves growth of mechanically small cracks that can be predicted by fracture mechanics methodology and is characterized by accelerating crack growth (Region BC in Fig. 1). Mechanically small cracks, which correspond to Stage II, or tensile, cracks are characterized by striated crack growth and a fracture surface normal to the maximum principal stress. Conventionally, the former has been defined as the initiation stage and is considered sensitive to stress or strain amplitude, and the latter has been defined as the propagation stage and is less sensitive to strain amplitude.

The characterization and understanding of both the crack initiation and crack propagation stage are important for accurate estimates of the fatigue lives of structural materials.
C

Crack Length

Mechanically Small Crack (Stage II Tensile Crack) B

8 2

8 1
Microstructurally Small Crack (MSC) (StageI Shear Crack)

8 2 > 8 1
0.6 0.8 1

0.2

0.4

Life Fraction
Mechanically Small Crack

Microstructurally Small Crack

8 3 8 2 8 1

8 1 Non Propagating Cracks

LEFM or EPFM

8 3 > 8 2 > 8 1
Crack Length

Fig. 1. Schematic illustrations of (a) growth of short cracks in smooth specimens as a function of fatigue life fraction and (b) crack velocity as a function of crack length. LEFM = linear elastic fracture mechanics; EPFM = elastic plastic fracture mechanics. Reduction of fatigue life in hightemperature water has often been attributed to easier crack initiation, because surface micropits that are present in hightemperature water act as stress raisers and provide preferred sites for the formation of fatigue cracks (Nagata et al., 1991). However, experimental data do not support this argument; the fatigue lives of carbon and lowalloy steel specimens that have been preoxidized at 288C in highDO water and then tested in air are identical to those of unoxidized specimens (Fig. 2) (Chopra and Shack, 1998a). If the presence of micropits was responsible for the reduction in life, specimens preexposed to highDO water and tested in air should show a decrease in life. Also, the fatigue limit of these steels should be lower in water than in air. Data obtained from specimens in highDO water indicate that the fatigue limit is either the same as, or than in air (Chopra and Shack, 1998a, b).

 KLJKHU LQ ZDWHU

25-4

0 t (%)

Air

Total Strain Range,

1.0

Total Strain Range,

F/F S/F Preoxidized F/F in air F/F in <10 ppb DO S/F in <10 ppb DO

0 t (%)

A106Gr B Steel 288C Water 0.50.8 ppm DO

Strain Rates (%/s) S: 0.004 & F: 0.4

A533Gr B Steel 288C Water 0.50.8 ppm DO

Air

1.0

Strain Rates (%/s) S: 0.004 & F: 0.4 F/F S/F Preoxidized F/F in air F/F in <10 ppb DO

0.1 10 1

10 2

10 3

10 4

10 5

10 6

10 7

0.1 10 1

10 2

10 3

10 4

10 5

10 6

10 7

Fatigue Life, N 25

Fatigue Life, N 25

Fig. 2. Effects of environment on formation of fatigue cracks in carbon and lowalloy steels. Preoxidized specimens were exposed at 288C for 30100 h in water with 060.8 ppm dissolved oxygen.
120 A106 GrB Carbon Steel 100 Number of Cracks 80 60 40 20 0 -20 0.2 0.4 0.6 0.8 1.0 Strain Range (%) 2.0
Air PWR >0.6 ppm DO Open Symbols: fast/fast test Closed Symbols: slow/fast or fast/slow test

120 A533 GrB LowAlloy Steel 100 Number of Cracks 80 60 40 20 0 -20 0.2 0.4 0.6 0.8 1.0 Strain Range (%) 2.0
Air PWR >0.6 ppm DO Open Symbols: fast/fast test Closed Symbols: slow/fast or fast/slow test

Fig. 3. Number of cracks >10 2m long along longitudinal section of fatigue specimens of (a) A106 Gr B carbon steel and (b) A533 Gr B lowalloy steel tested in various environments. Number of cracks represents the average value along a 7 mm gauge length. Furthermore, if reduction in life is caused by easier formation of cracks, the specimens tested in highDO water should show more cracks. Figure 3 shows plots of the number of cracks >10 2m long, along longitudinal sections of the gauge length of A106Gr B and A533Gr B specimens as a function of strain range in air, simulated PWR environment, and highDO water at two strain rates. The results show that with the exception of the LAS tested in simulated PWR water, environment has no effect on the frequency (number per unit gauge length) of cracks. For similar loading conditions, the number of cracks in the specimens tested in air and highDO water is identical, although fatigue life is lower by a factor of  LQ ZDWHU 'HWDLOHG metallographic evaluation of the fatigue test specimens indicates that the water environment has little or no effect on the formation of surface microcracks. Irrespective of environment, cracks in carbon and lowalloy steels initiate along slip bands, carbide particles, or at the ferrite/pearlite phase boundaries. The enhanced growth rates of long cracks in pressure vessel and piping steels in LWR environments have been attributed to either slip oxidation/dissolution (Ford, 1986) or hydrogeninduced cracking (Hnninen et al., 1986). Both mechanisms depend on the rates of oxide rupture, passivation, and liquid diffusion. Therefore, it is often difficult to differentiate between the two processes or to establish their relative contributions to crack growth in LWR environments. Studies on crack initiation in smooth fatigue specimens (Gavenda et al., 1997) indicate that the decrease in fatigue life of CSs and LASs in LWR environments is caused primarily by the effects of environment on the growth of cracks <100 2m deep. When compared with CGRs in air, growth rates in highDO water are nearly two orders of magnitude higher for cracks that are <100 2m deep and one order of magnitude higher for cracks that are >100 2m deep. Metallographic examination of test specimens indicates that in highDO water, surface cracks <100 2mdeep grow entirely as tensile cracks normal to the stress, whereas in air or simulated PWR environments, they are at an angle of 45 to the stress axis (Fig. 4) (Gavenda et al., 1997). Also, for CSs, cracks <100 2m deep propagate across both the soft ferrite and hard pearlite

25-5

(a) (b) Fig. 4. Photomicrographs of fatigue cracks along gauge sections of A106Gr. B carbon steel in (a) air and (b) highDO water at 288C

(a) (b) Fig. 5. Photomicrographs of fatigue cracks on gauge surfaces of A106Gr. B lowalloy steel in (a) air and (b) highDO water at 288C regions, whereas in air, they propagate only along soft ferrite regions. The crack morphology on the specimen surface also differs in air and water environments (Fig. 5); surface cracks in highDO water are always straight and normal to the stress axis, whereas in air or simulated PWR environments, they are mostly at 45 to the stress axis. The different crack morphology, absence of Stage I crack growth, and propagation of nearsurface cracks across pearlite regions indicate that in high-DO water, growth of MSCs occurs predominantly by the slip oxidation/dissolution process. In highDO water, crack initiation in CSs and LASs may be explained as follows: surface microcracks form quite early in fatigue life. During cyclic loading, the protective oxide film is ruptured at strains greater than the fracture strain of surface oxides, and the microcracks grow by anodic dissolution of the freshly exposed surface to crack lengths greater than the critical length of MSCs. These mechanically small cracks grow to engineering size, and their growth, which is characterized by accelerating rates, can be predicted by fracture mechanics methodology. FATIGUE SN DATA IN LWR ENVIRONMENTS The fatigue lives of both carbon and lowalloy steels are decreased in LWR environments; the reduction in life depends on temperature, strain rate, DO level in the water, and sulfur content of the steel (Chopra and Shack, 1998a, 1998b, 1999; Chopra and Muscara, 2000; Higuchi and Iida, 1991; Higuchi et al., 1997; Higuchi, 1999). Statistical models based on existing fatigue SN data have been developed for estimating fatigue lives of these steels in air and LWR environments (Chopra and Shack, 1998a, 1999; Chopra and Muscara, 2000). The effects of the critical parameters on fatigue life and their threshold values, as well as the statistical models for estimating fatigue life in air and LWR environments, have been presented in a companion paper at this conference.

INITIAL CRACK SIZE Studies on crack initiation in smooth fatigue specimens indicate that surface cracks form quite early in life. Smith et al. (1996) detected 102m deep surface cracks at temperatures up to 700oC in Waspalloy. Hussain et al. (1994) examined the

25-6

growth of 2mdeep surface cracks through four or more grains. Tokaji et al. (1986a, b, 1988, 1992) defined crack initiation as the formation of a 102mdeep crack. Gavenda et al. (1997) reported that in roomtemperature air, 102m deep cracks form early during fatigue life, i.e., <10% of fatigue life. Suh et al. (1985, 1990) reported that a crack is said to have initiated when any crack-like mark grows across a grain boundary, or when the separation of grain boundaries becomes clear. Based on these results, it is reasonable to assume the initial depth of MSCs to be  2m. TRANSITION FROM MICROSTRUCTURALLY SMALL TO MECHANICALLY SMALL CRACK Various criteria may be used to define the crack length for transition from microstructurally small to mechanically small crack. They may be related to the plastic zone size, crack lengthvs.fatigue life (aN) curve, Weibull distribution of the cumulative probability of fracture, stress rangevs.crack length curve, or grain size. The results indicate that the crack length for transition from MSC to mechanically small crack depends on applied stress and microstructure of the material. Plastic Zone de los Rios et al., (1992, 1996) and Lankford (1977, 1982, 1985) defined the transition from small to large cracks as the crack length at which the size of the linear elastic fracture mechanics (LEFM) plastic zone exceeds a grain diameter. Crack Lengthvs.Fatigue Life Curve Obrtlik et al. (1997) divided the fatigue crack length a vs.fatigue life N curves into two regimes: MSCs, in which the dependence of crack length on fatigue life can be represented by a straight line; and mechanically small cracks, in which fatigue crack growth is represented by an exponential function fit of the experimental data. Weibull Distribution of the Cumulative Probability of Fracture Suh et al. (1985, 1990) used the knee in the Weibull distribution of cumulative probability of fracture to define the transition from shear crack growth to tensile crack growth. The knee occurred in the range of 35 grain diameters. Stress Rangevs.Crack Length Curve Kitagawa and Takahashi (1976) and Taylor and Knott (1981) used the stress rangevs.crack length curve to discriminate a MSC from a mechanically small crack. For crack lengths >500 2m, plots of the threshold stress range for fatigue crack growth (8th) vs. crack length yield a straight line, i.e., the threshold stress intensity factor (Kth) is constant. For crack lengths < 500 2m, 8th deviates from the linear relationship and approaches a constant value as the crack length becomes smaller. The constant value of 8th is approximately equal to the fatigue limit of a smooth specimen of the material. The crack length at which the 8th vs. crack length curve changes from a linear relationship to a constant

value is used to define the transition from microstructurally small to mechanically small cracks. Grain Size Tokaji et al. (1986a, 1988, 1992) estimated the transition crack length to be about 8 times the microstructural unit size. Ravichandran (1997) reported that large fluctuations in crack shape or aspect ratio occur at crack lengths of approximately a few grain diameters (typically five or fewer grain diameters). Hussain et al. (1994) observed that fatigue CGRs decreased systematically at microstructural heterogeneities up to a length of three or four grain diameters. Dowling (1977) reported that the J-integral correlation is not valid for surface crack lengths <10 crystallographic grain diameters. The above studies indicate that the crack length for transition from MSC to mechanically small crack is a function of applied stress and microstucture of the material; actual value may range from 150 to 250 2m. A constant value of  2m was assumed for convenience, for both carbon and lowalloy steels; it is the initial size for mechanically small cracks.

FATIGUE CRACK GROWTH RATES Air Environment The growth rates da/dN (mm/cycle) of MSCs, i.e., from 10 to 200 2m, in air can be represented by the Hobson relationship (Hobson, 1982; Miller, 1985; Brown, 1986; Carbonell and Brown, 1986) da/dN = A1 8 n1 (d a), (1) where a is the length (mm) of the predominant crack, 8 is the stress range (MPa), constant A1 and exponent n1 are determined from existing fatigue SN data, and d is the material constant related to grain size. The values of A1 and n1 for carbon and lowalloy steels at room temperature and reactor operating temperatures are given in Table 1. A value of 0.3 mm was used for the material constant d. Also, because growth rates increase significantly with decreasing crack length, a constant growth rate was assumed for crack lengths smaller than 0.075 mm. The applied stress range 8 is determined from RambergOsgood relations given by Eqs. A1A5 of the Appendix, it represents the value at fatigue halflife. Table 1. Values of the constants A1 and n1 in Equation 1
Steel Type Carbon LowAlloy Temperature Room Operating Room Operating A1 3.33 x 1041 9.54 x 1034 1.45 x 1036 1.07 x 1043 n1 13.13 10.03 11.10 13.43

25-7

The growth rates of mechanically small cracks in air are estimated from Eq. A8 of the Appendix. A factor of 1.22 enhancement in growth rates was used at reactor operating temperatures.

da/dN = 7.03 x 105(0 0f)( 0 app )0.65,

(4a)

and from Eq. 3b for moderate environmental effects is given by da/dN = 4.00 x 103(0 0f),
0

(4b) (s1)

LWR Environment A model based on oxide film rupture and anodic dissolution (or slip dissolution/oxidation model) was proposed by Ford et al., (1993) to incorporate the effects of LWR environments on fatigue lives of CSs and LASs. The model considers that a thermodynamically stable protective oxide film forms on the surface to ensure that the crack will propagate with a high aspect ratio without degrading into a blunt pit, and that a strain increment is required to rupture the oxide film, thereby exposing the underlying matrix to the environment. Once the passive oxide film is ruptured, crack extension is controlled by dissolution of freshly exposed surfaces and by oxidation characteristics. Ford and Andresen (1995) proposed that the average crack growth rate da/dt (cm/s) is related to the crack tip strain rate 0 ct (s1) by the relationship da/dt = A2 2 , 0 ct
n

(2)

where the constant A2 and exponent n2 depend on the material and environmental conditions at the crack tip. A lower limit of crack propagation rate is associated with blunting when the crack tip cannot keep up with the general corrosion rate of the crack sides or when a critical level of sulfide ions cannot be maintained at the crack tip. The crack propagation rate at which this transition occurs may depend on the DO level, flow rate, etc. Based on these factors, the maximum and minimum environmentally assisted crack propagation rates have been defined by Ford et al. (1993), Ford and Andresen (1995), and Ford (1996). For cracktip sulfide ion concentrations above the critical level, CGR is expressed as da/dt = 2.25 x
104( 0
ct

where app is the applied strain rate and 0f is the threshold strain range needed to rupture the oxide film; 0f was assumed to be 0.0023 and 0.0029, respectively, for CSs and LASs. For strain rates > da/dN is lower from Eq. 4a than from Eq. 4b. Also, existing fatigue SN data indicate that strain rate effects on life saturate at 1998a). Therefore, Eq. 4a can be applied at rates between 0.003 and 0.00001 s1, 0 app is assumed to be 0.003 s1 for higher values, and 0.00001 s1 for lower values. Equations 4a and 4b assume that the stressfree state for the surface oxide film is at peak compressive stress. Studies on crack initiation and crack growth in smooth fatigue specimens indicate that the reference fatigue CGR curves (Fig. A1 in the Appendix) for carbon and lowalloy steels in LWR environments are somewhat higher than those determined experimentally from the growth of mechanically small cracks in LWR environments (Gavenda et al., 1997). Furthermore, using the reference CGR curves and fracture mechanics analyses to examine the fatigue S-N behavior of these steels in LWR environments yields conservative results. Therefore, the reference fatigue CGR curves were modified to estimate the growth rates of mechanically small cracks; the modified curves are shown in Fig. 6. The threshold values of K (MPam1/2) are given by

V

V &KRSUD DQG 6KDFN

Kb = 10.11 6
and



(5a) (5b)

Kc = 32.03 60.326, where rise time 6 is in seconds.


10 0 Carbon & LowAlloy Steels 10 -1 Crack Growth Rates (mm/cycle)
R=0 6 = 100 s da/dN = 7.67 x 106 Susceptible to EAC da/dN = 5.67 x 10 8

)0.35;

(3a)

for cracktip sulfide ion concentrations below the critical level, it is expressed as da/dt =
102( 0
ct

6 0.691  K0.949

10 -2

 K 3.07

)1.0.

(3b)

10 -3

However, the growth rates predicted by Eqs. 3a and 3b are somewhat higher than those observed experimentally (Gavenda et al., 1997). To be consistent with the experimental data, the constants in Eqs. 3a and 3b were decreased by a factor of 3.2 and 2.5, respectively. Assuming that 0 ct is approximately the same as the applied strain rate, 0 app , and crack advance due to mechanical fatigue is insignificant during the initial stages of fatigue damage, crack advance per cycle from Eq. 3a for significant environmental effects is given by

10 -4

Air Environment da/dN = 3.78 x 10 9

 K3.07

10 -5

Not susceptible to EAC da/dN = 4.91 x 109

 K3.07  Kc

10 -6 10 1

 Kb
10 2

 K (MPam

1/2

Fig. 6. Modified reference fatigue crack growth rate curves for carbon and lowalloy steels for LWR applications

25-8

Crack Growth Rate da/dN (2m/cycle)

10 1

LowAlloy Steel 288C 0.70% Strain Range 0.01%/s Strain Rate

10 0

Simulated BWR

Simulated PWR

10 -1

Air

10 -2 10

100 1000 Crack Length (2m)

Fig. 7. Crack growth rates during fatigue crack initiation in lowalloy steels in air and simulated PWR and BWR environments

mechanically small cracks are represented by the curves for materials not susceptible to environmentally assisted cracking (EAC), and those of MSCs, by either Eq. 4b or Eq. 1, whichever yields the higher value. For example, at high strain ranges, growth rates determined from Eq. 1 can be higher than those determined from Eq. 4b, i.e., mechanical factors control crack growth and environmental effects are insignificant. Environmental effects on fatigue life are significant when all of the threshold conditions are satisfied, e.g., temperature J 150C, DO J 0.05 ppm, strain rate <1%/s, and strain range is above the critical value. A minimum threshold sulfur content of 0.005 wt.% was also considered, i.e., sulfur content must also be >0.005 wt.% for significant environmental effects on fatigue life. When all five threshold conditions are satisfied, the growth rates of mechanically small cracks are represented by the curve for materials susceptible to EAC for K values below Kb, by the curve for materials not susceptible to EAC at K values above Kc, and by the transition curve for inbetween values of K. The growth rates of MSCs are represented by either Eq. 4a or Eq. 1, whichever yields the higher value.

103

102
Reactor Operating Temps. Strain Range 0.75% Air Strain Rate 0.1%/s BWR Water Open Symbols: LowAlloy Steel PWR Water Closed Symbols: Carbon Steel

101

1000

2000

3000 4000 5000 Number of Cycles, N

6000

7000

103

102
Reactor Operating Temps. Strain Range 0.75% Strain Rate 0.01%/s Open Symbols: LowAlloy Steel Closed Symbols: Carbon Steel

Air BWR Water PWR Water

101

1000

2000

3000 4000 5000 Number of Cycles, N

6000

7000

Fig. 8. Crack growth in carbon and lowalloy steels as a function of fatigue cycles at two strain rates Environmental effects on fatigue life are moderate when any one of the threshold environmental conditions is not satisfied, e.g., temperature <150C, DO <0.05 ppm, strain rate >1%/s, or strain range is below the critical value. For moderate environmental effects, the growth rates of

ESTIMATES OF FATIGUE LIFE The existing fatigue SN data for carbon and lowalloy steels in air and LWR environments were examined with the present model, in which fatigue life consists of the growth of MSCs and mechanically small cracks. The former may be defined as the initiation stage and represents the growth of MSCs from 10 to 200 2m. The growth of mechanically small cracks may be defined as the propagation stage and represents the growth of fatigue cracks from 200 to 3000 2m. During the initiation stage, the growth of MSCs is expressed by a modified Hobson relationship in air (Eq. 1) and by the slip dissolution/oxidation process in water (Eqs. 4a or 4b). During the propagation stage, the growth of mechanically small cracks is characterized in terms of the Jintegral range J and CGR data in air and LWR environments (Fig. 3). The correlations for calculating the stress range, stress intensity range K, J integral range J, and the CGRs for long cracks in air are given in the appendix. For the cylindrical fatigue specimens, the stress intensity ranges K were determined from the values of the Jintegral range J. Although J is often computed only for that portion of the loading cycle during which the crack is open, in the present study, the entire hysteresis loop was used when we estimated J (Dowling, 1977). The stress intensities associated with conventional cylindrical fatigue specimens were modified according to the correlations developed by O'Donnell and O'Donnell (1995). Typical CGRs and crack growth behavior during fatigue crack initiation in air and water environments are shown in Figs. 7 and 8, respectively.

2 Crack Length ( m)

2 Crack Length ( m)

25-9

Experimental values of fatigue life and those predicted from the present model in air and low and highDO water are plotted in Fig. 9. The predicted fatigue lives in air show excellent agreement with the experimental data; the predicted values in LWR environments, particularly in highDO water, are slightly lower than the experimental values. The differences in predicted and experimental fatigue lives in LWR environments are most likely due to crack closure effects that are expected to be significant at low strain amplitudes. The fatigue SN curves developed from the present model and those obtained from the statistical models in air and in PWR and BWR environments are shown in Figs. 10 and 11, respectively.

SUMMARY The influence of reactor environments on the mechanism of fatigue crack initiation is discussed. Decreased fatigue lives of carbon and lowalloy steels in highDO water are caused
10 7 Air Room Temperature 10 6

primarily by the effects of environment on the growth of small cracks <100 2m deep. In LWR environments, the growth of these small fatigue cracks in carbon and lowalloy steels occurs by a slip oxidation/dissolution process. To predict the fatigue lives of carbon and low-alloy steels in air and LWR environments, we used fracture mechanics approach in which we consider fatigue life to be composed of the growth of microstructurally and mechanically small cracks. The growth of the former cracks is very sensitive to microstructure and is characterized by decelerating crack growth, that of the latter, which can be predicted by fracture mechanics methodology, is characterized by accelerating crack growth, and has been characterized in terms of the Jintegral range J and CGR data in air and LWR environments. The growth of MSCs is expressed by a modified Hobson relationship in air and by the slip dissolution/oxidation process in water. The crack length for transition from microstructurally small crack to mechanically small crack was
10 7 Air Operating Temperature 10 6

Predicted Life (Cycles)

Predicted Life (Cycles)

10 5

10 5

10 4

10 4

10 3

10 3

10 2 Carbon Steel Low-Alloy Steel 10 1 10 1 10 2 10 3 10 4 10 5 10 6 10 7

10 2 Carbon Steel Low-Alloy Steel 10 1 10 1 10 2 10 3 10 4 10 5 10 6 10 7

Observed Life (Cycles) 10 7 Carbon Steel Water Operating Temperature 10 7

Observed Life (Cycles)

10 6

10 6

LowAlloy Steel Water Operating Temperature

Predicted Life (Cycles)

Predicted Life (Cycles)

10 5

10 5

10 4

10 4

10 3

10 3

10 2 <0.05 ppm DO 0.05 ppm DO 10 1 10 1 10 2 10 3 10 4 10 5 10 6 10 7

10 2 <0.05 ppm DO 0.05 ppm DO 10 1 10 1 10 2 10 3 10 4 10 5 10 6 10 7 Observed Life (Cycles)

Observed Life (Cycles)

Fig. 9. Experimentally observed values of fatigue life of carbon and lowalloy steels vs. those predicted by the present model in air and water environments

25-10

101 Air Carbon Steel Operating Temperture

101 Air LowAlloy Steel Operating Temperture

Strain Amplitude, 0a (%)

Strain Amplitude, 0a (%)

100

100

10-1 101 102

Statistical Model Present Model 103 104 Number of Cycles, N 105 106

10-1 101 102

Statistical Model Present Model 103 104 Number of Cycles, N 105 106

Fig. 10.

Fatigue strainvs.life curves developed from the present model for carbon and lowalloy steels in air
101 PWR Environment Carbon Steel Operating Temperture 101 PWR Environment LowAlloy Steel Operating Temperture

Strain Amplitude, 0a (%)

Strain Amplitude,0a (%)

100

100

10-1 101
101 Strain Rate 0.1%/s

Statistical Model Present Model 102 103 104 Number of Cycles, N 105 106

10-1 101 101


BWR Environment Carbon Steel Operating Temperture Statistical Model Present Model

Statistical Model Present Model 102 103 104 Number of Cycles, N 105 106

Strain Amplitude, 0a (%)

Strain Amplitude,0a (%)

Strain Rate 0.1%/s

BWR Environment LowAlloy Steel Operating Temperture Statistical Model Present Model

100

100

Strain Rate 0.001%/s

Strain Rate 0.001%/s

10-1 101 102 103 104 Number of Cycles, N 105 106

10-1 101 102 103 104 Number of Cycles, N 105 106

Fig. 11. Fatigue strainvs.life curves developed from the present model for carbon and lowalloy steels in LWR environments based on studies of small crack growth.. Fatigue lives estimated from the present model show good agreement with the experimental data for carbon and lowalloy steels in air and LWR environments. At low strain amplitudes, i.e., fatigue lives of >104 cycles, the predicted lives in water are slightly lower than those observed experimentally, most likely because of the effects of crack closure. APPENDIX Cyclic Stress Range The cyclic stressstrain response of carbon and lowalloy steels varies with steel type, temperature, and strain rate. In general, these steels exhibit initial cyclic hardening, followed by cyclic softening or a saturation stage. The CSs, with a pearlite and ferrite structure and low yield stress, show significant initial hardening. The LASs, which consist of tempered ferrite and a bainite structure, exhibit a relatively high yield stress, and show little or no initial hardening, may exhibit cyclic softening at high strain ranges. At 200370C, these steels exhibit dynamic strain aging, which leads to enhanced cyclic hardening, a secondary hardening stage, and

ACKNOWLEDGMENTS This work was sponsored by the Office of Nuclear Regulatory Research, U.S. Nuclear Regulatory Commission, Job Code W6610.

25-11

negative strain rate sensitivity. Under the conditions of dynamic strain aging, cyclic stress increases with decreases in strain rate. The cyclic stress range vs. strain range relationship is expressed by the modified RambergOsgood relationship given by

value, i.e., the crackdepthtospecimendiameter ratio can be as high as 0.4. Therefore, the geometrical correction factor Fs for a small semicircular surface crack was modified according to the correlation developed by O'Donnell and O'Donnell (1995): Fs = 0.6911 + 1.2685 (a/D) 5.6638 (a/D)2 + 21.511 (a/D)3, (A7) where D is specimen diameter. For conventional fatigue tests on cylindrical specimens, Fs may increase up to 1.7. The Jintegral range J is calculated from the ranges of cyclic stress and plastic strain, determined from stable hysteresis loops, i.e., at fatigue halflife. In general, J is computed only for that portion of the loading cycle during which the crack is open. For fully reversed cyclic loading, the crack opening point can be identified as the point where the curvature of the loadvs.displacement line changed before the peak compressive load. In the present study, evidence of a crack opening point was observed for cracks that had grown relatively large, i.e., near the end of fatigue life. Therefore, as recommended by Dowling (1977), the entire hysteresis loop was used in estimating J.

'H = 8 E + 8 A3

n3

(A1)

where E is Youngs modulus, constant A3 and exponent n3 are determined from the experimental data, and cyclic stress range corresponds to the value at halflife. At room temperature, the relationship of cyclic stress range (MPa) to strain range for CSs may be represented by

'H  'H = ('V/2010) + ('V/766.1)(1/0.207),


and for LASs, by

'V

'

'

(A2)

(A3) The effect of strain rate on the cyclic stressstrain curve is not considered at room temperature. At 288C, the cyclic stress strain curves may be represented by the correlations developed by Chopra and Shack (1998a). For CSs, the curve is given by the relationship
where Asig varies with the strain rate 0 (%/s) expressed as Asig = 1079.7 50.9 log( 0 ). For LASs, the curve is given by the relationship

'H = ('V/2010) + ('V/847.4)(1/0.173).

'

'H = ('V/1965) + ('V/Asig)(1/0.129),

(A4a) (A4b) (A5a) (A5b)

'H = ('V/1965) + ('V/Bsig)(1/0.110),


where Bsig is expressed as
Bsig = 961.8 30.3 log( 0 ).

Crack Growth Rate The fatigue CGRs da/dN of structural materials are characterized in terms of the range of applied stress intensity factor K and are given in Article A4300 of Section XI of the ASME Boiler and Pressure Vessel Code. For a stress ratio R in the range of 2 <R <0, the reference fatigue CGRs da/dN (mm/cycle) of carbon and lowalloys steels exposed to air environments are given by

'

Stress Intensity Factor Range For cylindrical fatigue specimens, the range of stress intensity factor K was determined from the value of the J integral range J, which, for a small semicircular surface crack, is given by Dowling (1977) as

' '

'J = 3.2 ('V2/2E) a + 5 ['V 'Hp/(S + 1)] a, where 'Hp is plastic strain range (%) (second ' 'K = (E 'J)1/2,

(A6a)

term in the Ramberg Osgood relationship) and S is the reciprocal of the strain hardening exponent n in Eq. A1. The stress intensity factor range K is obtained from (A6b) where E is the elastic modulus. Equation A6a incorporates a combined surface and flaw shape correction factor Fs of 0.714, which is derived from equivalent linear elastic solutions; Eq. 3a is valid as long as the crack size is very small when compared with the specimen diameter. For conventional fatigue tests, life is defined as the number of cycles for the tensile stress to decrease 25% from the peak or steadystate

(A8) da/dN = 3.78 x 109 ( K)3.07, where K = Kmax, the maximum stress intensity factor (MPam1/2). However, the effect of temperature is not considered in Eq. A8; CGRs are generally higher at 288C than at room temperature (Logsdon and Liaw, 1985). The results of Logsdon and Liaw indicate that for both CSs and LASs, CGRs are temperature. Section XI of the ASME Code also includes CGR curves for these steels exposed to LWR environments. The growth rates are represented by two curves for low and high values of K. However, the curves do not consider the effects of loading rate. Recent experimental results have shown the importance of key variables of material, environment, and loading rate on CGRs in LWR environments. Fatigue CGR correlations have been developed to explicitly consider the effects of loading rate, stress ratio R, K, and sulfur content in the steel (Eason et al., 1994). The new correlations, shown in Fig. A1, are divided into two categories: (a) for materials not susceptible to environmental effects, e.g., when sulfur content in the steel is low, CGRs are a factor of 2.8 higher than those in air; and

'

'

 KLJKHU DW & WKDQ DW URRP

'

'

25-12

(b) for materials susceptible to environmental effects, e.g., when sulfur content in the steel is high, CGRs are defined in terms of rise time 6, stress ratio R, and K.
10 0 Carbon & LowAlloy Steels 10 -1 Crack Growth Rates (mm/cycle)
R=0 6 = 100 s

da/dN = 1.18 x 10 5 6 0.691 DK0.949

10 -2

Susceptible to EAC da/dN = 1.56 x 10 7  K3.07

10 -3

10 -4

Air Environment da/dN = 3.78 x 10 9  K3.07 Not susceptible to EAC

10 -5

da/dN = 1.07 x 10 8  K3.07

 Ka  Kb
10 -6 10 1

 Kc  Kd
10 2

 K (MPam

1/2

Fig. A1. Proposed reference fatigue crack growth rate curves for carbon and lowalloy steels in LWR environments for a rise time of 100 s and R = 1 The correlations in Fig. A1 correspond to a rise time of 100 s and Kmin <0, e.g., fully reversed cyclic loading; R is set to zero. The various threshold values of K (MPam1/2) are given by

Ka = 14.156 60.125, Kb = 7.691 60.326, Kc = 27.186 60.326, Kd = 44.308 60.326, where rise time 6 is in seconds.

(A9a) (A9b) (A9c) (A9d)

REFERENCES Brown, M. W., 1986, Interface Between Short, Long and Non-Propagating Cracks, Mechanical Engineering Pub., pp. 423-439. Carbonell, E. P., and Brown, M. W., 1986, A Study of Short Crack Growth in Torsional LowCycle Fatigue for a Medium Carbon Steel, Fatigue Fract. Engng. Mater. Struct. 9, pp. 15-33. Chopra, O. K., 1999, Effects of LWR Coolant Environments on Fatigue Design Curves of Austenitic Stainless Steels, NUREG/CR5704, ANL98/31. Chopra, O. K., and Gavenda, D., J., 1998, Effects of LWR Coolant Environments on Fatigue Lives of Austenitic Stainless Steels, J. Pressure Vessel Technol. 120, pp. 116121.

Chopra, O. K., and Muscara, J., 2000, Effects of Light Water Reactor Coolant Environments on Fatigue Crack Initiation in Piping and Pressure Vessel Steels, Proc. 8th Intl. Conference on Nuclear Engineering, 2.08 LWR Materials Issue, Paper 8300, American Society of Mechanical Engineers, New York. Chopra, O. K., and Shack, W. J., 1997, Evaluation of Effects of LWR Coolant Environments on Fatigue Life of Carbon and LowAlloy Steels, Effects of the Environment on the Initiation of Crack Growth, ASTM STP 1298, W. A. Van Der Sluys, R. S. Piascik, and R. Zawierucha, eds., American Society for Testing and Materials, Philadelphia, pp. 247266. Chopra, O. K., and Shack, W. J., 1998a, Effects of LWR Coolant Environments on Fatigue Design Curves of Carbon and Low-Alloy Steels, NUREC/CR-6583, ANL97/18. Chopra, O. K., and Shack, W. J., 1998b, Low-Cycle Fatigue of Piping and Pressure Vessel Steels in LWR Environments, Nucl. Eng. Des. 184, pp. 49-76. Chopra, O. K., and Shack, W. J., 1999, Overview of Fatigue Crack Initiation in Carbon and Low-Alloy Steels in LWR Environments, J. Pressure Vessel Technol. 121, pp. 4960. Chopra, O. K., and Smith, J., L., 1998, Estimation of Fatigue StrainLife Curves for Austenitic Stainless Steels in Light Water Reactor Environments, Fatigue, Environmental Factors, and New Materials, PVP Vol. 374, H. S. Mehta et al., eds., American Society of Mechanical Engineers, New York, pp. 249259. de los Rios, E. R., Navarro, A., and Hussain, K., 1992, Microstructural Variations in Short Fatigue Crack Propagation of a C-Mn Steel, Short Fatigue Cracks, ESIS 13, Mechanical Engineering Pub., pp. 115-132. de los Rios., Wu, X. D., and Miller, K. J., 1996, A MicroMechanics Model of Corrosion-Fatigue Crack Growth in Steels, Fatigue Fract. Engng. Mater. Struct. 19, pp. 1383-1400. Dowling, N. E, 1977, Crack Growth During Low-Cycle Fatigue of Smooth Axial Specimens, ASTM STP 637, pp. 97-121. Eason, E. D., Nelson E. E., and Gilman, J. D., 1994, Modeling of Fatigue Crack Growth Rate for Ferritic Steels in Light Water Reactor Environments, Changing Priorities of Code and Standards, PVP 286, ASME, New York, pp. 131142. Ford, F. P., 1996, Quantitative Prediction of Environmentally Assisted Cracking, Corrosion 52, pp. 375-395. Ford, F. P., and Andresen, P. L, 1995, Corrosion in Nuclear Systems: Environmentally Assisted Cracking in Light Water Reactors, Marcel Dekker Inc., pp. 501-546. Ford, F. P., Ranganath, S., and Weinstein, D., 1993, Environmentally Assisted Fatigue Crack Initiation in

25-13

LowAlloy Steels A Review of the Literature and the ASME Code Requirements, EPRI Report TR102765. Gavenda, D. J., Luebbers, P. R., and Chopra, O. K., 1997, Crack Initiation and Growth Behavior of Carbon and Low-Alloy Steels, Fatigue and Fracture 1, PVP 350, ASME, New York, pp. 243-255. Hnninen, H., Trrnen, K., and Cullen, W. H., 1986, Comparison of Proposed Cyclic Crack Growth Mechanisms of Low Alloy Steels in LWR Environments, Proc. 2nd Int. Atomic Energy Agency Specialists' Meeting on Subcritical Crack Growth, NUREG/CP0067, MEA 2090, Vol. 2, pp. 7397. Higuchi, M., 1999, Fatigue Curves and Fatigue Design Criteria for Carbon and LowAlloy Steels in High Temperature Water, Probabilistic and Environmental Aspects of Fracture Mechanics, PVP Vol. 386, S. Rahman, ed., American Society of Mechanical Engineers, New York, pp. 161169. Higuchi, M., and Iida, K., 1991, Fatigue Strength Correction Factors for Carbon and LowAlloy Steels in Oxygen Containing HighTemperature Water, Nucl. Eng. Des. 129, pp. 293306. Higuchi, M., Iida, K., and Asada, Y., 1997, Effects of Strain Rate Change on Fatigue Life of Carbon Steel in High Temperature Water, Effects of the Environment on the Initiation of Crack Growth, ASTM STP 1298, W. A. Van Der Sluys, R. S. Piascik, and R. Zawierucha, eds., American Society for Testing and Materials, Philadelphia, pp. 216231. Hobson, P. D., 1982, The Formulation of a Crack Growth Equation for Short Cracks, Fatigue Fract. Engng. Mater. Struct. 5, pp. 323-327. Hussain, K., Tauqir, A., Hashmi, S. M., and Khan, A. Q., 1994, Short Fatigue Crack Growth Behavior in FerriteBainitic Steel, Metall. Trans. A 25A, pp. 2421-2425. Kitagawa, H., and Takahashi, S., 1976, Applicability of Fracture Mechanics to Very Small Cracks of the Cracks in the Early Stage, Proc. 2nd Int. Conf. on Mechanical Behavior of Materials, ASM, pp. 627-631. Lankford, J., 1977, Initiation and Early Growth of Fatigue Cracks in High Strength Steels, Eng. Fract. Mech. 9, pp. 617-624. Lankford, J., 1982, The Growth of Small Fatigue Cracks in 7075-T6 Aluminum, Fatigue Fract. Engng. Mater. Struct. 5, pp. 233-248. Lankford, J., 1985, The Influence of Microstructure on the Growth of Small Fatigue Cracks, Fatigue Fract. Engng. Mater. Struct. 8, pp. 161-175. Logsdon, W. A., and Liaw, P. K., 1985, Fatigue Crack Growth Rate Properties of SA508 and SA533 Pressure Vessel Steels and Submerged Arc Weldments in Room and Elevated Temperature Air Environments, Eng. Frac. Mech. 22, pp. 509-526.

Miller, K. J., 1985, Initiation and Growth Rates of Short Fatigue Cracks, Fundamentals of Deformation and Fracture, Cambridge Press, pp. 477-500. Nagata, N., Sato, S., and Katada, Y., 1991, LowCycle Fatigue Behavior of Pressure Vessel Steels in High Temperature Pressurized Water, ISIJ Intl. 31 (1), pp. 106114. Obrtlik, K., Polak, J., Hajek, M., and Vasek, A., 1997, Short Fatigue Crack Behavior in 316L Stainless Steel, Int. J. Fatigue 19, pp. 471-475. O'Donnell, T. P., and O'Donnell, W. J., 1995, Stress Intensity Values in Conventional S-N Fatigue Specimens, International Pressure Vessels and Piping Codes and Standards, PVP 313, pp. 195-197. Ravichandran, K. S., 1997, Effects of Crack Aspect Ratio on the Behavior of Small Surface Cracks in Fatigue; Part II. Experiments on a Titanium (Ti-8Al) Alloy, Metall. Trans. A 28A, pp. 157-167. Smith, R. A., Liu, Y., and Garbowski, L., 1996, Short Fatigue Crack Growth Behavior in Waspaloy at Room and Elevated Temperature, Fatigue Fract. Engng. Mater. Struct. 19, pp. 1505-1514. Suh, C. M., Lee, J. J., and Kang, Y. G., 1990, Fatigue Microcracks in Type 304 Stainless Steel at Elevated Temperature, Fatigue Fract. Engng. Mater. Struct. 13, pp. 487-496. Suh, C. M., Yuuki, R., and Kitagawa, H., 1985, Fatigue Microcracks in a Low Carbon Steel, Fatigue Fract. Engng. Mater. Struct. 8, pp. 193-203. Taylor, D., and Knott, J. F., 1981, Fatigue Crack Propagation Behavior of Short Cracks; The Effect of Microstructure, Fatigue Fract. Engng. Mater. Struct. 4, pp. 147-155. Tokaji, K., and Ogawa, T., 1992, The Growth Behavior of Microstructurally Small Fatigue Cracks in Metals, Short Fatigue Cracks, ESIS 13, Mechanical Engineering Pub., pp. 85-99. Tokaji, K., Ogawa, T., and Harada, Y., 1986a, The Growth of Small Fatigue Cracks in a LowCarbon Steel; The Effect of Microstructure and Limitation of Linear Elastic Fracture Mechanics, Fatigue Fract. Engng. Mater. Struct. 9, pp. 205-217. Tokaji, K., Ogawa, T., Harada, Y., and Ando, Z., 1986b, Limitation of Linear Elastic Fracture Mechanics in Respect of Small Fatigue Cracks and Microstructure, Fatigue Fract. Engng. Mater. Struct. 9, pp. 1-14. Tokaji, K., Ogawa, T., and Osako, S., 1988, The Growth of Microstructurally Small Fatigue Cracks in a FerriteBainitic Steel, Fatigue Fract. Engng. Mater. Struct. 11, pp. 331-342.

25-14

26
ENVIRONMENTAL FATIGUE EVALUATION ON JAPANESE NUCLEAR POWER PLANTS
Hitoshi Ohata Plant Operation Department, Plant Management Headquarters The Japan Atomic Power Company Tokyo, Japan

26-1

ENVIRONMENTAL FATIGUE EVALUATION ON JAPANESE NUCLEAR POWER PLANTS

Hitoshi Ohata Plant Operation Department, Plant Management Headquarters The Japan Atomic Power Company Tokyo, Japan

ABSTRACT Fatigue evaluations considering LWR environmental effects on key components for the oldest LWR in Japan, i.e., Tsuruga Unit 1, were carried out based on its operating experiences. Tsuruga Unit 1 is a BWR/2, non-jet pump type plant (see Table-1). A couple of evaluation methods, i.e., Higuchi-Iida formula and the design fatigue curves provided in NUREG/CR-6260, were adopted to calculate environmental fatigue usage factors. The evaluation results showed that all calculated fatigue usage factors were well below the allowable limit of 1.0. From the evaluation results, it was concluded that the results are mainly due to the low frequency of forced shutdowns, and that environmental fatigue will not be a significant safety issue even taking into account 60 years of plant operation. In this paper, the results of low-cycle thermal fatigue evaluation for key components considering environmental effects are discussed, since it has been suggested that the fatigue life of components in the LWR water environment could be significantly shorter than that in an air environment. Table-1 Outline of Tsuruga Unit 1
Plant Owner Reactor Type Main Contractor Thermal Output (MWh) Electric Output (Mwe) Construction permit First Criticality Commercial Operation Tsuruga Unit 1 The Japan Atomic Power Company BWR/2 GE 1070 357 Apr. 22, 1966 Oct. 03, 1969 Mar. 14, 1970

components and plant availability-related components were selected. In accordance with MITI Code Notification No.501, which is similar to ASME Section III, low-cycle fatigue at those locations was evaluated without considering environmental effects. In the second step, the locations that could be subjected to environmental effects, or in contact with primary coolant, were selected from the locations selected in the first step. Then, they were evaluated considering environmental effects. In this evaluation, assuming that critical points in the air condition are also severe in the water environment, only the severest point of each selected location was evaluated. In accordance with the selection steps described above, the main selected locations for environmental fatigue evaluation were as follows. Reactor feed water nozzles (low alloy steel) Reactor feed water piping and main steam piping (carbon steel) Recirculation piping (austenitic stainless steel) Reactor core shroud and support (austenitic stainless steel and Alloy 600) PLR pump casing (austenitic stainless cast steel)

As an example of the evaluation locations, the reactor feed water nozzle is shown in Fig.1 and Fig.2.

METHODOLOGY EVALUATION

FOR

ENVIRONMENTAL

FATIGUE

LOCATIONS TO BE EVALUATED With respect to environmental fatigue evaluation, locations to be evaluated for Tsuruga Unit 1 were selected in accordance with the following steps. In the first step, referring to the evaluation results in the design stage of the components of recent BWR plants, locations that are subjected to low-cycle thermal fatigue among safety-related

26-3

1
No. Locations in RPV Vessel Head Spray Nozzle Core Spray Nozzle Instrumantation Nozzle Recirculation Water Inlet Nozzle CRD Housing Stub ICM Housing Penatration

Table-2 Thermal transient cycles for Tsuruga Unit 1 (BWR)


1 2 Material Type SUS SUS SUS SUS Alloy 600 SUS SUS SUS LAS LAS SUS SUS Severity of Thermal Transient Medium Small Small Small Selected Location

Design basis thermal transient event

Design basis thermal transient cycles

actual thermal transient cycles in the past (before the end of

predicted thermal transient cycles after 60 years of operation 69 375(262) 208 (133) 93 (56)

12 11 10 9 8 7 6 4 5 2 3

3 4 5 6 7 8

FY1993)
Small Small Small Small Small

RPV Stud Bolt tensioning Hydraulic pressure test Start-up (Heat-up) Low power operation at night (no smaller than 75%) Low power operation at weekend (less than 75%) Rod pattern change Loss of feed water heating (by turbine trip)

123 400 240 10,000

28 162 (49) 92 (17) 45 (8)

Isolation Condenser Nozzle 12 Main Steam Nozzle 11

Small Small

Figure-1 Selection of the Evaluation Location on RPV in Tsuruga Unit 1 (BWR)

Nozzle (LAS)

Safe End (CS) C/ L Thermal Sleeve (Alloy 600)

Severest Point at Feedwater Nozzle Thermal Sleeve (Stainless Steel)

Figure-2 Evaluation Point on RPV Feedwater Nozzle in Tsuruga Unit 1 (BWR)

THERMAL TRANSIENT CYCLES For Tsuruga Unit 1 case, thermal transient cycles for environmental fatigue evaluation were determined as following. First, various thermal transient events which had occurred in Tsuruga Unit 1 since the pre-operational test period, were classified into the design basis thermal transient events which were usually utilized in the thermal fatigue evaluation in the design stage. Then, the average frequency or event cycles per year for each of the design basis thermal transient events was calculated. Finally, the predicted number of event cycles after 60 years of operation on each design basis event was obtained by linear extrapolation. Both the design basis thermal transient cycles and the predicted thermal transient cycles in Tsuruga Unit 1 are shown in Table-2. This table shows that all of the predicted thermal transient cycles after 60 years of operation are less than the design basis cycles.

26-4

'

Reactor Core Differntial Pressure mesuring Nozzle Recirculation Water Outlet Nozzle 9 CRD Water Return Nozzle 10 Feed Water Nozzle

Large

2,000

71 (26)

167 (122)

400 10

24 (15) 0 (0)

60 (51) 0 (0)

Loss of feed water heating (by feed water heater

70

0 (0)

0 (0)

bypass) Reactor scram (by turbine trip etc.) Reactor scram (by feed water pump trip) Reactor scram (by improper SRV actuation) Shutdown (Cooling down) RPV Stud Bolt 238 123 92 (17) 28 208 (133) 69 2 0 (0) 0 (0) 10 4 (0) 9 (5) 187 58 (9) 90 (41)

detensioning Note: Figures in parentheses are applicable to the evaluation for the RPV feed water nozzles. Since the inner surface metal of RPV feed water nozzles where thermal fatigue had been accumulated was removed in 1985, predicted thermal transient cycles after 60 years of operation were determined by using the actual thermal transient cycles after metal removal..

EVALUATION BASED ON MITI CODE NOTIFICATION NO.501 Fatigue evaluation without considering environmental effects for Tsuruga Unit 1 was conducted in accordance with MITI Code Notification No.501, as in the design stage. ENVIRONMENTAL FATIGUE EVALUATION METHODOLOGY The Higuchi-Iida formula was primarily selected as an environmental fatigue evaluation method [1]. In case that the HiguchiIida formula is not applicable because material is carbon steel or low alloy steel, or there is no data on the strain rate, the method in NUREG/CR-6260 was applied [2]. In the next article, details of their applicability of each method are described. EVALUATION BASED ON HIGUCHI-IIDA FORMULA The Higuchi-Iida formula is applicable to carbon steel and low alloy steel. In addition, it is necessary to know the corresponding strain rate at the location to be evaluated in order to calculate Fen

(fatigue life reduction factor for environmental effects) in the HiguchiIida formula by using the results of stress analysis. As a result, in the environmental fatigue evaluation for Tsuruga Unit 1, Higuchi-Iida formula was only applied to reactor feed water nozzle (low alloy steel). EVALUATION BASED ON NUREG/CR-6260 NUREG/CR-6260 is applicable to all kinds of material, therefore, it is applicable to all evaluation locations, including the locations where the Higuchi-Iida formula was not applied. As a result, NUREG/CR-6260 was applied to the components which were made of austenitic stainless steel and Alloy 600, e.g., reactor core shroud and its support. In addition, NUREG/CR-6260 was also applied to the locations where data on strain rates were not available, e.g., recirculation system piping, feed water system piping and main steam piping. EVALUATION RESULTS After the selection of the locations to be evaluated, determination of the predicted number of thermal transient cycles after 60 years of operation, and the selection of environmental fatigue evaluation method, the environmental fatigue calculation was carried out (see Table-3). The results showed that all of the calculated CUFs, considering environmental effects for Tsuruga Unit 1, were well below the allowable limit of 1.0 (see Table-4). Table-3 Evaluation locations and applied evaluation methods for Tsuruga Unit 1
Component/ Evaluation location Reactor Pressure Vessel (RPV)/ Feed water nozzles (low alloy steel) Selected reason Severest point in RPV from the viewpoint of fatigue Applied evaluation method and evaluating condition Method Temperature (C) DO (ppm) Strain rate (%/sec.) Primary system piping/ T joint, downstream of PLR pump outlet valve (austenitic Strain rate (%/sec.) Highest as inside to point peak Method Temperature (C) DO (ppm) Strain rate (%/sec.) 0.1 Saturated curve in Fig. 3-5 stainless steel) Main steam system piping/ Main steam header tee (carbon steel) Containment Vessel Loop Highest as to point peak Method Temperature (C) DO (ppm) Recirculation (PLR) 0.1 DO 0.2 Higuchi-Iida formula 100 T 200

Table-4 CUFs after 60 years of operation considering environmental effects


CUFs after 60 years of Component/ Evaluation location RPV/Feed water nozzles PLR piping/ T joint, downstream of PLR pump outlet valve Main steam piping/ Main steam header tee inside Containment Vessel 0.449 operation considering environmental effects 0.057 0.146

DISCUSSION All of the calculated CUFs in the environmental fatigue evaluation conducted for Tsuruga Unit 1 were well below the allowable limit of 1.0. The reason for such favorable results is considered to come from the fact that Tsuruga Unit 1 has been operated steadily, for instance, the number of forced shutdowns were considerably less than design basis thermal transient cycles as shown in Table-2. Among the calculated CUFs considering environmental effects after 60 years of operation, the value of 0.449 for the main steam piping (carbon steel) in Table-4 seems to be relatively high. This is because, in this evaluation, the design fatigue curve in NUREG/CR6260 for single-phase water was utilized instead of that in a watersteam two phase condition. This produced a significant fatigue life reduction. Therefore, the actual CUF is believed to be less than the calculated CUF. Finally, as thermal transients utilized for stress analyses in the design stage are conservatively determined [3], it is believed that there is further margin to the limit from the viewpoint of fatigue. CONCLUSION From the results of environmental fatigue evaluation for the oldest LWR in Japan, i.e., Tsuruga Unit 1 (BWR), it is concluded that the low-cycle fatigue considering environmental effects will not be a significant safety issue even taking into account 60 years of plant operation. REFERENCES [1] M. Higuchi and K. Iida, Fatigue strength correction factors for carbon and low-alloy steels in oxygen-containing high-temperature water Nuclear Engineering and Design 129 (1991) 293-306, NorthHolland [2] A. G. Ware, D. K. Morton, M. E. Nitzel, Application of NUREG/CR-5999 Interim Fatigue Curves to Selected Nuclear Power Plant Components, NUREG/CR-6260, INEL-95/0045 [3] T. Sakai, K. Tokunaga, G. L. Stevens and S. Ranganath, Implementation of Automated, On-Line Fatigue Monitoring in a Boiling Water Reactor, PVP Volume 252, page 67-74, The 1993 ASME Pressure Vessel and Piping Conference, Denver, Colorado, 1993, ASME, New York

Mean value for a pair of strain peaks NUREG/CR-6260 Applicable to all temperatures Applicable to all DO levels 0.001%/s curve in Fig. 3-18 NUREG/CR-6260 288C

stress intensity

stress intensity

26-5

26-6

26-7

26-8

26-9

26-10

26-11

26-12

26-13

26-14

26-15

26-16

26-17

26-18

26-19

26-20

27
EVALUATION OF ENVIRONMENTAL EFFECTS ON FATIGUE LIFE OF PIPING
F.A. Simonen and M.A. Khaleel Pacific Northwest National Laboratory Richland, Washington 99352 D.O. Harris and D. Dedhia Engineering Mechanics Technology, Inc. San Jose, California 95129 D.N. Kalinousky and S.K. Shaukat U.S. Nuclear Regulatory Commission Washington, D.C. 20555

27-1

EVALUATION OF ENVIRONMENTAL EFFECTS ON FATIGUE LIFE OF PIPING

F.A. Simonen and M.A. Khaleel Pacific Northwest National Laboratory Richland, Washington 99352 D.O. Harris and D. Dedhia Engineering Mechanics Technology, Inc. San Jose, California 95129 D.N. Kalinousky and S.K. Shaukat U.S. Nuclear Regulatory Commission Washington, D.C. 20555

27-3

EVALUATION OF ENVIRONMENTAL EFFECTS ON FATIGUE LIFE OF PIPING

F.A. Simonen and M.A. Khaleel Pacific Northwest National Laboratory Richland, Washington 99352 D.O. Harris and D. Dedhia Engineering Mechanics Technology, Inc. San Jose, California 95129 D.N. Kalinousky and S.K. Shaukat U.S. Nuclear Regulatory Commission Washington, D.C. 20555

Abstract Recent data indicate that light water reactor environments can significantly reduce the fatigue resistance of materials, and thereby show that design fatigue curves may be unconservative for reactor coolant environments. This paper describes calculations of probabilities of fatigue failures for a sample of components from five PWR and two BWR plants. The calculations used an enhanced version of the pc-PRAISE probabilistic fracture mechanics code. A wide range of failure probabilities were predicted, with some components having end-of-life probabilities of through-wall crack of nearly 100 percent and others with probabilities less than 10 -6. Inclusion of environmental effects (water versus air) can increase calculated through-wall crack probabilities by as much as two orders of magnitude. Sources of the uncertainties in the calculated probabilities were identified as coming from assumptions made in the fracture mechanics model itself and also from the inputs to the model. Uncertain inputs include the design-basis data for cyclic stresses which could differ from the actual stresses occurring during plant operation, and assumptions regarding strain rates and environmental variables used for predicting of fatigue crack initiation and growth. This paper describes sensitivity calculations that address the various sources of uncertainty.

Introduction The American Society of Mechanical Engineers (ASME) Code Section III requires a fatigue evaluation of the components of the reactor-coolant pressure boundary. Aspects of the code fatigue methodology have come under review because recent test data indicate that the effects of light-water reactor (LWR) environments could significantly reduce the 27-4

fatigue resistance of materials and show that the ASME design fatigue curves may not always be conservative for nuclear power plant primary system environments. As a result Argonne National Laboratory (ANL) has developed revised fatigue curves based on laboratory test data from small, polished specimens cycled to failure in water with temperatures, pressures, and chemistries that simulate LWR conditions (NUREG/CR6335) (Keisler et al. 1995). Using curves developed by ANL, the Idaho National Engineering Laboratory (INEL) investigated the significance of the interim fatigue curves as published in NUREG/CR5999 (Majumdar et al. 1993) by performing deterministic fatigue evaluations for a sample of components in the reactor coolant pressure boundary of LWRs. Cumulative usage factors (CUFs) for each component were reported NUREG/CR-6260. ANL has also developed statistical models for estimating the effects of various material, loading, and environmental conditions on the fatigue life of these materials. Probabilistic fatigue strain versus life (S-N) data for carbon steel (CS), low-alloy steel (LAS), and austenitic stainless steels (SS) were published in NUREG/CR-6237. In the present paper the statistical models from the ANL work were used by Pacific Northwest National Laboratory to estimate the probability of fatigue-crack initiation. The objective was to calculate component-failure probabilities rather than deterministic usage factors. The methodology and results of the PNNL study are documented in NUREG/CR-6674. This NUREG report includes a description of the plants and components that are addressed by the fatigue analyses. This is followed by a discussion the fracture-mechanics methodology along with documentation of the probabilistic equations from ANL that were used to predict the number of cycles to crack initiation. Another section of the report focuses on the consequences of small and large leaks and describes how calculations were performed to estimate core damage frequencies (CDF). These evaluations of CDFs are outside the scope of the present paper. The final section of the NUREG report summarizes the results in terms of absolute and relative failure probabilities, giving particular attention to how these calculated probabilities differ for a 40-year versus 60year plant life. Failure probabilities for water versus air environments are then compared. Appendices to NRUGEG/CR-6674 describe actual inputs and results of the calculations. Details of the modified pc-PRAISE code and a review the accuracy of crack-tip stressintensity factors calculated by pc-PRAISE are documented in other appendices. Discussions of the fracture-mechanics model include assumptions made to account for the effects of through-wall stress gradients on crack propagation and describe methods used to estimate the fractions of through-wall cracks that become small leaks and large leaks. A final appendix of NURER/CR-6674 describes sensitivity calculations that evaluate effects of uncertainties in the fracture-mechanics calculations, and calculations that exercise the cracking linking model to show how calculated crack lengths change when the inputs to the linking model are changed.

27-5

The present paper begins with a summary of the probabilistic fracture mechanics methodology and the enhanced version of the pc-PRAISE computer code. Results and conclusions from the component specific calculations are then presented. The remainder of the paper is devoted to sensitivity calculations that are helpful in reconciling the relatively high values of calculated failure probabilities with more favorable service experience that has not shown evidence of cracking or leakage.

Methodology for Through-Wall Crack Calculations The calculations estimated the probability that fatigue cycles will result in through-wall cracks and leaks of various sizes in pressure boundary components of reactor coolant systems (RCSs) of PWR and BWR plants. Only the contribution of initiated fatigue cracks was addressed, which excluded the contributions of pre-existing cracks. The methodology consisted of two parts. The first part calculated the probability that a fatigue crack (3 mm deep) will initiate as a function of time over the life of the plant. The second part evaluated the probability that these initiated cracks will grow to become through-wall cracks. Additional calculations for probabilities of large versus small leaks and for core damage frequencies are described in NUREG/CR-6674. These calculations are beyond the scope of the present paper. Stress amplitudes and the numbers of stress cycles for the selected components during a 40-year plant life were taken directly from NUREG/CR-6260 (Ware et al. 1995). The types of transients and cyclic stresses for the 60-year plant life were assumed to be the same as those during the 40-year plant life. The 60-year number of accumulated cycles was calculated by multiplying the 40-year number of cycles by a factor of 1.5. The number of cycles to crack initiation was a function of the material type, water/air environment, temperature, dissolved oxygen content, sulfur content and strain rate. The material types were carbon steel, low-alloy steel, 304/316 austenitic stainless steel and 316NG stainless steel. The statistical models of NUREG/CR-6335 (Keisler et al. 1995) were used to calculate the number of cycles to crack initiation corresponding to given probabilities (or percentiles) of the material S-N curves. The ANL statistical distributions predicted the number of cycles to initiate a 3-mm crack for a given cyclic stress amplitude. The parameters of the lognormal probabilistic fatigue initiation curves were based on the ANL revised fatigue curves published in NUREG/CR6335 (Keisler et al. 1995). The crack propagation was assumed to start from a 3-mm deep initiated flaw, which can was then grown to the critical size (through-wall) for component failure. This 3-mm size was the estimated crack size that can give a measurable 25 percent load drop in the testing of standard fatigue specimens. Sensitivity calculations were performed to evaluate the effect of changing this crack depth from 3 mm to 2 mm or 4 mm. The resulting changes in the calculated probabilities of through-wall cracks were at most a factor of two. It was decided not to include the initial crack depth as a simulated variable, in part because the 27-6

uncertainty in the initial crack depth is indirectly captured by the statistical scatter in the data based on the load drop approach used to define crack initiation. The cyclic stress levels from the INEL report were used to calculate both crack initiation and crack growth. These stresses included effects of stress concentrations in a manner prescribed by the ASME Code approach of stress indices. In many cases the stress indices addressed very high local stresses (e.g. weld root stress concentrations) with values up to 2.0. Because such surface stresses are not indicative of internal stress levels remote from the stress concentration, adjustments were made for deeper cracks to account for the effects of the through-wall stress. Cyclic stresses were broken into a component of uniform stress and a component of thermal gradient stress in accordance with a set of rules. Fatigue Crack Initiation The present work used a crack-initiation model (NUREG/CR-6335) (Keisler et al. 1995). This model estimates the probability of initiating a 3-mm deep fatigue crack based on existing fatigue (S-N) data, foreign and domestic, for carbon, low-alloy and stainless steels used in the construction of nuclear power plant components. ANL Crack-Initiation Correlation Only data obtained on smooth specimens tested under fully reversed loading conditions were considered. A statistical distribution was fitted by ANL to S-N data to describe the scatter in the fatigue data. The ANL statistical distributions of cycles to initiate a 3-mm crack for a given cyclic stress were lognormal. The parameters of the probabilistic fatigue initiation curves were based on the ANL revised fatigue curves published in NUREG/CR6335 (Keisler et al. 1995). The equations for stainless steels included recent updates for fatigue life correlations provided by ANL. The equations for cycles to failure were coded into a Fortran subroutine for implementation into probabilistic fracture mechanic codes such as pc-PRAISE. The calling program needs to provide values for the stress amplitude, the material type, the sulfur content (for ferritic steels), temperature, whether the environment is water or air, oxygen content of the water, and the strain rate for the stress cycle. Figure 1 was generated from data obtained from a series of calls to the subroutine. Each of the curves corresponds to the indicated percentile of data having cycles to failure less than or equal to the indicated percentile. The solid curve of Figure 1 is the median or 50th percentile curve for cycles to crack initiation. Inputs for strain rates, oxygen, and sulfur were all assigned as bounding values that are unlikely to be present simultaneously at these maximum values for any given component. The enhanced version of pc-PRAISE addresses crack initiation at multiple sites by subdividing the pipe circumference into a set of 5.08-cm (2-in.)-long zones. The amplitude of cyclic stresses at each site can vary in a manner specified by user input such 27-7

that the fatigue cracks may initiate at some sites much sooner or later than at other sites. The model also assumes no correlation between the random scatter in crack initiation times from one site to the next. Thus, a different selection from the family of S-N curves (as shown by the example of Figure 1) is sampled at random for each of the various sites around the pipe circumference. Treatment of Size Effects The equations developed by ANL to predict probabilities of fatigue-crack initiation are based on a statistical treatment of data from small specimen tests. An additional term of ln(4) is included to bring the equation into empirical agreement with some test data on 22.86-cm (9-in.)-diameter vessels. This term is intended to account for size, geometry, and surface-finish differences between small fatigue test specimens and actual components. The present calculations made use of the ANL equation, including the ln(4) term, for those cases in which the model assumed only one initiation site. However, the revised pcPRAISE model accounts for multiple initiation sites with each site covering some 5.08 cm (2 in.) of the pipe circumference. The probability of crack initiation therefore increases as the number of specified initiation sites is increased. This means that the fracture-mechanics model itself indirectly accounts for size effects, and inclusion of the ln(4) term in the ANL equation can result in a double counting of size effects. The ln(4) term of the ANL equation was modified when used to address crack initiation at multiple sites. The pc-PRAISE multiple-site model was calibrated to achieve agreement of calculated cycles to crack initiation with experimental data from the tests of the 22.86-cm (9-in.)-diameter vessels described in the ANL reports. The conclusion from this calibration effort was that the cycles to failure from the ANL equation needed to be increased by a factor of about 3.0. The net result was a factor of 3/4 applied to the number of cycles to failure from the small specimen data. In contrast, the ANL equation uses a factor of 1/4, but bases the fatigue-life prediction on consideration of a single initiation site. Fatigue Crack Growth As for the calculations for crack initiation, the calculations for fatigue crack-growth were based on data that accounted for the effects of environment on the growth rates. The equations for fatigue crack growth rates were unchanged from the prior version of the pcPRAISE code (Harris and Dedhia 1992). These equations did not address the specific factors that enhance the crack-growth rates in the same level of detail and rigor as in the Argonne correlations for crack initiation.

27-8

Treatment of Correlations It was assumed that the random variations in fatigue crack-growth rates were not correlated with the corresponding random variations in the cycles to crack initiation. If such correlations were to exist, the predictions for probabilities of through-wall cracks could be unconservative as indicated by sensitivity calculations described below. The simplifying assumption greatly facilitated the calculations, and has a good technical basis because the technical literature (Wire and Li 1996) provides evidence to support the assumption of independence. In general, crack initiation and crack growth involve different material damage mechanisms, such that environment and loading rates affect the mechanisms for crack initiation and growth differently. It was assumed that crack growth occurred under conditions of zero R-ratio. While this assumption will be conservative for transients with very high stress amplitudes, crackgrowth rates for cases of high cycle/low stress fatigue could be underestimated. The pcPRAISE model for predicting multiple crack initiations from site-to-site around the circumference of a given weld also assumed that random variations in the number of cycles to crack initiation were uncorrelated from site-to-site. Treatment of Through-Wall Stress Gradients The inputs for cyclic stress were the same stresses that were used in the NRC-funded research project at INEL as described in NUREG/CR-6260 (Ware et al. 1995). The data gave only peak stresses for the surface locations at which the initiation of fatigue cracks was to be evaluated and did not describe the corresponding variations of the stresses through the section thickness. It was appropriate to use these peak stresses for the initiation aspect of fatigue cracking. However, it was unrealistic to always assume that these peak stresses were uniformly distributed through the component thickness. The pc-PRAISE calculations used a number of conservative assumptions and inputs, which were in part balanced by unconservative assumptions and inputs. The inputs for stress cycles were taken from the INEL report (NUREG/CR-6260 [Ware et al. 1995]) and are believed in most cases to conservatively bound the stresses experienced during actual plant operation. These stresses assumed bounding conditions for the severity of thermal transients and other loads. The method used by INEL to derive load pairs from the transients assumed a worst-case sequencing of loads. The method used to estimate through-wall stress distributions (uniform tension versus through-wall gradient) was intended to overestimate the stress assigned to the uniform tension category. The approach used for the present calculations was to decompose the peak stress into a component of uniform stress and a component of through-wall gradient stress. Details of the approach are described in NUREG/CR-6674. A standardized (quadratic) stress gradient was developed on the basis of stress solutions for heating and cooling ramps and step changes in surface temperatures.

27-9

Since results of detailed stress calculations were not available from the work of NUREG/CR-6260 (Ware et al. 1995), rules were developed to assign a fraction of the peak stress to the uniform stress category. The remaining fraction was assigned to the through-wall gradient category. In many cases, the values of peak stresses were greater than 100 ksi, which implied that most of the stress was due to heating and cooling transients or were due to geometric stress concentrations. In other cases, the number of stress cycles was very large, which also suggested thermal transients. Another consideration was that the ASME code limits do not permit membrane stresses (including secondary stresses) to be more than three times the code design stress (i.e. < 3S m). For typical piping materials, the 3Sm limit implied that all stress ranges (or 2S a) greater than 45 ksi should be treated as gradient stresses. The rules used to assign stresses to the uniform and gradient categories were: seismic stresses were treated as 100 percent uniform stress. stresses greater than 45 ksi were treated as having a uniform component of 45 ksi with the remainder being assigned to the gradient category. For transients with more than 1000 cycles over a 40-year life, it was assumed that 50% of the stress was uniform stress and 50% through-wall gradient stress. In addition the uniform stress component was not permitted to exceed 10 ksi.

These rules were less conservative than assuming uniform stresses through the full component sections. The approach ensured that shallow initiated cracks would at first be subjected to the peak surface stresses, but allowed for a reduction in crack-growth rates as the cracks grow into regions of lower stress levels. Crack and Linking of Cracks at Multiple Sites The enhanced version of pc-PRAISE simulated the initiation of fatigue cracks at multiple sites around the circumference of a weld, a phenomenon often seen in service induced cracking of pipe welds. Details of the aspect of the fracture mechanics model are described in NUREG/CR-6674. The crack growth calculations simulate the growth (depth and length) of the individual cracks and combine adjacent cracks into a single larger crack in accordance with proximity rules. One of the sensitivity calculations described below provides an example of the crack linking.

27-10

Treatment of Inservice Inspections The calculations of through-wall crack frequencies did not account for potential benefits of inservice inspections or maintenance programs, even though the predicted cumulative probabilities of failure for many of the components attained levels late in plant life that exceeded 50 percent. On the other hand, the pc -PRAISE model does take credit for l eak detection, but this only decreased the probability that short through-wall cracks with inconsequential leaks will grow over time to cause much larger leaks. Leak detection had no effect on the probabilities of through-wall cracks. Calculations for Selected Components PNNL performed calculations for the five pressurized-water reactor (PWR) plants and two boiling-water reactor (BWR) plants listed in Table 1 and for the components listed in Table 2. Inputs to Calculations The fracture-mechanics calculations were based on data compiled by INEL from stress reports for actual plants. However, NUREG/CR-6260 (Ware et al. 1995) did not reveal the identities of these plants. For each plant, four to nine locations were investigated, including locations within the reactor pressure vessel. Calculations with the pc-PRAISE predicted probabilities of crack initiation and probabilities of through-wall cracks as a function of time for plant operating periods up to 60 years. For the PWR plants, the curves for high-sulfur steel (0.015 weight percent) and a lowoxygen environment (0.01-ppm) were used. For the BWR plants, the curves for high sulfur steel and a high-oxygen environment (0.10 ppm) were used. The strain rates for both PWR and BWR components (low alloy and carbon steel) were assumed to be 0.001% (see NUREG/CR-6260) (Ware et al. 1995). For 316 stainless steel, the strain rate was 0.004%. For all components, the temperature was assumed to be 290 C. Summary of Results Table 3 provides the final results for all the components. Results are given for both a 40year and a 60-year operating period. Many of the components have cumulative probabilities of both crack initiation and through-wall cracks that approach unity. Other components, often with similar values of fatigue usage factors, show much lower failure probabilities. As discussed in NUREG/CR-6674 the maximum failure rate (through-wall cracks per year) is about 510-2, and the maximum core-damage frequency based on these calculated failure rates is about 1.010-6 per year. These maximum values do not change significantly from 40 years to 60 years. In contrast, failure rates for other components with much lower baseline failure rates are seen to increase by as much as an order of magnitude from 40 years to 60 years.

27-11

Correlation of Failure Probabilities with Usage Factors Figure 2 shows the correlation between calculated probabilities of through-wall cracks for each of the 47 components with the fatigue usage factors reported in NUREG/CR-6260 (Ware et al. 1995). The correlation is only approximate because the usage factors address only crack initiation. As such the usage factor calculations do not address factors that determine how likely each initiated crack will grow to become a through-wall flaw. The CUFs as reported in this paper were taken directly from the INEL work of NUREG/CR-6260 (Ware et al. 1995), which made use of the fatigue (S-N) curves of NUREG/CR-5999 (Majumdar et al. 1993). These curves accounted for environmental effects and predicted significant reductions in life compared to the fatigue curves of the ASME code. As such, the INEL fatigue usage factors for the sample components are generally greater than the usage factors calculated when the components were originally designed. The plots of Figure 2 indicate a potential for through-wall cracks (probability of failure greater than say >10-1) even for usage factors less than one. Usage factors greater than one can sometimes correspond to essentially 100 percent failure probability. On the other hand, for usage factors of 0.1 or less Figure 2 indicates that the probabilities of failure become relatively low (10-3 or less). These overall trends are consistent the viewpoint that code usage factors should not be treated as precise predictors of cycles to fatigue failure, but rather as a method to establish acceptable designs. It should furthermore be noted that plant operating experience has shown few if any fatigue failures for the types of loading conditions addressed by the design calculations. Instead, fatigue failures have generally been due to stresses (vibration, thermal fatigue, etc.) that were not anticipated during plant design. Probabilities at 60 Years Versus 40 Years Figure 3 shows trends of the calculated results of Table 3. The data points above the dashed diagonal line indicate the wide range of the calculated failure probabilities and compare the failure probabilities at the end of a 60-year plant life with the corresponding probabilities for a 40-year plant life. The range of the through-wall crack probabilities is about seven orders of magnitude. The failure probabilities corresponding to a 60-year plant life can be a factor of 10 or greater than those for a 40-year plant life. It should be noted that the increases are greatest for components that have relatively small values at 40 years. In contrast, there are only small increases when the 40-year probabilities are already quite large. In these cases, the fatigue cracks initiate relatively early in life and there is a high potential of leaking before the end of a 40-year operating period. However, it is unlikely in practice that such high levels of fatigue damage would go unnoticed and unmitigated at a significant number of fatigue sites. It is more likely that corrective action programs consisting of augmented inspections, repairs and replacements, along with changes to 27-12

plant operating practice would be implemented before the end of a 40-year operating period. Such programs would significantly decrease the failure frequencies from those calculated here. Water Versus Air Environment The main objective of the PNNL study was to compare predicted probabilities of fatigue failures for a 60-year life versus a 40-year life with the effect of reactor coolant environments included in both evaluations. However, additional calculations were performed to establish the extent of environment effects independent of the issue of 40 years versus 60 years pant life. The data points plotted below the dashed diagonal line of Figure 3 compare the calculated fatigue lives for water versus air environments. The 40-year life for the water environment was used as the baseline case. The data show that changing to an air environment gives lower probabilities of through-wall cracks (by about a factor of 100). In contrast, changing from a 40-year life to a 60-year life increased the probabilities, but the relative increase (a factor of about 10) is not nearly as large as that associated with the environmental effects. Sensitivity Calculations for Surge Line Elbow A detailed study was performed for the surge-line location in the newer-vintage Combustion Engineering plant. This was the location with the highest calculated failure probability of any of the 47 locations addressed by PNNL. Figure 4 presents the failure probabilities predicted by pc-PRAISE for the surge-line elbow in terms of probabilities of crack initiation and of through-wall cracks, both as a function of time. It is seen that cracks can initiate rather early in the plant life, with about a 50-percent probability of initiating a fatigue crack after only 10 years of operation. Over this same 10 year period, about 50 percent of the initiated cracks are predicted to become through-wall cracks. The frequency of through-wall cracks (lower curve) increases significantly over this 10 year period and then remains relatively constant over the remainder of the 60-year plant life. Figure 5 evaluates effects of the critical inputs that specify through-wall stress gradients. It was recognized that peak surface stresses, apply only to the initiation of fatigue cracks, and are not representative of the stresses that grow these initiated cracks through the thickness of the pipe wall. A range of assumptions on stress gradients was made for the calculations of Figure 5. It should be emphasized that the probabilities of crack initiation remained the same for all the calculations because the cyclic surface stresses that govern crack initiation were the same for all cases. The solid curve of Figure 5 shows the baseline through-wall crack probabilities as reported in Table 3. This calculation assumed that peak stresses greater than 45 ksi 27-13

should be treated as thermal gradient stresses. The most conservative assumption would treat the peak stress as entirely uniform tension stress. As indicated by Figure 5, the more conservative assumption increased the calculated failure probabilities by a factor of about 2.0. The other bounding assumption was to treat the peak stress as 100 percent thermal gradient. This reduced the failure probabilities by a factor between five and ten. Perhaps the most realistic assumption addressed by Figure 5 considered all stresses greater than 15 ksi as thermal gradient stresses. This assumption decreased the through-wall crack probabilities by a factor of about 2.0 relative to the baseline case.

A Study of Crack Lengths and Linking as Predicted by pc-PRAISE The pc-PRAISE model for fatigue crack initiation was applied to simulate the initiation, growth and linking of thermal fatigue cracks for a small diameter pipe. The objective was to demonstrate the ability of pc-PRAISE to predict realistic lengths for circumferential cracks. A second objective was to perform sensitivity calculations to evaluate the effects of modeling assumptions and alternative inputs. The calculated crack lengths were then compared to the size and shape of the cracking reported for a small diameter pipe of the high-pressure injection system at the Oconee 2 plant (USNRC 1997). The calculations addressed a stainless steel pipe with an inner diameter of 2.9 inch and a wall thickness of 0.3 inch. The baseline case was intended to correspond to inputs and assumptions used for the 47 PWR and BWR components of the PNNL study. The following describes the baseline case: 100% of the cyclic stress assigned to the thermal gradient category no circumferential variation of cyclic stress cycles to crack initiation were sampled independently at each circumferential site 5 sites for crack initiation around the circumference of pipe 1 percent probability that the length of the initiated fatigue crack will exceed the length of the initiation site

Table 4 lists the calculations for five variations of the input parameters from those of the baseline case. The magnitudes and numbers of the cyclic stresses were assigned to give calculated probabilities of through-wall cracks that approached 100 percent. The objective was not to predict failure probabilities per se, but to predict the circumferential extent of cracking that develops late in the life of a highly stressed component. It was not possible to make a direct comparison of the probabilistic calculations with the single observed case of Oconee-2 pipe failure (see Figure 6). The comparisons were 27-14

therefore based on calculated distributions of cracks corresponding to the time at which the probability of through-wall cracking attained a value of 50 percent. A further complication was that the ANL fatigue correlations assume that initiated fatigue cracks start immediately with a depth of 3 mm. For the small 3-inch diameter pipe, this 3mm crack is about 40 percent of the pipe wall. Much of the cracking in the Oconee-2 event had depths less than 30 percent of the pipe wall. The output of the computer code fails to define the depths of these small cracks, prior to such time that they attain the threshold depth of 3-mm. Therefore, that portion of the Oconee cracking with depths less than 3 mm was treated as uncracked in the context of the approach taken by the pcPRAISE fracture mechanics model. The crack length at the outer surface of the Oconee-2 pipe had become 21 percent of the pipe circumference (Figure 6) when the leak rate caused the plant operators to bring the plant into a shutdown mode. The pipe also had part-through cracking around the remaining circumference. About 47 percent of the pipe circumference had cracking that exceeded the 3-mm threshold of the ANL correlations. The columns of Table 4 further summarize trends from the detailed computer output. Several global measures of the extent of circumferential cracking are used to describe the circumferential cracking as indicated by the column headings. The baseline case of Table 4 predicts that 40 to 60 percent of the pipe circumference will be cracked (at a probability of about 50 percent), with the percentage depending on the measure selected to described the circumferential cracking. The calculated percentages are generally consistent with the observed cracking of the Oconee-2 pipe. The higher numbers (60%) of the final two columns of Table 4 are based on a weld-by-weld summary of the simulated data. These tabulations characterize only the total amount of circumferential cracking and do not consider whether this cracking is in the form of one large crack or from the sum of several smaller unconnected cracks. The second calculation assumed that part of the cyclic stress is uniform tension (i.e. 20%) rather than a pure through-wall thermal gradient stress. The predicted probability for long cracks decreases markedly, suggesting that the cracking pattern of the Oconee failure is characteristic of a pure thermal gradient stress. The third calculation of Table 4 increased the number of potential crack initiation sites from 5 to10. The length of the sites decreased from about 2 inch to about 1 inch. There is little predicted change in the overall amount of circumferential cracking, but the lengths of the individual cracks tended to decrease. This means more cracks with the average length of each crack becoming shorter. The fourth calculation of Table 4 kept the number of initiation sites at five, but increased the assigned probability that an initiated crack will span the entire 2-inch length of the initiation site. The median length of the initiated crack was not changed from the baseline value of 0.6 inch. Table 4 shows little change in the calculated probabilities for larger lengths of circumferential cracking. 27-15

The fifth calculation of Table 4 assumed a perfect correlation between the number of cycles to crack initiation for all sites in a given weld. That is , if one of the five sites becomes cracked, all of the other sites will crack at the same time. This assumption ignores the characteristic scatter in fatigue data. The predictions of Table 4 show very high probabilities for cracking around a large fraction of the pipe circumference. The final calculation of Table 4 expands on the previous calculation. The site-to-site randomness of fatigue lives is again taken to be zero, but there is a 20% circumferential variation in the cyclic stress level. The circumferential variation gives a modest reduction in the probability for cracking around large fractions of the pipe circumference compared to the previous case. In summary the pc-PRAISE predictions have been compared to the cracking observed at Oconee-2. The service failure had deep cracking over 20 to 50 percent of the pipe circumference, whereas the pc-PRAISE baseline calculation predicted deep cracking over some 40 to 60 percent of the pipe circumference. It is concluded that the predicted cracking is generally consist and somewhat conservative relative to `the observed cracking.

Sensitivity Calculations Using Latin Hypercube Method A comprehensive set of calculations for all locations of the seven plants addressed the effects of changing various inputs and modeling assumptions. These calculations applied the Latin hypercube method described in an ASME paper (Khaleel and Simonen 1995), which permitted rapid calculations even for components with very small failure probabilities. The streamlined Latin hypercube method was successfully benchmarked against the pc-PRAISE code. Baseline Parameters Inputs for the baseline cases of the sensitivity calculations were essentially the same as the inputs later used for the final pc-PRAISE runs of the NUREG/CR-6674 report. Initial Flaw Depth The baseline depth of the initiated crack was assigned a deterministic value of 3 mm. Variations from this depth were part of the sensitivity calculations. Correlations Between Crack Initiation and Crack Growth The baseline case assumed that the random variations in crack growth rates (da/dN) were independent of random variations in the number of cycles to crack initiation. This assumption was used for the calculations with the pc-PRAISE code. Sensitivity calculations addressed the effect of a perfect correlation between crack initiation and crack growth. These calculations used the same random number to sample from the distributions for crack initiation and crack growth. Oxygen Content of Reactor Water The baseline cases used a reactor water oxygen content of 0.010 ppm (10 ppb) for PWR plants and 0.100 ppm (100 ppb) for BWR plants. 27-16

These values were considered to be realistic levels but somewhat higher than expected for typical plant operating conditions. Sensitivity calculations considered somewhat lower and more typical values for water chemistries. Sulfur Content The baseline value of sulfur content was 0.015 weight percent for low alloy steels. For stainless steels the sulfur content does not appear in the equations for fatigue crack initiation. Sensitivity calculations considered the effects of somewhat lower and more typical values of sulfur than the bounding value of 0.015 percent. Strain Rate Lower strain rates result in fewer predicted cycles to crack initiation. The baseline cases assumed a common strain rate of 0.001 percent per second for all components and all transients. Somewhat lower values were used for sensitivity calculations. Sensitivity to Initial Flaw Depth The ANL fracture mechanics model defines crack initiation as a surface flaw with a depth of about 3 mm. This depth was related to the 25 percent load drop method used to detect the presence of a crack in the fatigue tests. The actual depth of the crack could be less than or greater than 3 mm. To address uncertainties in the initial crack depth, calculations were performed for crack depths of 2 mm and 4 mm. Each point on Figure 7 corresponds to one of the components from the seven representative plants. If initial flaw depth had no effect on the probability of through-wall cracks, then all points would fall on the diagonal line. However Figure 7 shows that the calculated probabilities of through-wall crack increases somewhat for the 4-mm crack and decreases somewhat for the 2-mm crack. The changes in through-wall crack probabilities are at most by a factor of 2, and are insignificant for components having relatively high failure probabilities. Simulation of the initial flaw depth as an additional stochastic variable was not considered appropriate, because 1) the calculated failure probabilities were relatively insensitive to the assumed value of flaw depth, and 2) uncertainties in the flaw depths corresponding to experimental data on cycles to crack initiation were considered to be adequately accounted for in the variability in cycles to crack initiation. Sensitivity to Initial Flaw Length or Aspect Ratio The baseline Latin hypercube model assumed that all initiated cracks had aspect ratios of 10:1 (a length of 30 mm for the initial flaw depth of 3 mm). It was also assumed that the aspect ratio of growing fatigue crack remained at 10:1 as the crack increased in depth. Figure 8 shows the effect of replacing the 10:1 aspect ratio with a ratio of 3:1. The smaller value was selected as bounding for growing fatigue cracks. The calculated probabilities of through-wall cracks decreased by a factor as great as 10. The differences were greatest 27-17

for low failure probabilities and were relatively small for components with very high failure probabilities. The results indicate that precise inputs for modeling of flaw lengths are not critical to the calculations of through-wall crack probabilities (provided that assumed values of aspect ratio are not taken to be unrealistically small) but are critical to probabilities of large leaks and breaks. Effect of Wall Thickness The results of Figure 9 were generated by arbitrarily changing the wall thickness of each component to 1.0 inch and then to 8.0-inch. Whereas the calculated probabilities of initiating a fatigue crack do not change, the thicker component was expected to have a greater fatigue life because more stress cycles are needed to growth the crack through the thicker metal path. The calculated results are consistent with this expectation. The differences in failure probabilities are about a factor of 10 for relatively low failure probabilities but become insignificant when the failure probabilities are relatively large. The results were based on a uniform through-thickness distribution of cyclic stress. The presence of large stress gradients will tend to offset the wall thickness effect seen in Figure 9. Effect of Through-Wall Stress Gradients The baseline case conservatively assumed a uniform distribution of stress through the wall thickness, which meant that the peak surface stress that governs crack initiation also is available to grow the small initiated crack to become a through-wall crack. However, for most stress transients, the peak surface stress is associated with stress gradients. Figures 10-13 show the sensitivity of calculated probabilities of through-wall cracks to the magnitude of the stress gradient. Figure 10 assumed a relatively modest gradient consisting of a linear distribution of stress such that the stress at the outer surface is 50 percent of the peak stress at the inner surface. This modest gradient has only a small effect (factor of 2 or less) in terms of decreasing the calculated probabilities of through-wall cracks. Figure 11 increased the linear stress gradient such that the outer surface stress became zero. In this case the calculated probabilities of through-wall cracks decreased by as much as a factor of 10 relative to the baseline case. Figure 12 was based on a rather extreme assumption whereby the stress used for calculating crack growth rates was assumed to be uniformly distributed through the wall thickness, but with this stress reduced to 50% of the peak surface stress used for the initiation of the crack. Such an assumption could approximate the situation where crack initiation is from very localized stress concentration. The predicted effect on through-wall crack probabilities is substantial and amounts to 3-4 orders for magnitude of cases with very low failure probabilities. The effect is much smaller for components with the higher failure probabilities, but gives a factor of about 10 reduction in failure probability.

27-18

Figure 13 was based on an assumed stress gradient that was a more realistic estimate of a stress gradient produced by a transient thermal stress. The predicted effect of this stress gradient that is about two orders of magnitude for lower failure probability components and about one order of magnitude for components with higher failure probabilities. It was concluded that realistic predictions of through-wall crack probabilities require realistic modeling of through-wall stress gradients. Correlation of Crack Initiation and Crack Growth The source code was modified to assume a perfect correlation between the random variations in cycles to crack initiation with the corresponding variations in the crack growth rates. Such correlations were expected to increase the probability that a crack which initiates at a relatively small number of stress cycles will subsequently grow at a faster than average crack growth rate. The data points of Figure 14 confirm this expectation. The correlation increases the calculated failure probabilities by up to an order of magnitude, with the effect being largest when the failure probabilities are small. At failure probabilities greater than 1.0E-01 the effect becomes negligible. Environment and Material Characteristics A number of inputs for environmental parameters must be assigned for application of the Argonne National Laboratory equations for fatigue crack initiation. The calculations of Figures 15-20 address the effects of uncertainties in these inputs on calculated probabilities of through-wall cracks. The inputs addressed in sensitivity calculations were the oxygen content of the reactor water, the sulfur content of the steel, and the strain rate associated with the cyclic stresses. The calculations did not address the effects of temperature. Figure 15 addresses the effect of oxygen content with the baseline being an oxygen level of 0.01 ppm that is typical of PWR conditions. The sensitivity calculations increased this level to 0.10 ppm (BWR conditions). It is should be noted that these calculations arbitrarily assigned the same oxygen level to all of the 47 components without regard whether they corresponded to a PWR or BWR plant. It is seen in Figure 15 that increasing the oxygen level over the selected range of uncertainty has at most an order of magnitude effect on calculated probabilities of through-wall cracks. In some cases the probabilities increased (ferritic steel components) and other cases the calculated probabilities decreased (stainless steel components). Figures 16 and 17 show the sensitivity of calculated failure probabilities to strain rates, and indicate that low strain rates can result in higher values of calculated failure probabilities. The default strain rate used in the baseline calculations was 0.001. Figure 16 shows the effect of using a strain rate of 0.01, which would correspond to an actual transient with a relatively rapid loading rate. Figure 15 shows that some of the calculated failure probabilities decreased by as much as a factor of 10. Failure probabilities for another grouping of components (ferritic steel) show little if any change. Figure 16 increased the 27-19

strain rate by a factor of 1000, and these results show more substantial decreases in calculated failure probabilities for the higher strain rate. Figure 18 shows the effect of reducing the sulfur content of the steel to 0.0 percent from 0.015 percent. Some failure probabilities show no increase (stainless steel) whereas other components show only a modest increase (less that a factor of 2.0). Figure 19 (low alloy steel) and Figure 20 (stainless steel) show the effect of changing the inputs from bounding values that govern environmentally assisted fatigue to more moderate values that may be more typical of actual plant operation. The calculations used baseline inputs of strain rates of 0.001 %/sec, oxygen of 100 ppb and sulfur of 0.015 percent. In constructing these figures the material type for all components was arbitrarily set to either low alloy steel (Figure 19) or stainless steel (Figure 20). Within the relatively small range of uncertainty (as considered here) the changes in calculated failure probabilities were relatively small (factor of 2 or less). The assumption in these calculations was that the total exclusion of environmental effects could not be justified for any of the components. Other sensitivity calculations of Figure 3 compared calculated through-wall crack probabilities for air environment versus probabilities for water environment, which shows differences by a factor of 10 or more. Summary and Conclusions Probabilities of fatigue failures have been estimated for RPV and piping components of five PWR and two BWR plants. These calculations were made possible by the development of a new version of the pc-PRAISE probabilistic fracture mechanics code that has the capability to simulate the initiation of fatigue cracks in combination with a simulation of the subsequent growth of these fatigue cracks. The calculations gave a wide range of failure probabilities for the selected components, with some components having end-of-life probabilities for through-wall cracks of nearly 100 percent and others with probabilities of less than 10 -6. It is recognized that there are uncertainties in the calculated failure probabilities. Uncertainties come from assumptions made in the fracture mechanics models and from the inputs to the models. In addressing these uncertainties, the intent of the baseline calculations was to perform best-estimate calculations. When best-estimate inputs were not available, the approach was to select conservative values. In particular, the inputs for cyclic stresses were based on design-basis data, which could differ from the stresses occurring during actual plant operation. Other areas of uncertainty included inputs for strain rates and environmental variables used to predict fatigue-crack initiation. Calculations were performed to address the effects of reactor water environments (versus air) and to compare these effects to the effects of extended plant operation from 40 years to 60 years. The environmental effects were predicted to increase through-wall crack probabilities by as much as two orders of magnitude.

27-20

The calculated through-wall crack probabilities for the components with the very highest probabilities of failure show essentially no increase in failure probability from 40 to 60 years. On the other hand, other components with lower fatigue uasage showed significant increases in failure probabilities. Here again, the increases associated with the waterenvironment are a factor of 10 or more greater than the corresponding increases associated with extended plant operation from 40 to 60 years. Further work is needed to reconcile the relatively high calculated failure probabilities for some components with the relatively good service experience for these components. It is expected that inputs based on better data on the frequency and severity of cyclic stresses in combination with a better characterization of environmental factors will remove most of the differences between predicted cycles to failure versus service experience. The higher values of though-wall crack probabilities were also based on conservative assumptions relative to the scenario that governs the inspection and maintenance of highfatigue locations. The model assumed that components with high-failure frequencies will remain in operation as components fail one-by-one or are retired at the end of plant life. In practice, the first failure of a group of similar components will cause other members of the group to be subject to an aggressive program of corrective actions such that the probability of repeat failures is greatly reduced. Such corrective actions would include frequent inservice inspections and changes to plant operational practices to reduce stress levels. The present fracture-mechanics model does not address the benefits of such corrective actions in reducing failure frequencies. Acknowledgments Work Supported by the U.S. Nuclear Regulatory commission under Contract DE-AC0676RLO1830; NRC JCN W6671. References Harris, D. O. and D. Dedhia. 1992 . A Probabilistic Fracture Mechanics Code for Piping Reliability Analysis (pc-PRAISE code), NUREG/CR-5864, U.S. Nuclear Regulatory Commission, Washington D.C. Keisler, J., O.K. Chopra and W. J. Shack. 1994. Statistical Analysis of Fatigue StrainLife Data for Carbon and Low-Alloy Steels, NUREG/CR-6237, U.S. Nuclear Regulatory Commission, Washington D.C. Keisler, J., O.K. Chopra, and W. J. Shack. 1995. Fatigue Strain-Life Behavior of Carbon and Low-Alloy Steels, Austenitic Stainless Steels, and Alloy 600 in LWR Environments, NUREG/CR-6335, U.S. Nuclear Regulatory Commission, Washington D.C.

27-21

Khaleel, M. A. and F. A. Simonen. 1995. A Model for Predicting Vessel Failure Probabilities Due to Fatigue Crack Growth, ASME PVP Vol. 304, pp. 401 -416, Fatigue and Fracture Mechanics in Pressure Vessels and Piping. Khaleel, M.A., F.A. Simonen, H.K. Phan, D.O. Harris and D. Dedhia 2000. Fatigue Analysis of Components for 60-Year Plant Life, NUREG/CR-6674, U.S. Nuclear Regulatory Commission, Washington D.C. Majumdar, S., O. K. Chopra, and W. J. Shack. 1993. Interim Fatigue Curves to Selected Nuclear Power Plant Components, NUREG/CR-5999, U.S. Nuclear Regulatory Commission, Washington D.C. USNRC 1997. Unisolable Reactor Coolant System Leak at Oconee-2 April 21, 1997, Licensee Event Report LER No. 270/97-001, U.S. Nuclear Regulatory Commission. Ware, A. G., D. K. Morton, and M. E. Nitzel. 1995. Application of NUREG/CR-5999 Interim Fatigue Curves to Selected Nuclear Power Plant Components, NUREG/CR6260, U.S. Nuclear Regulatory Commission, Washington D.C. Wire, G. L., and Y. Y. Li. 1996. Initiation of Environmentally-Assisted Cracking in Low Alloy Steels, Fatigue and Fracture Volume 1, PVP-Vol.323, pp. 269-289, American Society of Mechanical Engineers, New York.

27-22

Table 1. Plants Evaluated in the Fatigue Study


PWRs Babcock and Willcox (B&W) Combustion Engineering (CE) Newer Vintage CE Older Vintage Westinghouse (W) Newer Vintage W Older Vintage BWRs

General Electric (GE) Newer Vintage


GE Older Vintage

Table 2. Components Selected for Fatigue Analysis PWRs


1. Reactor pressure vessel shell and lower head. Reactor vessel inlet and outlet nozzles. Pressurize surge line (including hot leg and pressurizer nozzles). Reactor coolant piping charging system nozzle. Reactor coolant piping safety injection nozzle. Residual heat removal (RHR) system class 1 piping. 1. BWRs Reactor pressure vessel shell and lower head. Reactor vessel feedwater nozzle.

2.

2.

3.

3.

Reactor recirculation piping (including inlet and outlet nozzles). Core spray line reactor vessel nozzle and associated class 1 piping. Residual heat removal class 1 piping.

4.

4.

5.

5.

6.

6.

Feedwater line class 1 piping.

27-23

Table 3 Summary of Results for All Seven Plants Water Environment


PLANT CE-NEW CE-NEW CE-NEW CE-NEW CE-NEW CE-NEW CE-NEW CE-NEW CE-NEW CE-OLD CE-OLD CE-OLD CE-OLD CE-OLD CE-OLD CE-OLD B&W B&W B&W B&W W-NEW W-NEW W-NEW W-NEW W-NEW W-NEW W-OLD W-OLD W-OLD W-OLD W-OLD W-OLD W-OLD W-OLD GE-NEW GE-NEW GE-NEW GE-NEW GE-NEW GE-NEW GE-OLD GE-OLD GE-OLD GE-OLD GE-OLD GE-OLD GE-OLD COMPONENT RPV LOWER HEAD/SHELL RPV INLET NOZZLE RPV OUTLET NOZZLE SURGE LINE ELBOW CHARGING NOZZLE NOZZLE CHARGING NOZZLE SAFE END SAFETY INJECTION NOZZLE NOZZLE SAFETY INJECTION NOZZLE SAFE E SHUTDOWN COOLING LINE ELBOW RPV LOWER HEAD/SHELL RPV INLET NOZZLE RPV OUTLET NOZZLE SURGE LINE ELBOW CHARGING NOZZLE SAFE END SAFETY INJECTION NOZZLE SAFE E SHUTDOWN COOLING LINE ELBOW RPV NEAR SUPPORT SKIRT RPV OUTLET NOZZLE MAKEUP/HPI NOZZLE SAFE END DECAY HEAT REMOVAL/REDUCING T RPV LOWER HEAD/SHELL RPV INLET NOZZLE RPV OUTLET NOZZLE CHARGING NOZZLE NOZZLE SAFETY INJEC NOZZLE NOZZLE RESIDUAL HEAT INLET TRAN RPV LOWER HEAD SHELL RPV INLET NOZZLE INNER SURFACE RPV INLET NOZZLE OUTER SURFACE RPV OUTLET NOZZLE INNER SURF RPV OUTLET NOZZLE OUTER SURF CHARGING NOZZLE NOZZLE SAFETY INJECTION NOZZLE NOZZLE RESIDUAL HEAT REMOVAL TEE RPV NEAR CRDM PENETRATION FEEDWATER NOZZLE SAFE END RECIRC SYS - TEE SUCTION PIPE CORE SPRAY LINE SAFE END EXT RHR LINE STRAIGHT PIPE FEEDWATER LINE ELBOW RPV LOWER HEAD TO SHELL RPV FEEDWATER NOZZLE BORE RECIRC SYSTEM RHR RETURN LINE CORE SPRAY SYSTEM NOZZLE CORE SPRAY SYSTEM SAFE END RESIDUAL HEAT TAPERED FEEDWATER LINE - RCIC TEE MAT LAS LAS LAS 304/316 LAS 304/316 LAS 304/316 304/316 LAS LAS LAS 304/316 304/316 304/316 304/316 LAS LAS 304/316 304/316 LAS LAS LAS 316NG 316NG 304/316 LAS LAS LAS LAS LAS 304/316 304/316 304/316 LAS LAS 304/316 LAS LAS LAS LAS LAS 304/316 LAS 304/316 304/316 LAS DET DET USEAGE@40 USEAGE@60 1.40E-02 4.75E-01 4.72E-01 2.60E+00 1.04E-01 5.02E-01 4.57E-01 2.86E-01 4.87E-01 1.30E-02 1.72E-01 5.53E-01 6.61E-01 5.62E-01 3.17E-01 8.40E-02 2.23E-01 4.69E-01 1.05E+00 5.30E-01 1.80E-02 2.90E-01 6.58E-01 3.37E+00 1.46E+00 2.73E+00 8.91E-01 3.02E-01 4.96E-01 4.99E-01 3.47E-01 3.19E-01 3.27E-01 2.05E-01 6.28E-01 1.88E+00 8.30E-01 4.36E-01 1.13E+01 3.69E+00 7.90E-02 3.17E+00 3.90E+00 5.20E-01 1.77E+00 4.78E-01 6.98E+00 2.10E-02 7.12E-01 7.08E-01 3.90E+00 1.56E-01 7.53E-01 6.85E-01 4.29E-01 7.30E-01 1.95E-02 2.58E-01 8.29E-01 9.92E-01 8.43E-01 4.75E-01 1.26E-01 3.35E-01 7.04E-01 1.58E+00 7.95E-01 2.70E-02 4.35E-01 9.87E-01 5.06E+00 2.19E+00 4.10E+00 1.34E+00 4.53E-01 7.44E-01 7.48E-01 5.20E-01 4.79E-01 4.90E-01 3.08E-01 9.42E-01 2.82E+00 1.25E+00 6.54E-01 1.69E+01 5.53E+00 1.19E-01 4.75E+00 5.85E+00 7.80E-01 2.65E+00 7.17E-01 1.05E+01 CUMULATIVE PI@40 YR 7.89E-06 1.40E-02 4.22E-01 9.95E-01 9.56E-04 1.06E-02 1.01E-03 8.68E-03 1.13E-02 2.68E-06 1.88E-03 5.91E-01 9.39E-01 1.18E-02 7.56E-03 3.94E-02 8.25E-03 7.74E-01 1.30E-01 5.72E-02 3.21E-05 2.49E-03 8.62E-01 9.51E-01 4.34E-03 9.58E-01 1.11E-01 3.91E-01 6.81E-02 4.90E-01 1.63E-01 4.67E-04 1.88E-03 1.34E-02 7.89E-05 1.04E-01 4.23E-02 3.83E-04 4.73E-01 1.59E-01 2.71E-10 7.27E-02 9.43E-01 1.41E-04 3.33E-01 1.47E-03 3.86E-01 CUMULATIVE PI@60 YR 4.82E-05 4.44E-02 6.89E-01 9.99E-01 3.84E-03 6.75E-02 4.81E-03 3.16E-02 5.75E-02 1.93E-05 7.89E-03 8.46E-01 9.87E-01 5.31E-02 3.59E-02 1.19E-01 2.50E-02 8.99E-01 4.79E-01 2.08E-01 1.71E-04 1.05E-02 9.49E-01 9.83E-01 3.69E-02 9.99E-01 1.28E-01 6.44E-01 1.11E-01 7.53E-01 2.38E-01 3.75E-03 1.31E-02 5.16E-02 3.49E-04 2.53E-01 1.39E-01 1.27E-03 6.71E-01 3.65E-01 2.76E-08 2.42E-01 9.99E-01 7.89E-04 7.64E-01 7.89E-03 7.82E-01 CUMULATIVE PTWC@40 YR 6.71E-15 5.90E-05 1.74E-03 9.81E-01 2.61E-06 9.00E-05 1.00E-06 2.61E-06 2.00E-05 6.36E-16 4.11E-07 7.05E-02 6.27E-01 4.10E-05 1.40E-05 2.10E-04 7.85E-06 1.83E-01 2.10E-03 3.00E-03 7.52E-13 9.17E-07 3.65E-01 8.72E-01 5.00E-04 7.80E-01 7.20E-07 4.38E-03 4.48E-04 9.33E-03 7.77E-03 3.00E-07 4.00E-06 1.15E-04 7.88E-12 1.31E-03 4.80E-04 1.45E-07 4.10E-01 1.01E-03 0.00E+00 1.00E-05 7.12E-01 1.91E-08 1.46E-02 9.21E-05 2.99E-03 CUMULATIVE TWC/YEAR TWC/YEAR PTWC@60 YR @40 YR @60 YR 1.44E-12 9.01E-04 2.90E-02 9.98E-01 5.50E-05 1.03E-03 1.90E-05 5.50E-05 4.53E-04 1.85E-13 1.33E-05 3.53E-01 8.85E-01 5.98E-04 2.00E-04 2.36E-03 1.52E-04 5.44E-01 3.09E-02 2.54E-02 9.64E-11 2.84E-05 7.42E-01 9.63E-01 1.09E-02 9.80E-01 1.11E-05 5.04E-02 3.32E-03 9.60E-02 3.60E-02 5.20E-06 8.80E-05 1.14E-03 6.82E-10 1.47E-02 4.67E-03 3.25E-06 6.21E-01 1.46E-02 0.00E+00 8.80E-04 9.85E-01 8.84E-07 1.10E-01 1.02E-03 5.92E-02 1.13E-15 7.50E-06 3.58E-04 7.60E-02 3.46E-07 1.75E-05 3.75E-07 3.46E-07 7.00E-06 1.07E-16 5.87E-08 8.98E-03 4.36E-02 8.75E-06 2.25E-06 4.38E-05 1.04E-06 1.94E-02 5.88E-04 4.26E-04 1.23E-13 1.30E-07 3.17E-02 5.38E-02 5.33E-05 6.25E-02 8.38E-08 7.53E-04 4.75E-05 1.56E-03 6.99E-04 7.50E-08 8.75E-07 1.63E-05 1.25E-12 2.38E-04 7.13E-05 1.97E-08 1.35E-02 1.69E-04 0.00E+00 2.50E-06 7.20E-02 2.85E-09 2.08E-03 1.07E-05 6.96E-04 1.90E-13 7.59E-05 2.57E-03 9.38E-02 5.06E-06 1.15E-04 1.50E-06 5.06E-06 4.40E-05 1.85E-13 1.33E-05 2.27E-02 5.48E-02 5.05E-05 1.85E-05 1.98E-04 1.36E-05 3.35E-02 2.22E-03 1.79E-03 1.21E-11 2.83E-06 4.50E-02 5.66E-02 1.30E-03 1.18E-01 9.08E-07 3.96E-03 2.18E-04 7.54E-03 1.83E-03 6.00E-07 1.05E-05 9.26E-05 8.26E-11 1.23E-03 3.66E-04 3.04E-07 2.25E-02 1.35E-03 0.00E+00 9.76E-05 1.23E-01 9.51E-08 8.04E-03 7.82E-05 5.54E-03 CDF @40 YR 1.13E-16 2.03E-11 9.65E-10 2.17E-06 2.77E-12 1.40E-09 1.88E-12 1.73E-11 1.89E-10 1.08E-17 1.58E-13 2.42E-08 1.24E-06 7.00E-10 1.13E-10 1.18E-09 1.04E-07 5.25E-08 2.94E-08 1.15E-08 1.24E-14 3.51E-13 8.57E-08 4.31E-07 2.67E-10 1.69E-06 8.44E-09 2.03E-09 1.28E-10 4.21E-09 1.89E-09 6.00E-13 4.38E-12 8.13E-10 1.25E-13 3.57E-11 1.07E-11 7.09E-15 2.54E-11 3.04E-09 0.00E+00 3.75E-14 1.08E-08 6.41E-17 4.68E-10 1.61E-12 1.04E-10 CDF @60 YR 1.91E-14 2.05E-10 6.93E-09 2.67E-06 4.05E-11 9.21E-09 7.50E-12 2.53E-10 1.19E-09 1.86E-14 3.59E-11 6.13E-08 1.56E-06 4.04E-09 9.25E-10 5.35E-09 1.36E-06 9.03E-08 1.11E-07 4.82E-08 1.50E-12 7.64E-12 1.22E-07 4.53E-07 6.50E-09 3.17E-06 9.15E-08 1.07E-08 5.89E-10 2.04E-08 4.94E-09 4.80E-12 5.25E-11 4.63E-09 8.26E-12 1.84E-10 5.49E-11 1.09E-13 2.03E-10 5.06E-09 0.00E+00 1.46E-12 1.84E-08 2.14E-15 1.81E-09 1.17E-11 8.30E-10

27-24

Table 4 Comparison of Alternative pc-PRAISE Calculations with Oconee-2 Event


Measure of the Circumferential Extent of Cracking

Percent of deep cracks (a/t >80%) that are the result of 2 or more linking of cracks from adjacent sites (Q1)

Percent of deep cracks (a/t > 80%) that are longer than 40% of the circumference (Q4)

Percent of cracked welds that have cracking over more than 60% of the inner circumference (Q2)

Percent of the welds that have deep cracking (a/t > 80%) that extends the deep cracking over more than 60% of the circumference (Q3)

Oconee-2

100%

0% if cracking <3-mm is neglected 100% if crack < 3-mm is included 51.8% 0.2%

0%

Baseline Calculation Calculation with 80% thermal gradient stress and 20% uniform tension stress Calculation with 5 circumferential sites for crack initiation increased to 10 sites Calculation with length of initiated crack increased from 1% probability for 2 inch crack to 10% probability

78.0% 31.1%

73.4% 3.4%

52.6% 0.0% 59.3%

19.8%

24.4%

64.0%

78.9%

81.5%

59.0%

58.1%

Cycles for crack initiation 100% correlated from site-to-site (but no circumferential variation in cyclic stress)

100.0%

98.8%

84.2%

92.8%

Cycles for crack initiation 100% correlated from site-to-site (but a 20% circumferential variation in cyclic stress)

89.4%

92.8%

71.1%

77.4%

27-25

1000
N0.0001%

N0.1%

N1%

Stress Amplitude, ksi

100

N2%

N5%

N10%

N25%

10

N50%

N75%

N90%

N9999%

Fatigue of Carbon Steel in Water at 550F; Low Strain Rate; High Sulfur; High Oxygen

1 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06

Cycles for Crack Initiation

Figure 1 Example of Probabilistic S-N Curves for Low-Alloy Steel


1.E+00

Cumulative Probability of Through-Wall Crack at 40 Yrs

1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-12 0 1 2 3 4 5 6 7 8 9 10

Fatigue Usage Factor at 40 Years

Figure 2 Comparison of Calculated Usage Factors with Calculated Through-Wall Crack Probabilities 27-26

1.E+00 1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11 1.E-10 1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00
7

Crack Initiation and Crack Growth Statistically Independent pc-PRAISE

Cumulative Probability of TWC - Alternative Case

Cumulative Probability of TWC at 60-Years (Water Environment)

Cumulative Probability of TWC at 40-Years (Air Environment)

No Failures in 10 Trials

Cumulative Probability of Through-Wall Crack at 40-Years (Water Enviroment)

Figure 3 Comparison of Probabilities of Through-Wall Cracks for Air Versus Water Environment and for 40-Year Life and 60-Year Life
1.00

Cumulative Probability of TWC and TWC per Year

0.90 0.80 0.70 0.60 0.50 0.40 0.30 0.20 0.10 0.00 0

Cumulative Probability of Crack Initiation Cumulative Probability of Through-Wall Crack

New Vintage Cumbustion Engineering Plant Surge Line Elbow

Initiated Cracks per Year

Through-Wall Cracks per Year

10

20

30

40

50

60

70

80

Time, Years

Figure 4 Calculated Probabilities of Crack Initiation and Through-Wall Crack for the Surge-Line Elbow of the Newer Vintage CE Plant 27-27

1.00

Cumulative Probability of TWC and TWC per Year

0.90 0.80 0.70 0.60 0.50 0.40 0.30 0.20 0.10 0.00 0

Cumulative Probability of Crack Initiation

Cumulative Probability of Through-Wall Crack Membrane Stress < 45 ksi (Baseline Case) TWC All Uniform Tension TWC Membrane Stress < 15 ksi

TWC All Gradient Stress

Newer Vintage Cumbustion Engineering Plant Surge Line Elbow


10 20 30 40 50 60 70 80

Time, Years

Figure 5 Calculated Probabilities of Through-Wall Crack for the Surge-Line Elbow of the Newer Vintage CE Plant for Alternative Through-Wall Stress Distributions

27-28

Figure 6 Small Diameter Pipe with Cracking Caused Thermal Fatigue Stresses (LER No. 270/97-001)

27-29

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent

Cumulative Probability of TWC at 40-Years (Depth of Initiated Crack = 2-mm and 4-mm)

1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 Latin Hypercube Method 1.E-10 1.E-11 1.E-11

a0 = 4-mm

a0 = 2-mm

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (Depth of Initiated Crack = 3-mm)

Figure 7 Effects of Initial Crack Depth

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent

Cumulative Probability of TWC at 40-Years (Crack Growth With Constant 3:1 Aspect Ratio)

1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 Latin Hypercube Method 1.E-10 1.E-11 1.E-11

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (Crack Growth With Constant 10:1 Aspect Ratio)

Figure 8 Effect of Flaw Aspect Ratio

27-30

1.E+00 1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11 Latin Hypercube Method Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent

Cumulative Probability of TWC at 40-Years (All Components Assumed to be 8.0-Inch Thick)

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (All Components Assumed to be 1.0-Inch Thick)

Figure 9 Effect of Wall Thickness

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent Latin Hypercube Method 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11

Cumulative Probability of TWC at 40-Years

Linear Stress Gradient 1.0 to 0.5

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (Stress for Crack Growth = Stress of Crack Initiation)

Figure 10 Effect of Small Linear Stress Gradient

27-31

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent Latin Hypercube Method 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11

Cumulative Probability of TWC at 40-Years

Linear Stress Gradient 1.0 to 0.0

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (Stress for Crack Growth = Stress of Crack Initiation)

Figure 11 Effect of Larger Linear Stress Gradient

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent Latin Hypercube Method 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11

Cumulative Probability of TWC at 40-Years

Uniform Stress 50%

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (Stress for Crack Growth = Stress of Crack Initiation)

Figure 12 Effect of Reduced Stress for Growth of Fatigue Crack

27-32

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent Latin Hypercube Method 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08

Cumulative Probability of TWC at 40-Years

Nonlinear Stress Gradient


1.E-09 1.E-10 1.E-11 1.E-11

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (Stress for Crack Growth = Stress of Crack Initiation)

Figure 13 Effect of Nonlinear Stress Gradient

1.E+00

Cumulative Probability of TWC at 40-Years (Initiation Time and Crack Growth Rates Correlated)

1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11

Crack Initiation Probabilities by Method 2 (Monte Carlo Simulation)

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of TWC at 40-Years (Initiation Time and Crack Growth Rates Statistically Independent)

Figure 14 Effect of Correlation Between Crack Initiation and Crack Growth

27-33

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent

Cumulative Probability of TWC at 40-Years (All Components With High Oxygen = 0.10 PPM)

1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11 Latin Hypercube Method

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (All Components With Low Oxygen = 0.01 PPM)

Figure 15 Effect of Increased Oxygen Content

1.E+00

Cumulative Probability of TWC at 40-Years (All Components With Medium Strain Rate = 0.01 %/sec)

1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11

Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent

Latin Hypercube Method

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (All Components With Low Strain Rate = 0.001 %/sec)

Figure 16 Effect of Small Increase in Strain Rate

27-34

1.E+00

Cumulative Probability of TWC at 40-Years (All Components With High Strain Rate = 1.00 %/sec)

1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11

Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent

Latin Hypercube Method

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (All Components With Low Strain Rate = 0.001 %/sec)

Figure 17 Effect of Large Increase in Strain Rate

1.E+00 1.E-01 Crack Initiation Probabilities by Method 2 Crack Initiation and Crack Growth Statistically Independent

Cumulative Probability of TWC at 40-Years (All Components With High Sulfur = 0.015)

1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11 Latin Hypercube Method

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Through-Wall Crack at 40-Years (All Components With Low Sulfur = 0.0)

Figure 18 Effect of Reduced Suffer Content

27-35

1.E+00 1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11 Latin Hypercube Method Crack Initiation Probabilities by Method 2 Effects of EAC Moderated Sulfur = 0.007; EDOT = 0.01 %/sec; O2 = 10 PPB

Cumulative Probability of Crack Iniation 40-Years (EAC Effects Moderated or Removed)

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Crack Initiation at 40-Years (All Components Assumed to be of Low Alloy Steel)

Figure 19 Effect of Reduced EAC for Low Alloy Steel

1.E+00 1.E-01 1.E-02 1.E-03 1.E-04 1.E-05 1.E-06 1.E-07 1.E-08 1.E-09 1.E-10 1.E-11 1.E-11 Latin Hypercube Method Crack Initiation Probabilities by Method 2 Effects of EAC Moderated Sulfur = 0.007; EDOT = 0.01 %/sec; O2 = 10 PPB

Cumulative Probability of Crack Iniation 40-Years (EAC Effects Moderated or Removed)

1.E-10

1.E-09

1.E-08

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

Cumulative Probability of Crack Initiation at 40-Years (All Components Assumed to be Stainless Steel)

Figure 20 Effect of Reduced EAC for Stainless Steel

27-36

28
FATIGUE LIFE REDUCTION IN PWR WATER ENVIRONMENT FOR STAINLESS STEELS
K. Tsutsumi H. Kanasaki T. Umakoshi Mitsubishi Heavy Industries, Ltd., Japan T. Nakamura S. Urata H. Mizuta S. Nomoto The Kansai Electric Power Co., Inc., Japan

28-1

28-3

28-4

28-5

28-6

28-7

28-8

28-9

28-10

28-11

28-12

28-13

28-14

ENVIRONMENTAL FATIGUE II

29
AN APPROACH FOR EVALUATING THE EFFECTS OF REACTOR WATER ENVIRONMENTS ON FATIGUE LIFE
Robert E. Nickell Applied Science & Technology Poway, CA David A. Gerber Gary L. Stevens Structural Integrity Associates San Jose, CA

29-1

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life
By Robert E. Nickell Applied Science & Technology Poway, CA And David A. Gerber Gary L. Stevens Structural Integrity Associates San Jose, CA Presented At The INTERNATIONAL CONFERENCE ON FATIGUE OF REACTOR COMPONENTS July 31-August 2, 2000 Silverado Country Club & Conference Center Napa, California
I. Background. Generic Safety Issue No. 166 (GSI-166), later renumbered as Generic Safety Issue No. 190 (GSI-190), was identified by U. S. Nuclear Regulatory Commission (NRC) staff [1] because of concerns about the effects of reactor water environments on fatigue life during the period of extended operation. The GSI was closed in December 1999, based on a memorandum from NRC-RES to NRC-NRR [2]. During the time when the issue was actively being evaluated, both the NRC and the industry carried out numerous studies to determine the scope and magnitude of the effects. Some of these studies were generic, such as those conducted by EPRI and its contractors. Others, by necessity, were plant specific. For example, the first two license renewal applicants Baltimore Gas & Electric Company for the Calvert Cliffs Nuclear Power Plant (CCNPP) and Duke Energy for the Oconee Nuclear Station (ONS) were required to address the GSI in their applications on a plant-specific basis. Each of the applicants developed responses to the NRC staff questions on environmental fatigue without the benefit of information from the GSI closure. With the closure of the issue, industry efforts have now shifted to the management of potentially increased through-wall cracking frequency and associated potential increases in leakage from fatigue-sensitive component locations. Management of these potential environmental fatigue effects is a license renewal issue, based upon communications with

29-3

the NRC staff. For example, an informal handout by the NRC license renewal staff at an industry meeting held on November 17, 1999 [4] identified an additional requirement for license renewal applicants . . . to incorporate reactor water environmental effects into fatigue aging management program decision criteria. In other words, the determination of whether a particular component location is to be included in a license renewal program for managing the effects of fatigue, and the characteristics of that program (e.g., inservice inspection method, frequency, and acceptance criteria) should incorporate reactor water environmental effects for the 60-year period. Reference 2 noted that the 60-year fatigue design basis for advanced light-water-cooled reactors (ALWRs) was considered by the NRC staff to be adequate without consideration of reactor water environmental effects, because of the conservatism in ASME Boiler and Pressure Vessel Code, Section III, Subsection NB-3000 fatigue design procedures. However, Reference 2 also found that the potential for increased frequency of throughwall cracking and leakage from environmentally-assisted fatigue was such that aging effects management programs for license renewal would need to consider reactor water environmental acceleration of fatigue cracking. This paper addresses the remaining regulatory concerns on reactor water environmental effects for fatigue of metal components during the license renewal period based on three considerations. First, the nuclear industry has an incentive to minimize through-wall cracking and leakage for both economic and safety reasons during the license renewal period. Second, most nuclear plants have fatigue monitoring programs that can manage through-wall cracking and leakage, provided that appropriate interpretation of reactor water environmental effects is factored into those programs. Third, an industry consensus on that interpretation is useful from both a cost and a regulatory stability perspective. It is the opinion of the authors that the industry program described in this paper is responsive to all three of these concerns. II. Fatigue Monitoring Program Screening Criteria. Time Limited Aging Analysis. Fatigue design, when evaluated in terms of either explicitly defined or implied numbers and severity of design-basis transients or operational cycles, has been determined to be a Time Limited Aging Analysis (TLAA), even though no explicit time-limited assumptions are involved. The rules of the ASME Boiler and Pressure Vessel Code, Section III, Subsection NB, for Class 1 nuclear power plant components, represent a typical explicit fatigue design basis. The rules of ANSI B31.1, with fatigue strength reduction factors applied to piping stress allowables that are measured in terms of design-basis numbers of equivalent full-range thermal cycles, represent a typical implicit fatigue design basis. Part 54.21(c)(1) of the License Renewal Rule identifies the following three acceptable methods for resolving TLAAs: i. The fatigue design analyses can be shown to remain valid for the period of 60 years (e.g., the number and severity of explicitly defined design-basis transients for 60 years of operation are bounded by the number and severity of

29-4

explicitly defined design-basis transients for 40 years); ii. The fatigue design analyses have been projected to the end of the period of extended operation (e.g., the cumulative fatigue usage factor (CUF) can be shown to remain less than 1.0 for 60 years); or

iii. The effects of aging on the intended function(s) will be adequately managed for the period of extended operation. In the latter case, the license renewal applicant is required to demonstrate that the elements of a program to manage the effects of fatigue are in place or will be in place during the period of extended operation. The fatigue TLAA evaluation is shown on the right side of Figure 1 as a licensing basis assessment. Plants with fatigue monitoring programs are able to show, in almost all cases, that the number of design-basis transients assumed in the fatigue design basis envelopes the actual number of transient events experienced for either 40 or 60 years of operation. In such a case, a formal finding that 10 CFR 54.21(c)(1)(i) is satisfied can be made. However, license renewal applicants then must introduce reactor water environmental effects into a subsequent fatigue evaluation of potentially fatigue-sensitive component locations. This evaluation is shown on the left side of Figure 1, and has been labeled "Technical Evaluation" since it is outside the bounds of the plant licensing basis. This brings in one or both of the other two options for addressing fatigue either an analytical re-evaluation of the cumulative usage factor (CUF) for 60 years (10 CFR 54.21(c)(1)(ii)) or a demonstration that fatigue effects are adequately managed (10 CFR 54.21(c)(1)(iii)). Potential Approaches. Five possible approaches for incorporating reactor water environmental effects into the subsequent fatigue evaluations have been used to-date by utilities that have submitted license renewal applications to the NRC. Two of these approaches are defined as explicit, and three are defined as implicit. The two implicit approaches are exemplified by the Oconee and Hatch License Renewal Applications (LRAs). One of these implicit approaches proposes to track the number of design-basis transients during the period of extended operation, and to reduce the number of those transients or cycles that would trigger aging management program action by a factor. This factor would conservatively incorporate known reactor water environmental effects. The other implicit approach proposes to include component locations in the plant fatigue monitoring program based on a CUF that is reduced by a factor. Again, the factor would conservatively incorporate known reactor water environmental effects. A common feature of these two implicit approaches is that design-basis transient severity is used, as opposed to actual operating transient profiles. Design-basis transient severity has been shown to provide conservatism on the order of 10 to 20, and often more, in explicit fatigue design calculations [5]. The three explicit approaches are described in the following. The first of these explicit approaches deterministically recalculates CUFs for a number of fatigue-sensitive component locations, and incorporates reactor water environmental effects in either the calculated increment of fatigue usage or in the fatigue design allowable. This explicit method, referred to as the Fen multiplier approach, was developed by EPRI and the

29-5

General Electric Company (GE) [6], and has been endorsed by the Pressure Vessel Research Committees (PVRCs) Cyclic Life Environmental Effects (CLEE) Steering Committee. The method has been used by EPRI and its contractors to recalculate CUFs for fatigue-sensitive component locations in an early-vintage Combustion Engineering (CE) PWR [5], an early-vintage Westinghouse PWR [7], a late-vintage GE BWR [8], and an early-vintage GE BWR [9]. This approach allows for the selective application of reactor water environmental reductions to incremental CUF calculations, depending on transient values of dissolved oxygen, temperature, strain range, and strain rate. References 5, 7, and 8 used actual plant transient profiles (as opposed to design-basis transient severity), as measured by plant instrumentation, in order to take advantage of the transient definition conservatism. All of the EPRI studies took advantage of actual numbers of plant transients. The second explicit method involves a reduced set of fatigue design curves, such as those proposed by Argonne National Laboratory (ANL) [11, 12]. This method has been used by Idaho National Engineering Laboratory (INEL, now INEEL) to recalculate the CUFs for fatigue-sensitive component locations in early and late vintage CE PWRs, early and late vintage Westinghouse PWRs, early and late vintage GE BWRs, and B&W PWRs. The results of the INEEL calculations were published in NUREG/CR-6260 [13]. The INEEL calculations took advantage of the conservatism available in the numbers of actual plant transients, relative to the numbers of design-basis transients, but did not recalculate stress ranges based on actual plant transient profiles. Another explicit approach can be used. This approach probabilistically determines the fatigue-sensitive component locations to be included in the fatigue aging management program. For a plant that has incorporated risk-informed methodology to redefine its plant inservice inspection (ISI) program, this involves assessing the impact of reactor water environmental effects on the number of locations to be inspected, the type of examination method to be used, and the frequency of examination. This explicit approach is exemplified by the probabilistic calculations carried out by Pacific Northwest National Laboratory (PNNL) [3] in support of the NRC staff evaluation of GSI-190. All of the fatigue-sensitive component locations evaluated in NUREG/CR-6260 were re-evaluated, in probabilistic terms, in Reference 3. It should be pointed out that a variant on the probabilistic approach is possible; that is, the deterministic results from NUREG/CR-6260 and the EPRI generic studies can be reviewed and incorporated as expert opinion in the risk-informed inservice inspection (RI-ISI) process. The justifications for the two implicit approaches depend upon negotiating the degree of conservatism in the fatigue design cycle or fatigue CUF reduction factor with the NRC staff. Industry calculations using the F en (intermittently applied reactor water environmental effects) approach would argue for a knockdown factor of 2 to 2.5, which would set a CUF limit of 0.4 to 0.5 as the decision criterion limit for fatigue monitoring or inservice examination. This limit is approximately the same as the Examination Category B-J threshold for Class 1 piping weld examination in Subsection IWB of the ASME Code Section XI. If NRC contractor calculations (reduced S-N curves) are used, knockdown factors from 6 to 10 are possible, which would argue for a CUF decision threshold of 0.1 to 0.15.

29-6

In the following sections, an approach is outlined that uses all of the evaluations performed to-date, coupled with plant-specific fatigue evaluation and ISI program results, to systematically address the fatigue TLAA for license renewal and the incorporation of reactor water environmental effects in a manner consistent with that shown in Figure 1. III. NUREG/CR-6260 Results. As a part of the effort to close GSI-166 (later GSI-190) for operating nuclear power plants during the current 40-year licensing term, INEEL evaluated fatigue-sensitive component locations at plants designed by all four U. S. nuclear steam supply system (NSSS) vendors Westinghouse (PWR), CE (PWR), GE (BWR), and B&W (PWR). For the first three vendor designs, component locations from both early and late vintage designs were evaluated. Reference 13 provides the results of those evaluations. The early-vintage Westinghouse PWR calculations are chosen here for illustrating the available information. The fatigue-sensitive component locations chosen for evaluation in Reference 13 for the early-vintage Westinghouse PWR calculations were: (1) the reactor vessel shell and lower head; (2) the reactor vessel inlet and outlet nozzles; (3) the pressurizer surge line, including the pressurizer and hot leg nozzles; (4) the reactor coolant system piping charging system nozzle; (5) the reactor coolant system piping safety injection nozzle; and (6) the residual heat removal system Class 1 piping. For the latter three component locations, INEEL performed representative design basis fatigue calculations, since earlyvintage Westinghouse PWRs typically utilized an ANSI B31.1 design basis for most of the Class 1 piping. Reference 13 found that four of the six fatigue-sensitive component locations had cumulative usage factors (CUFs) less than 1.0 for both 40 and 60 years of operation, including the effects of the reactor water environment through a reduced set of fatigue design curves [11]. However, in two of these four cases (the charging nozzle and the safety injection nozzle), the evaluators needed to use NB-3200 (design by finite element analysis) methods, instead of NB-3600 piping analysis methods, in order to obtain more realistic stress distributions. Reference 13 states, in part: Whereas the NB-3200 and NB3600 results were comparable for the S alt computed for the nozzle-to-charging system junction region, the Salt was reduced by more than a factor of four in the nozzle body (considered to be a branch connection in the NB-3600 analysis) region using the NB-3200 finite element analysis. Similarly, the report states that the corresponding reduction in Salt for the safety injection nozzle body was more than a factor of 10. These findings were corroborated in Reference 9, where a recirculation system tee connection in a BWR plant was evaluated by both NB-3600 and NB-3200 analysis methods. The reduction in the CUF at the most fatigue-sensitive location (the inside corner of the tee) was approximately a factor of 10. Such reductions in CUF are typical of NB-3200 calculations, providing a clear demonstration of a portion of the conservatism inherent in standard ASME Code NB-3600 piping fatigue analyses. The remaining two fatigue-sensitive component locations from the Reference 13 evaluation of early-vintage Westinghouse PWRs had the following results. The design-

29-7

basis CUF for the inside surface of the reactor vessel lower head near the shell-to-head transition, where core support guides are welded to the interior of the shell, was determined to be 0.290. When the NUREG/CR-5999 interim fatigue curves were applied, the 40-year CUF increased to 0.891, with the 60-year CUF at 1.337. Since the major contributors to these CUFs were 200,000 hypothetical alternating cycles of frictional and vibration loads, and since these design-basis cycles are conservatively defined, this component location is not considered to represent an issue. The conservatism in the definition of the design-basis cycles more than compensates for the reactor water environmental effects. This is not the case for the other most fatigue-sensitive component location the inside surface of the hot leg to surge line nozzle safe end. The design-basis CUF, including thermal stratification loads, was determined to be 0.900. When the revised interim fatigue curves of NUREG/CR-5999 were used, with actual numbers of cycles instead of designbasis numbers of cycles, the 40-year CUF increased to 4.248. The 60-year CUF increased to 6.372. The INEEL calculations were also based on 30-second stresses, implying that thermal stratification loads approximate a thermal shock. Some reduction of the CUF values might be possible with more realistic transient definitions. These results from Reference 13 for early vintage Westinghouse plants are typical of the findings for other plant types. In general, the majority of fatigue-sensitive component locations were found to have 60-year CUFs less than 1.0. A relatively few component locations were found to have environmentally-adjusted CUFs for 60 years that exceeded 1.0, using the reduced S-N curve approach. In this sense, the Reference 13 results match up very well with the probabilistic results of Reference 3. IV. EPRI Generic Studies. The calculations reported in Reference 13 were based on the interim reduced fatigue design curves given in Reference 11. Such an approach penalizes the component location fatigue analysis unnecessarily, since research has shown that a combination of environmental conditions is required before reactor water environmental effects become pronounced. The strain rate must be sufficiently low and the strain range must be sufficiently high to cause continuing rupture of the passivation layer that protects the exposed surface area. Temperature, dissolved oxygen content, metal sulfur content, and water flow rate are additional variables to be considered. In order to take these parameters into consideration, EPRI and the General Electric Company jointly developed a method, called the Fen approach [6], that permits reactor water environmental effects to be applied intermittently, as justified by parameter combinations. The Fen approach was then applied by EPRI and its contractors to fatigue sensitive component locations in four types of nuclear power plants an early-vintage CE PWR [5], an early-vintage Westinghouse PWR [7], and both early-vintage [9] and late-vintage [8] GE BWRs. Component locations similar to those evaluated in Reference 13 were examined in these generic studies. A later study on reactor water environmental effects on Class 1 branch piping also used the F en approach [10]. The early-vintage Westinghouse PWR results from Reference 7 provide an excellent

29-8

example of the benefits of the F en approach. Actual transient information from plant instrumentation (e.g., hot leg temperature, pressurizer water temperature) over three cycles of operation (1994, 1995, and 1996) was extrapolated both backward and forward in time, in order to calculate the environmental factor to be applied to the design-basis CUF. The maximum value for any of the surge line locations pressurizer shell, pressurizer surge nozzle, pressurizer spray nozzle, pressurizer water temperature instrument nozzle, RCS hot leg surge nozzle, and charging nozzle was 1.91. This compares to an environmental multiplier of over 7 from Reference 13. The intermittent application of the environmental effects provides a reduction in CUF of almost 4. The effect of using actual transient information instead of design-basis transient definitions is even greater, approaching a factor of 10. These findings were confirmed in the other PWR study of an early-CE PWR [5]. Again, the pressurizer surge line was studied in detail. This calculation provided a direct comparison with the same component location evaluated in Reference 13. Section 5.2.3 of Reference 13 describes the results of calculating the fatigue CUF for the pressurizer surge elbow in older vintage CE plants. The component location is fabricated from austenitic stainless steel, with a design-basis CUF of 0.705 for 40 years of operation. Reference 13 cites a value for the 40-year CUF of 8.07 when the ANL interim environmental fatigue design curves [11] were used. This CUF value is more than ten times the design-basis CUF. Based on the F en approach, Reference 5 showed that the environmental multiplier was actually only about two. The difference between the two environmental multipliers is again about a factor of 4. Reference 13 also provided the results of CUF recalculation based on the removal of two different elements of conservatism: (1) the use of actual numbers of operating transients (as opposed to prescribed numbers of design-basis transients), and (2) the use of revised interim environmental fatigue design curves recommended by ANL (as opposed to the NUREG/CR-5999 [11] interim environmental fatigue design curves). Examining only the second of these elements of conservatism, the difference between the interim and the revised interim fatigue curves is shown in Figures 3-18 and 3-19 of Reference 13 and can be on the order of a factor of 2. In particular, the difference is especially significant for stress ranges up to 80 ksi. The alternating stress ranges for the two dominant load pairs are 44.89 and 36.99 ksi, respectively, so that the allowable number of design cycles are 4,878 cycles and 15,639 cycles, respectively, when using the NUREG/CR-5999 [10] curve. These number of allowable cycle values increased to 7,095 cycles and 35,238 cycles, respectively, when using the revised (0.001 %/sec strain rate) curve. Because the number of design-basis transients was 15,000 cycles and 72,025 cycles, respectively, for the two load pairs, the contributions to the CUF decrease to 2.114 and 1.987, respectively, with a total CUF of approximately 4.5. This recalculation shows that the excess conservatism in the NUREG/CR-5999 [11] stainless steel interim fatigue design curve is of the order of 1.5 to 2.2, and could have been higher. The ratio between the Reference 5 environmental multiplier and the revised Reference 13 environmental multiplier is now about 2, instead of 4. Collectively, the EPRI generic studies show that the difference between intermittent and continuous application of reactor water environmental effects is approximately a factor of
29-9

four. This ratio, which represents the difference between the industry F en method and the reduced S-N curve method, is one of the factors that can be used to implicitly or explicitly incorporate reactor water environmental effects into fatigue evaluations and aging management program decision criteria. The ratio of about 10 that represents the effect of actual thermal transients versus severe design-basis thermal transients is another factor. A factor representing the ratio of the number of actual thermal cycles experienced versus the number of design-basis thermal cycles is also available. V. EPRI Branch Piping Studies. Reference 10 evaluated six Class 1 small-bore branch piping systems at Oconee Units 1, 2, and 3, including reactor water environmental effects. The piping systems evaluated were: Core flood piping and the associated reactor pressure vessel nozzle Pressurizer spray piping and the spray nozzle in the pressurizer High pressure emergency injection (HPI) and normal makeup piping Decay heat removal piping Letdown piping, and Loop drain piping.

Because of the low amplitudes of cyclic stress, many of these piping systems were found to be exempt from Class 1 fatigue analysis. These exemptions included: Portions of the core flood piping outboard of the isolation valve Decay heat line piping, including the hot leg nozzle Letdown piping, including the cold leg nozzle Loop drain piping, including cold leg nozzles, and Cold leg pressurizer spray nozzles.

The three systems requiring fatigue evaluation were the pressurizer spray piping, the HPI/emergency piping, and the HPI/makeup piping. For the HPI/emergency and HPI/makeup piping locations evaluated, reactor water environmental effects led to CUFs on the order of 1.5 to 2.2. If a moderate environmental effects factor of either 1.5 or 2.0 is available for these locations, fatigue enhanced by reactor water environmental effects is not a concern. The pressurizer spray system for PWRs represents a potential thermal fatigue concern because of severe thermal transient conditions, including potential stratification. Component locations for this system were not analyzed for the older vintage W plant in NUREG/CR-6260 [13]. However, the pressurizer spray nozzle and two other pressurizer spray line locations were evaluated for thermal fatigue at the Oconee plant [10], including reactor water environmental effects and stratification transients. The procedure used for these environmental fatigue calculations in the Oconee study was similar to that used for the Calvert Cliffs pilot study [5], in that on-line monitoring of critical thermocouple locations was performed to determine actual thermal transient behavior. The evaluation

29-10

also included a virtual dissolved oxygen instrument based on Oconee plant experience, calculation of Fen multipliers for the thermal transients during the baseline measurement period, and extrapolation to 40 and 60 years of operation. All locations, except for the spray nozzle, were determined to have CUF values less than 1.0 for 60 years, including significant Fen multipliers that ranged from 3.9 to 7.4. Therefore, with the exception of the spray nozzle, branch piping locations do not present a major environmental fatigue concern. For the Oconee pressurizer spray nozzle, the original design-basis CUF of 0.60 bounded the CUF calculated for the inside and outside surfaces of the base metal (not exposed to the reactor water environment). For the stainless steel cladding, which is exposed to the reactor water environment, the design-basis CUF was 0.15. Environmental and thermal stratification effects were such that, for auxiliary spray cooldown, the design basis CUF of 0.15 increased to 0.973 (an Fen multiplier of 6.488) when the worst case thermal transients were used. Extrapolating the environmentally-enhanced CUF for the stainless steel cladding to 60 years would cause the CUF to exceed 1.0 (CUF = 1.46). If a moderate environmental effects factor of either 1.5 or 2.0 is available for consideration, even the spray nozzle would be of no concern for reactor water environmental effects. VI. NRC Staff Issues. The NRC staff raised a number of issues relative to the EPRI generic studies that have been the subject of a number of meetings over the past two years. Those issues can be separated into four topics: (1) The NRC staff has asked for revised F en calculations for carbon and/or lowalloy steel components analyzed in the EPRI generic studies, using an alternating strain range threshold of 0.07 % (as opposed to 0.10 % argued by the industry), above which the more recent ANL data would apply. (2) The NRC staff has asked that the moderate environmenta l effects factor be reduced to 3.0, rather than the factor of 4.0 argued by the industry. (3) The NRC staff has asked for revised F en calculations for austenitic stainless steel components analyzed in the EPRI generic studies, using an alternating strain range threshold of 0.097 % (as opposed to 0.10 % argued by the industry), above which the more recent ANL data would apply. (4) The NRC staff has asked that the moderate environmental effects factor be reduced to 1.5, rather than the factor of 2.0 argued by the industry. In addition, the NRC staff has asked that the industry explicitly address differences between the mean ASME air curve used as the basis for the ASME Code fatigue design curve and the mean air curve established by ANL. In essence, this implies a double environmental correction one correction to account for differences in fatigue life between ANL air data and ANL simulated reactor water data, and the other correction to account for differences between the ASME Code design basis mean air curve and the ANL mean air curve. VII (a). EPRI Generic Study Recalculations -- Carbon and Low-Alloy Steels. More
29-11

recent laboratory fatigue data in simulated LWR reactor water environments have been generated by ANL for carbon and low-alloy steels, and published in NUREG/CR-6583 [14]. These data do not differ substantially from the data used in the EPRI generic studies. However, the change in strain threshold may have a significant effect, and that effect has been evaluated. The recalculation is based on one of the examples contained in EPRI TR-105759 [6], a BWR carbon steel feedwater piping location with a design-basis fatigue usage factor of 0.1409 for 40 years. An alternating stress threshold of 30 ksi (approximating the alternating strain threshold of 0.10 %) was used initially to adjust the incremental fatigue usage for eight (8) out of thirty-one load pairs, giving an additional (environmental) fatigue usage of 0.0477, for a 40-year adjusted total of 0.1886. The overall F en multiplier in this case was 1.38 (1.68 for the eight affected load pairs). Reducing the alternating stress threshold to 21 ksi (approximating the alternating strain threshold of 0.07 %) would require an environmental adjustment for at least six additional load pairs. Assuming that the F en multiplier of 1.68 would continue to apply for the 14 load affected load pairs, the estimate for the adjusted fatigue usage factor would be 0.1409 0.0803 + 1.68 (0.0803) = 0.1955. The overall F en multiplier increases only to 1.39. Because the additional load pairs that would have to be included contribute relatively small increments to the total CUF, it is unlikely that changing the strain range threshold would present a significant issue. It should be pointed out also that the intermittent F en multiplier is less than a moderate environmental effects factor of either 3.0 or 4.0. VII (b). EPRI Generic Study Recalculations -- Austenitic Stainless Steels. More recent laboratory fatigue data in simulated LWR reactor water environments have also been generated by ANL for austenitic stainless steels, and published in NUREG/CR-5704 [15]. These data are substantially more penalizing than the data used in the EPRI generic studies. The updated analytical expressions for the mean fatigue initiation life of austenitic stainless steel in both air and the laboratory-simulated light water reactor (LWR) environment are given below. For the laboratory air environment, the number of cycle for fatigue crack initiation, N, is: ln (N) = 6.703 - 2.030 ln (a - 0.126) + T 1**, (1)

where a is the strain amplitude (%) and T 1* and * are transformed temperature and strain rate, respectively, defined as follows: T1* = 0 T1* = [(T-250)/525]0.84 T = temperature, C * = 0 * = ln(1/0.4) (T < 250C) (250 < T < 400C)

( > 0.4%/sec) (0.0004 < < 0.4%/sec)


29-12

* = ln(0.0004/0.4)

( < 0.0004%/sec)

For the laboratory simulation of the reactor water environment, ln (N) = 5.768 - 2.030 ln (a - 0.126) + T 2**O*, (2)

where T2* and O* are transformed temperature and dissolved oxygen (DO), respectively, defined as follows: T2* = 0 T2* = 1.0 O* = 0.260 O* = 0.172 (T < 200C) (T > 200C) (DO < 0.05 ppm) (DO > 0.05 ppm)

The analytical expression for Fen is then obtained as the following: Fen = exp [0.935 + *(T1* - T2*O*)] (3)

For the case of relatively low temperature (< 200oC), a low (bounding) strain rate, and either high or low dissolved oxygen, the environmental shift is 2.55. For relatively high temperature (> 200oC), low dissolved oxygen, and a low (bounding) strain rate, the environmental shift may be as high as 15.35, although there is a reduction above 250 oC where the environmental factor decreases to about 3.20 at 340 oC. For most of the component locations evaluated in the EPRI generic studies, these most recent data do not pose a problem for the demonstration that the 60-year CUF is less than 1.0, including reactor water environmental effects. Again, a significant benefit accrues to the Fen approach in this regard, since most of the penalizing thermal transients lie below the threshold temperature of 200 oC. Therefore, the environmental shift is relatively low, provided that a different multiplier is used for the portions of the transient that are above and below 200oC. However, for the most fatigue sensitive locations, (e.g., surge line elbow in PWRs), the environmentally-adjusted CUF increases over that calculated in the EPRI generic studies by a factor of about 2. If a moderate environmental effects factor of 2.0 is available, there are no component locations for which the CUF cannot be shown to be less than 1.0 for 60 years. If a moderate environmental effects factor of 1.5 is available, the environmentally-adjusted CUF for one or two locations exceeds 1.0 for 60 years. If no moderate environmental effects factor is available, the most fatigue-sensitive component locations will have 60-year CUFs that approach 2.0. A CUF greater than 1.0 does not guarantee fatigue cracking, due to other conservatisms in ASME Code Class 1 fatigue design analyses. However, there is some potential for fatigue cracking to occur when the most recent data on austenitic stainless steels are used in the environmental fatigue evaluations for one or two of the most fatigue sensitive component locations.
29-13

VIII. NUREG/CR-6674 Results. Reference 3 based the probabilistic risk assessments on the component location fatigue calculations of NURGE/CR-6260 [13]. Forty-seven (47) component locations in seven different types of light-water-cooled power reactors (e.g., new vintage Combustion Engineering PWRs, older vintage Combustion Engineering PWRs, Babcock & Wilcox PWRs, new vintage Westinghouse PWRs, older vintage Westinghouse PWRs, new vintage General Electric BWRs, and older vintage General Electric BWRs) were evaluated. The probability of through-wall cracking after either 40 years or 60 years of operation, using a very conservative approach, is fairly high for nine of those component locations: Surge line elbows in new vintage and older vintage CE plants, RPV outlet nozzles in older vintage CE plants, B&W plants, and new vintage W plants, Charging system nozzles in new vintage W plants, RHR inlet transitions in new vintage W plants, RHR line pipe straight pipe in new vintage GE plants, and Recirculation system RHR return lines in older vintage GE plants

However, for many of the other component locations, typical cumulative probabilities of through-wall cracking for either 40 years or 60 years of operation are on the order of 10 -5 to 10-8. Such probabilities do not warrant any extraordinary inservice inspection requirements. Even for a component location such as a charging nozzle safe end in a new vintage CE plant, with a cumulative 60-year probability of through-wall cracking of about 10-3, the efficacy of extraordinary inservice inspection may not be warranted. The 40-year probability is about 10-4. Therefore, based upon an interpretation of the very conservative through-wall cracking probabilities calculated in Reference 3, only a few component locations warrant a review and possible revision of the plant inservice inspection program to accelerate the inspection frequency. The reason for the relatively large sample of component locations with a cumulative probability of through-wall cracking can be traced to the very conservative assumptions used in the probabilistic calculations. While these assumptions are reasonable as the basis for regulatory bounding calculations, the assumptions are too conservative to serve as the basis for risk-informed inservice inspection (RI-ISI). One example should suffice. Detailed stress calculations in particular, the stress distribution across the component thickness were not available from NUREG/CR-6260. Only the peak cyclic stresses for surface locations, many of which represent discontinuities and stress concentrations, were available. Therefore, Reference 3 assumed that, for peak cyclic stresses with surface amplitude greater than 310 Mpa (45 ksi), the uniform (e.g., membrane) component was 310 Mpa (45 ksi), with the remainder of the stress amplitude assigned to a gradient category. The gradient category was defined as a nonlinear distribution of stress across the component thickness typical of step change thermal loading. This assumption is much

29-14

less conservative than assuming a uniform membrane stress across the component thickness equal to the peak cyclic stress. However, uniform thermal stress amplitudes are typically very small, in comparison to the linear component of stress and, in particular, the nonlinear component of stress. By reducing some of these conservatisms to more accurate best estimate levels, it would be possible to estimate the cumulative probability of through-wall cracking on a more realistic basis, thereby providing improved information to RI-ISI programs at the plants. Until such modified probabilistic calculations are completed, Reference 3 has identified a conservative set of component locations that can be used as input for RI-ISI consideration, based on a cumulative through-wall crack probability for 40 years of 10-3. IX. Section XI Program. The results from NUREG/CR-6260, NUREG/CR-6674, and from the EPRI generic studies can be used to demonstrate that, for most component locations, the TLAA resolution requirements of either 10CFR54.21(c)(1)(i) or 10CFR54.21(c)(1)(ii) are met. Only a few exceptions can be identified, depending upon the conservatism of the analytical methods used. Two general classes of component locations have been identified for which aging management program demonstrations would be required: Fatigue-Sensitive PWR Pressurizer Surge Line Locations Fatigue-Sensitive PWR Class 1 Branch Piping Locations

In addition, for some PWRs, the pressurizer spray line could be an exception because of thermal stratification loads. For these three sets of fatigue-sensitive component locations, the requirements of 10CFR54.21(c)(1)(iii) should be used. In other words, it is necessary to demonstrate that the effects of fatigue damage on the intended function(s) will be adequately managed for the period of extended operation. A sample of the surge line welds is examined as a part of the plant ISI program every ten years, in accordance with the requirements of the ASME Boiler and Pressure Code, Section XI, Subsection IWB, Examination Categories B-J and B-P. Surge line welds selected for the ISI program, by nature of their size, require a volumetric examination. A number of the welds should have been examined ultrasonically during the first two examination periods at PWR plants, and a somewhat larger sample of welds, and perhaps base metal, may need to be examined ultrasonically during subsequent ten-year periods. The justification for the enlarged sample could be based on RI-ISI methods. Most RI-ISI programs implemented to-date do not explicitly include reactor water environmental effects. Therefore, for the most fatigue-sensitive locations, the ten-year ISI interval may be called into question. This issue can be addressed by using the results of flaw tolerance evaluations, coupled with previous ultrasonic (UT) examinations, assuming they have not revealed any indications. In addition to the fatigue-sensitive pressurizer surge line locations, the other set of component locations that are deemed to be fatigue-sensitive, especially when reactor water environmental effects are considered, is fatigue-sensitive Class 1 small-bore piping. This set is defined as Class 1 piping and fittings with nominal pipe size ( NPS) less than

29-15

four inches in diameter. In particular, the concern is for those small-bore piping locations where either (1) a geometric discontinuity is present and the thermal loading is represented by thermal shock, or (2) stratified flow conditions are known to exist. In order to manage fatigue crack initiation and growth for this set of fatigue-sensitive small-bore piping locations, the plant may choose to develop a One-Time Inspection for ASME Section III Class 1 Piping Less Than 4 Inch License Renewal NPS Program. X. Summary. The objective of this paper is to examine existing fatigue evaluations and establish a process for evaluating the fatigue TLAA for both license renewal and reactor water environmental effects. Results from NUREG/CR-6260 have been coupled with results from industry generic and plant-specific studies. First, utilizing the results of fatigue monitoring programs, it can usually be shown that the number of design-basis transients assumed in the fatigue design basis envelopes the actual number of transient events experienced for either 40 or 60 years of operation. Thus, the TLAA on fatigue for license renewal can be shown to be acceptable based on 10 CFR 54.21(c)(1)(i) of the License Renewal Rule. Second, a technical assessment that introduces reactor water environmental effects into the fatigue evaluation can be performed based on the evaluations contained in plantspecific evaluations, NUREG/CR-6260, and the EPRI generic studies. The results of this technical assessment show that relatively few component locations are left that do not meet allowable acceptance criteria. The major exceptions are the PWR surge line elbows, welds, and nozzles. Another possible exception is Class 1 small-bore piping under significant thermal stratification loading. Thus, many components are considered acceptable through an analytical re-evaluation of the cumulative usage factor (CUF) for 60 years (10 CFR 54.21(c)(1)(ii) of the Rule). Finally, for those locations remaining that do not satisfy acceptance criteria when reactor water environmental effects are included, an approach that demonstrates that fatigue effects are adequately managed for the intended operating period is used (10 CFR 54.21(c)(1)(iii) of the Rule). This is accomplished by incorporating all appropriate fatigue-sensitive locations into relevant ISI programs. RI-ISI programs can consider reactor water environmental effects during the expert panel decision criteria step to further support this approach. Collectively, this process, which is outlined in Figure 1, can be used to address the TLAA on fatigue and also reactor water environmental effects and justify operation for an extended period. In order to address the remaining fatigue issues associated with license renewal, four additional generic tasks could be undertaken. The first task would be to complete the assembly of information and interpretation of data to support current license renewal applications. The second task would be to review the technical basis for the Reference 3 risk study, removing excess conservatism so that the results can be used more directly in support of ISI programs for detection and sizing of environmentally-assisted fatigue crack initiation and growth at fatigue-sensitive component locations. Third, for those fatiguesensitive component locations for which the probabilities of environmentally-assisted

29-16

fatigue crack initiation, through-wall crack growth, and subsequent leakage are sufficiently high such as surge line elbows in PWR plants, a cooperative lead-plant inspection program could be initiated. Since surge line elbows are fabricated from either wrought or cast austenitic stainless steel, the technical difficulties of ultrasonic examination to detect and size such cracks needs to be evaluated. Finally, the feasibility and cost of testing laboratory-scale and perhaps full-scale fatigue test articles under realistic reactor water environmental conditions should be assessed. This last task could be considered for potential interaction with a much larger effort called the Japanese Environmental Fatigue Testing Program. XI. References. 1. SECY-95-245, Completion of the Fatigue Action Plan , James M. Taylor, Executive Director for Operations, U. S. Nuclear Regulatory Commission, Washington, DC, September 25, 1995. Memorandum, Ashok C. Thadani, Director, Office of Nuclear Regulatory Research, to William D. Travers, Executive Director for Operations, Closeout of Generic Safety Issue 190, Fatigue Evaluation of Metal Components for 60 Year Plant Life, U. S. Nuclear Regulatory Commission, Washington, DC, December 26, 1999. NUREG/CR-6674 (PNNL-13227), Fatigue Analysis of Components for 60-Year Plant Life, Pacific Northwest National Laboratory for the U. S. Nuclear Regulatory Commission, Washington, DC, June 2000. Informal Handout at a Meeting Between U. S. Nuclear Regulatory Commission Staff and Industry, Christopher I. Grimes, Director, License Renewal Project Directorate, U. S. Nuclear Regulatory Commission, Washington, DC, November 17, 1999. Evaluation of Thermal Fatigue Effects on Systems on Systems Requiring Aging Management Review for License Renewal for the Calvert Cliffs Nuclear Power Plant, Report No. EPRI TR-107515, Structural Integrity Associates for EPRI, Palo Alto, CA, January 1998. An Environmental Factor Approach to Account for Reactor Water Effects in Light Water Reactor Pressure Vessel and Piping Fatigue Evaluations, EPRI TR105759, General Electric Company for EPRI, Palo Alto, CA, December 1995. Evaluation of Environmental Fatigue Effects for a Westinghouse Nuclear Power Plant, Report No. EPRI TR-110043, Structural Integrity Associates for EPRI, Palo Alto, CA, April 1998. Evaluation of Environmental Thermal Fatigue Effects on Selected Components in a Boiling Water Reactor Plant, Report No. EPRI TR-110356, Structural Integrity Associates for EPRI, Palo Alto, CA, April 1998. Environmental Fatigue Evaluations of Representative BWR Components, Report No. EPRI TR-107943, General Electric Company for EPRI, Palo Alto, CA, May 1998.

2.

3.

4.

5.

6.

7.

8.

9.

29-17

10.

Effect of Environment on Fatigue Usage for Piping and Nozzles at Oconee Units 1,2, and 3, Report No. EPRI TR-110120, Structural Integrity Associates for EPRI, Palo Alto, CA, December 1999. Interim Fatigue Design Curves for Carbon, Low-Alloy, and Austenitic Stainless Steels in LWR Environments, NUREG/CR-5999 (ANL-93/3), Argonne National Laboratory for the U. S. Nuclear Regulatory Commission, Washington, DC, April 1993. Statistical Analysis of Fatigue Strain-Life Data for Carbon and Low-Alloy Steels, NUREG/CR-6237 (ANL-94/21), Argonne National Laboratory for the U. S. Nuclear Regulatory Commission, Washington, DC, August 1994. Application of NUREG/CR-5999 Interim Fatigue Curves to Selected Nuclear Power Plant Components, NUREG/CR-6260 (INEL-95/0045), Idaho National Engineering Laboratory for the U. S. Nuclear Regulatory Commission, Washington, DC, March 1995. Effects of LWR Coolant Environments on Fatigue Design Curves of Carbon and Low-Alloy Steels, NUREG/CR-6583 (ANL-97/18), Argonne National Laboratory for the U. S. Nuclear Regulatory Commission, Washington, DC, March 1998. Effects of LWR Coolant Environments on Fatigue Design Curves of Austenitic Stainless Steels, NUREG/CR-5704 (ANL-98/31), Argonne National Laboratory for the U. S. Nuclear Regulatory Commission, Washington, DC, April 1999.

11.

12.

13.

14.

15.

29-18

Figure 1. Approach for License Renewal Fatigue Assessment

29-19

Robert E. Nickell David A. Gerber Gary L. Stevens

AN APPROACH TO EVALUATING THE EFFECTS OF REACTOR WATER ENVIRONMENTS ON FATIGUE LIFE

International Conference on Fatigue of Reactor Components


Silverado Country Club & Conference Center Napa, California

July 31- August 2, 2000


Robert E. Nickell, Applied Science & Technology David A. Gerber and Gary L. Stevens, Structural Integrity Associates

Reactor Water Environmental Effects on Fatigue Life

29-20

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Implicit Methods for Incorporating Reactor Water Environmental Effects x Reduce the number of design-basis transients or cycles that would initiate aging management program actions by a factor that conservatively incorporates known and applicable reactor water environmental effects. x Use the design-basis CUF, reduced by a factor that conservatively accounts for known and applicable reactor water environmental effects, as a decision criterion for initiating aging management program actions. x A common feature of both implicit approaches is that design-basis transient severity is used, as opposed to actual operating transient profiles. Design-basis transient severity has been shown to provide conservatism on the order of a 10 to 20, or even more, relative to actual operating transients. x Industry calculations support a factor of 2.0 to 2.5 (intermittent application of environmental effects); regulatory calculations support a factor of 6 to 10.
Reactor Water Environmental Effects on Fatigue Life
29-21

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Explicit Methods for Incorporating Reactor Water Environmental Effects x Deterministically recalculate the CUF for a number of fatigue-sensitive component locations, incorporating reactor water environmental effects in the recalculation through an environmental multiplier, Fen. x Deterministically recalculate the CUF for a number of fatigue-sensitive component locations, incorporating reactor water environmental effects in the recalculation through a reduced set of fatigue design curves. x Probabilistically evaluate the fatigue-sensitive component locations to be included in the aging management program, with reactor water environmental effects incorporated into the probabilistic calculations. x In all three cases, generic calculations are available to assist in the evaluation (NUREG/CR-6260, NUREG/CR-6674, EPRI generic studies). x Calculations feed into risk-informed ISI programs as expert information.

Reactor Water Environmental Effects on Fatigue Life

29-22

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

NUREG/CR-6260 Generic Results x Six or More Fatigue-Critical Component Locations Evaluated for Seven Reactor Types (Early and Late Vintage W PWRs; Early and Late Vintage GE BWRs; Early and Late Vintage CE PWRs; B&W PWRs) x For many component locations, direct application of the ANL fatigue curves produced CUFs greater than 1.0 for both 40 and 60 years x By reducing the number of cycles from the design basis to (extrapolated) actual cycles, or by making strain rate adjustments to the ANL fatigue curves, the CUF was reduced below 1.0 for both 40 and 60 years for most of the affected component locations x NB-3200 finite element calculations for NB-3600 piping locations were able to demonstrate substantial reductions in the design-basis CUF, even with reactor water environmental effects x For a few component locations (e.g., surge lines), further calculations (e.g., elastic-plastic analysis) were thought to be necessary to reduce the CUF below 1.0

Reactor Water Environmental Effects on Fatigue Life

29-23

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

NUREG/CR-6674 Generic Results x The PNNL risk study used appropriate bounding assumptions to estimate contributions to core damage frequency (CDF) for 47 of the most fatiguesensitive Class 1 component locations in seven types of LWRs x The extension of these bounding calculations showed only six component locations (e.g., PWR surge line elbows, PWR RPV outlet nozzles, BWR recirculation system RHR return lines) with sufficiently high through-wall cracking frequency to serve as the basis for riskinformed inspection x The extension of the CDF calculations to through-wall cracking frequency and leakage rates, using the bounding assumptions, may be too conservative to be used as the basis for risk-based management of potential reactor water environmental effects x Reduction of some of the conservatism will limit the number of locations to be inspected even further

Reactor Water Environmental Effects on Fatigue Life

29-24

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

NUREG/CR-6674 Generic Results (Continued) x 60-year cumulative through-wall cracking probabilities > 10-4

Surge-line elbows in newer vintage and older vintage CE plants RPV outlet nozzles in older vintage CE plants, B&W plants, and newer vintage W plants Charging system nozzles in newer vintage W plants RHR inlet transitions in newer vintage W plants RHR line piping in newer vintage GE plants Recirculation system RHR return lines in older vintage GE plants

x Most significant conservatism is probably the stress assumptions across the component wall thickness for crack growth calculations

Reactor Water Environmental Effects on Fatigue Life

29-25

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

EPRI Initial Generic Studies: Reactor Water Environmental Effects x On-Line Transient and Fatigue Usage Monitoring FatiguePro: On-Line Fatigue Usage Transient Monitoring System, Report No. EPRI NP-5835M, Electric Power Research Institute (1988)

Environmental module (e.g., virtual dissolved oxygen instrumentation) added to on-line monitoring system in 1997

x Reactor Water Environmental Effects Evaluation (Fen) Methodology An Environmental Factor Approach to Account for Reactor Water Effects in Light Water Reactor Pressure Vessel and Piping Fatigue Evaluations, Report No. EPRI TR-105759, Electric Power Research Institute (December 1995)

Reactor Water Environmental Effects on Fatigue Life

29-26

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

EPRI Generic Applications Studies: Reactor Water Environmental Effects x Evaluation of Thermal Fatigue Effects on Systems on Systems Requiring Aging Management Review for License Renewal for the Calvert Cliffs Nuclear Power Plant, Report No. EPRI TR-107515, Electric Power Research Institute (January 1998) x Evaluation of Environmental Thermal Fatigue Effects on Selected Components in a Boiling Water Reactor Plant, Report No. EPRI TR110356, Electric Power Research Institute (April 1998) x Evaluation of Environmental Fatigue Effects for a Westinghouse Nuclear Power Plant, Report No. EPRI TR-110043, Electric Power Research Institute (April 1998) x Environmental Fatigue Evaluations of Representative BWR Components, Report No. EPRI TR-107943, Electric Power Research Institute (May 1998)

Reactor Water Environmental Effects on Fatigue Life

29-27

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Generic Findings: Reactor Water Environmental Effects x Increase in CUF from Fen is between 1.0 to 1.6, and typically less than 3.0, for older vintage CE plants, older vintage Westinghouse plants, and both older and newer vintage BWR plants x Almost all modified CUFs fall within the moderate environmental effects margin in the ASME Code Section III fatigue design curves x The conservatism associated with design-basis thermal transient definitions versus actual thermal transient profiles is worth a factor of the order of 2 to 20 x The actual number of cycles versus numbers of design-basis cycles represents an added conservatism

Reactor Water Environmental Effects on Fatigue Life

29-28

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Major Generic Findings: Reactor Water Environmental Effects x The combination of actual thermal transients and actual numbers of thermal transients, plus the application of the Fen methodology and detailed finite element analysis, where necessary, provided a generic demonstration that CUFs are less than 1 for both 40 and 60 years for almost all fatigue-sensitive locations x Concern about austenitic stainless steel component locations during periods of operation at low temperature and low dissolved oxygen levels remains, and has a significant effect on EPRI generic findings x Field experience does not confirm the generic austenitic stainless steel findings, probably because of applicability of some of the laboratory simulations of reactor conditions, including flow rates and surface strain distributions

Reactor Water Environmental Effects on Fatigue Life

29-29

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

EPRI Generic Studies: NRC Staff Issues x Revised Fen calculations for carbon/low-alloy steels using an alternating strain threshold for reactor water environmental effects of 0.07 %, rather than 0.10 %, and NUREG/CR-6583 data (A recalculation from EPRI TR105759 shows that these changes would have little effect on the EPRI generic results). x Reduced moderate environmental effects factor of 3.0 (instead of 4.0) for carbon/low-alloy steels and 1.5 (instead of 2.0) for austenitic stainless steels (Recalculations show that these changes would have little effect on the EPRI generic results). x Revised Fen calculations for austenitic stainless steels using an alternating strain threshold for reactor water environmental effects of 0.097 %, rather than 0.10 %, and NUREG/CR-5704 data (Recalculations show that the NUREG/CR-5704 data have an effect on the order of a factor of 2, thus significantly affecting the EPRI generic results).

Reactor Water Environmental Effects on Fatigue Life

29-30

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

EPRI Generic Applications Studies: Reactor Water Environmental Effects x Effect of Environment on Fatigue Usage for Piping and Nozzles at Oconee Units 1, 2, and 3, Report No. EPRI TR-110120, Electric Power Research Institute (December 1999)

Exemption from Class 1 fatigue analysis for portions of the core flood piping outboard of the isolation valve; decay heat line piping, including the hot leg nozzle; letdown piping, including the cold leg nozzle; loop drain piping, including cold leg nozzles; and cold leg pressurizer spray nozzles. Three systems required fatigue evaluation: (1) pressurizer spray piping; (2) HPI/emergency piping; and (3) HPI/makeup piping. All locations, except for the pressurizer spray nozzle, were found to have CUFs (including reactor water environmental effects) < 1.0 for 60 years, even with environmental multipliers of 3.9 to 7.4. Pressurizer spray nozzle had 60- year CUF, including reactor water environmental effects, of 1.46 (Fen = 6.49).

Reactor Water Environmental Effects on Fatigue Life

29-31

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Application to a Westinghouse PWR x NUREG/CR-6260 evaluated (1) the reactor vessel and lower head; (2) the reactor vessel inlet and outlet nozzles; (3) the pressurizer surge line, including the pressurizer and hot leg nozzles; (4) the reactor coolant system piping charging system nozzle; (5) the reactor coolant system piping safety injection nozzle; and (6) the residual heat removal system Class 1 piping. x Four of the six locations had CUFs < 1.0 for both 40 and 60 years, including reduced fatigue design curves. Two of these four locations that were found to be < 1.0 needed NB-3200 calculations to reduce the CUF. x The remaining locations were: (1) the inside surface of the hot leg to surge line nozzle safe end (40-year CUF = 4.25; 60-year CUF = 6.37), and (2) inside surface of the vessel lower head near the shell-to-head transition, where core support guides are welded to the shell interior (40year CUF = 0.89; 60-year CUF = 1.34).

Reactor Water Environmental Effects on Fatigue Life

29-32

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Application to a Westinghouse PWR (Continued) x Application of the EPRI/GE Fen approach (intermittent reactor water environmental effects, as opposed to reduced fatigue design curves) reduces the CUF by a factor of between 2 and 4. x Application of laboratory data from NUREG/CR-5704 increases the CUF by a factor of about 2. x The number of mechanical loading transients for the core support guides is very conservative x Therefore, the inside surface of the lower head of the vessel is not a major fatigue issue. x The inside surface of the hot leg to surge line nozzle safe end must be included in the plant inservice inspection program, with an inservice inspection interval justified by flaw tolerance calculations.

Reactor Water Environmental Effects on Fatigue Life

29-33

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

SUMMARY x At least five (two implicit and three explicit) approaches are available for implementation by license renewal applicants as the basis for decision criteria to determine the fatigue-sensitive component locations for which fatigue crack initiation and growth, including the effects of reactor water environments, must be managed. x Generic information available from both NRC and EPRI contractor reports is able to show that almost all component locations have CUFs for 60 years, including reactor water environmental effects, that are less than 1.0. x The remaining component locations (e.g., the inside surface of the hot leg to surge line nozzle safe end in PWRs) must be included in the plant inservice inspection program, with an inservice inspection interval justified by flaw tolerance calculations, including appropriate reactor water environmental effects.

Reactor Water Environmental Effects on Fatigue Life

29-34

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

SPARES

Reactor Water Environmental Effects on Fatigue Life

29-35

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Presentation Outline x Recent Developments - NRC Staff -- 11/17/99: Risk Study Results, Grimes Statement -- 12/1/99: Craig Letter to ASME BNCS -- 12/3/99: NRC Staff Presentation to ACRS, Issue Closure -- 12/26/99: Memorandum From Thadani to Travers x Recent Developments - Industry -- 1/10/00: Industry Action Meeting at NEI -- 1/27/00: Industry Presentation to ASME BNCS -- 2/25/00: EPRI Letter to ASME BNCS -- 2/28-29/00: ASME Code Meetings, Daytona Beach

Reactor Water Environmental Effects on Fatigue Life

29-36

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Recent Developments x GSI-190 has been determined not to be a safety issue and is closed x Licensees continue to be asked by NRC staff to address the effects of reactor water environments on fatigue life as aging management programs are formulated in support of license renewal x The PNNL risk study shows no significant difference in calculated core damage frequencies (CDFs) between 40 and 60 years of life x The NRC staff has determined that ALWRs certified to 10 CFR 52 that are designed to ASME Code fatigue requirements for 60 years of life have sufficient design conservatism to compensate for reactor water environments x The industry argues that license renewal regulatory oversight is not required for environmental fatigue other than a TLAA demonstration that the fatigue CLB is applicable for 60 years of service

Reactor Water Environmental Effects on Fatigue Life

29-37

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Recent NRC Fatigue Activities x Extension of probabilistic risk study to 60 years of life has now been completed x Fatigue Analysis of Components for 60-Year Plant Life, F. A. Simonen, et al., has been published as NUREG/CR-6674, June 2000. x Results were discussed at the NRC Fatigue Workshop on November 17, 1999, which also included a distributed statement by Chris Grimes on the requirements for license renewal applicants. x Results were used as the basis for closure of GSI - 190 and discussed with the Advisory Committee on Reactor Safeguards (ACRS) on December 3, 1999 x ACRS issued favorable letter on the NRC staff findings

Reactor Water Environmental Effects on Fatigue Life

29-38

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Chris Grimes Statement, Fatigue Meeting, 11/17/99 x In particular, we anticipate that fatigue damage beyond the existing 40-year license term may not have sufficient impact on the core damage frequency to justify imposing a requirement to revise the current licensing basis to include an environmental factor in all of the fatigue design locations. x At least for the critical Class 1 locations, the staff would expect that an environmental factor would be accounted for in the process by which licensees decide when to take corrective action in their fatigue monitoring programs.

Reactor Water Environmental Effects on Fatigue Life

29-39

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Chris Grimes Statement, Fatigue Meeting, 11/17/99 (Continued) x We encourage the industry .. to establish an industry consensus on decision criteria to ensure that corrective actions will be taken before fatigue damage jeopardizes the ability of plant systems and components to perform their intended functions. x We would expect to continue to review the fatigue monitoring programs on a plant-specific basis, relative to some accounting for environmental effects in the decision criteria, until an industry consensus is established and endorsed by the NRC.

Reactor Water Environmental Effects on Fatigue Life

29-40

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Memorandum, Ashok C. Thadani, Director, Office of Nuclear Regulatory Research, to William D. Travers, Executive Director for Operations, dated December 26, 1999 x The advanced light water reactors (ALWRs) that have been certified under 10 CFR Part 52 were designed for a 60-year life expectancy. The associated fatigue analyses accounted for the design cycles based on a 60-year plant life but did not account for the environmental effects as addressed in GSI - 190. However, the staff has concluded that there is sufficient conservatism in the fatigue analyses performed for the generic 60-year ALWR life to account for environmental effects.

Reactor Water Environmental Effects on Fatigue Life

29-41

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Initial Industry Studies: Analytical Conservatism - Transient Definitions x Fatigue design analysis conservatism study Evaluation of Conservatisms and Environmental Effects in ASME Code, Section III, Class 1 Fatigue Analysis, Report No. SAND940187, Sandia National Laboratories (August 1994) x Findings Conservatisms in the transient definitions, analytical procedures, fatigue design curves, and other elements of the ASME Code explicit fatigue design process appear to more than compensate for reactor water environmental effects

Reactor Water Environmental Effects on Fatigue Life

29-42

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Recommendations, January 10, 2000, Industry Meeting (Continued) x The PNNL reactor water environmental effects risk study should be evaluated further by the industry -- Reduce conservative assumptions used in bounding analyses -- Perform sensitivity studies -- Single component Monte Carlo calculations for benchmarking x EPRI should continue its planning exercise to prepare for testing laboratory-scale fatigue test articles under realistic reactor water environmental conditions, while examining the potential for coordination with the Japanese EFT program

Reactor Water Environmental Effects on Fatigue Life

29-43

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

Recommendations, January 10, 2000, Industry Meeting (Continued) x As an alternative to a very expensive test program, EPRI should continue its evaluation of a cooperative industry program for inservice examination of the most fatigue-sensitive component locations (e.g., surge line elbows), considering base metal inspection capability, wrought versus cast stainless steel issues, and the potential for a lead plant program. x EPRI should prepare a short white paper in support of the SNOC Hatch LR proposal to use reactor water environmental fatigue effects management decision criteria -- i.e., fatigue design basis CUF threshold less than 1.0 (e.g., 0.1) for determining locations to be managed, with design-basis transient cycle monitoring and including non-CUF approaches as needed.

Reactor Water Environmental Effects on Fatigue Life

29-44

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

ASME Code Meetings, February 28-29, 2000, Daytona Beach, FL x ASME Section XI Task Group on Operating Plant Fatigue Assessment, Monday, February 28, 2000 -- Presentation, Stan Rosinski, EPRI, on January 10th Recommendations -- Presentation, Mike Davis, Duke Energy, on EPRI MRP Thermal Fatigue x ASME BNCS has directed Subcommittee III to take responsibility for the issue, with Dick Barnes (SG Design) in charge of oversight group; meeting of oversight group on Monday night to gather information; NRC staff (Joe Muscara) claim industry is stonewalling the issue, while NRC has worked on the issue for over ten years x Subgroup on Fatigue Strength (SG-FS) of the service Subcommittee on Design (SC-D) took several actions to implement reduced S-N curves

Reactor Water Environmental Effects on Fatigue Life

29-45

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

SUMMARY x GSI - 190 is closed; no actions relative to the ASME Section III Class 1 fatigue design basis could be justified by the NRC staff, for either 40 or 60 years of service, because of the small conditional contributions to core damage frequency x The bounding risk calculations show a measurable increase in throughwall cracking and potential leakage for a few of the most fatigue-sensitive component locations; the NRC LR staff feel justified in requesting that reactor water environmental effects be incorporated into aging management program decision criteria x This increase in through-wall cracking and leakage has not been observed in actual service, probably because the risk study was extremely conservative and because the laboratory data on reactor water environmental effects are not directly applicable

Reactor Water Environmental Effects on Fatigue Life

29-46

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

NRC Fatigue Activities x Generic Safety Issue (GSI) 78, Monitoring of Design Basis Transient Fatigue Limits for Reactor Coolant System, was identified in June 1983 -- determine whether transient monitoring (cycle counting) is necessary at operating plants x GSI-166, Adequacy of Fatigue Life of Metal Components, was identified in April 1993, as a consequence of license renewal fatigue issues raised for current operating plants x -- CUF > 1, reactor water environmental effects, Class 1 B31.1 piping, adequacy of inservice inspection x NRC Fatigue Action Plan, SECY-95-245, Completion of the Fatigue Action Plan, September 25, 1995

Reactor Water Environmental Effects on Fatigue Life

29-47

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

NRC Fatigue Activities x Memorandum, Eric S. Beckjord, Director, Office of Nuclear Regulatory Research, to Ashok C. Thadani, Associate Director for Inspection & technical Enforcement, Office of Nuclear Reactor Regulation, September 23, 1994 x Addressed Generic Issue No. 78, Monitoring of Fatigue Transient Limits for the Reactor Coolant System x The impact on core-damage frequency of fatigue failure in piping, using the environmentally-adjusted fatigue curves of NUREG/CR-5999, was found to be negligible in comparison to that from RPV failure. x Crack initiation does not insure through-wall cracking; the conditional probabilities of through-wall cracking are small (flaw tolerant piping). x Contribution to core-damage frequency is insensitive to CUF, even up to CUF = 10.

Reactor Water Environmental Effects on Fatigue Life

29-48

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

NRC Fatigue Activities x The probabilistic risk study results from Generic Issue 78 were later incorporated into the NRC Fatigue Action Plan, which was completed and documented in SECY-95-245. The staff concluded that the risk from fatigue failure of the primary coolant pressure boundary components was very small, based on a plant life of 40 years. The impact of a license renewal period of 20 years on fatigue of metal components was to be considered in the resolution of Generic Safety Issue 166. Later, Generic Safety Issue 190 was established to address this subject.

Reactor Water Environmental Effects on Fatigue Life

29-49

Robert E. Nickell David A. Gerber Gary L. Stevens

An Approach for Evaluating the Effects of Reactor Water Environments on Fatigue Life

NRC SECY-95-245 Findings

x the [NRC] staff believe that no immediate staff or licensee action is necessary to deal with the fatigue issues addressed by the [Fatigue Action Plan]; x fatigue failure of piping is not a significant contributor to core-melt frequency; x the [NRC] staff does not believe it can justify requiring a backfit of the environmental fatigue data to operating plants; and x the [NRC] staff believe that the [Fatigue Action Plan] issues should be evaluated for any proposed extended period of operation for license renewal.

Reactor Water Environmental Effects on Fatigue Life

29-50

30
DESIGN BASIS ENVIRONMENTAL FATIGUE EVALUATION AT OCONEE
Stan T. Rosinski EPRI 1300 Harris Boulevard Charlotte, NC 28262 Arthur F. Deardorff William F. Weitze Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557 Tim Brown Duke Energy Corporation 7800 Rochester Highway Seneca, SC 29672-0752

30-1

DESIGN BASIS ENVIRONMENTAL FATIGUE EVALUATION AT OCONEE

Stan T. Rosinski EPRI 1300 Harris Boulevard Charlotte, NC 28262

Arthur F. Deardorff William F. Weitze Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557

Tim Brown Duke Energy Corporation 7800 Rochester Highway Seneca, SC 29672-0752

ABSTRACT A Class 1 fatigue evaluation of piping systems was conducted at the Oconee nuclear plant. This analysis included an assessment of the effects of proposed environmental fatigue rules. The development of environmental correction factors (Fen) for application to individual load set pairs considered strain rates based on quasi-steady, thermal transient and dynamic loading effects. The approach was applied to several components and showed that environmental effects would be quite significant when using the proposed environmental correction to the fatigue curves. A critical evaluation of the applicability of the environmental correction factors is presented. INTRODUCTION Recent studies have determined that reactor water environment can affect component fatigue life. An evaluation methodology has been developed by EPRI where environmental correction factors (F en) can be applied to the partial usage factors determined in a conventional ASME Code fatigue analysis [1]. This proposed methodology has been used as part of an industry position regarding the management of environmental fatigue. This methodology was applied in several system-specific analyses to demonstrate use of the methodology and the general impact of reactor water effects on component fatigue [2-5]. During these studies, modified procedures were developed as new research results became available [6], including revised equations for carbon/low-alloy steel [7] and for stainless steel [8]. To assess the effects of environment on fatigue in representative pressurized reactor components, a Class 1 fatigue analysis was performed on the nonisolable stainless steel piping attached to the reactor coolant system at the Oconee Nuclear Station. The analysis included the effects of thermal stratification in some portions of the piping that had been identified during in-plant testing. The proposed environmental rules were then used to develop environmental factors that could be applied to the ASME Class 1 usage factor analysis. COMPONENTS EVALUATED ASME Code fatigue analysis was performed on Class 1 piping systems attached to the reactor coolant system out to and including the first isolation valve. Fatigue usage factors at a number of locations in the Core Flood, Pressurizer Spray and High Pressure Injection/Makeup piping systems were shown to require detailed Class 1 fatigue evaluation. The remaining piping systems were shown to be exempt from the requirements for a cyclic fatigue analysis and were not evaluated further since fatigue effects would be expected to be minimal. ENVIRONMENTAL CORRECTION FACTOR, Fen The environmental correction factor (Fen) is defined as the ratio between the fatigue life in air to that in water, both at the same temperature. It may be a function of several variables, including material,

30-3

water oxygen content, strain rate, and temperature. (For ferritic materials, the sulfur content is also a factor.) In performing an environmental fatigue evaluation, a separate Fen factor may be applied to each of the partial usage factors (associated with each load set pair) determined in a conventional ASME Code fatigue analysis. The equations for describing Fen for stainless steels [11] used in this evaluation are as follows: Fen = exp [0.935 + * (T1* - T 2* O*)] where: * * * (for > 0.4%/sec) (for 0.0004 0.4%/sec) (for < 0.0004 %/sec)

= = = =

0 ln (/0.4) ln (0.0004/0.4) strain rate, %/sec 0 [(T 250)/525]0.84 0 1.0 material temperature, C 0.260 0.172 dissolved oxygen, ppm

T1* = T1* = T2* = T2* = T = O* = O* = DO =

(for T < 250C) (for 250 T < 400C) (for T < 200C) (for T 200C)

(for DO < 0.05 ppm) (for DO 0.05 ppm)

A load set pair strain amplitude threshold was assumed (0.097 percent) below which there would be no effect of environment on fatigue usage [6]. ENVIRONMENTAL EVALUATION OF LOAD SET PAIRS As shown in the above equations, the environmental correction factor is a function of temperature, strain rate, and dissolved oxygen content. These factors are easily definable in laboratory testing. However, in real component fatigue analysis, it is not so apparent as to how to choose the appropriate variables to use in the analysis. The dissolved oxygen level may be different for the two events of a load set pair. Temperature and strain rate are highly variable during a thermal transient. The stress peaks and valleys of the transient pair may be far removed from one another (in time) and there may be other less significant stress change transients between the load set pair peaks and valleys. Thus, some assumptions must be made in determining how to apply laboratory data to actual component evaluation. Effective Environmental Factor (Fen) In actual reactor operating conditions, the conditions that exist for the peak and valley of a load set pair may be different and the strain rate and fluid and local surface temperature will vary during each of the contributing transients. Reference 7 presented an improved approach that determines an effective environmental factor as follows: Fen = 1 +

max

th

Fen 1 d max th

or

Fen =

max

th

Fen d max th

30-4

where: Fen max th

= = = =

instantaneous environmental factor based on current conditions during transient strain, relative to that at the most compressive stress state algebraic maximum strain for load set range pair strain threshold value

For a load set pair, the strain range, equivalent to max - min, may be determined as: max min = where: min = Salt E = = 2Salt E

algebraic minimum value of strain for load set pair, arbitrarily taken as zero for the integration process alternating stress from fatigue analysis, ksi modulus of elasticity from fatigue curve, ksi

This equation offers the opportunity to evaluate the time history of conditions that occur during a transient pair, rather than just assuming the worst possible value for the conditions that might exist, when determining environmental effects. Piping Analysis Equations The peak stress range and alternating stress amplitude for a load set pair for piping analysis is given by Equations 11 and 14 of ASME Section III, NB-3650 [9]: Sp = K1C1 D 0 P0 D + K 2 C 2 0 M i + K 3C3 E ab a Ta b Tb 2t 2I 1 1 + K 3 E T1 + E T2 2(1 ) 1

and

Salt =

K e Sp 2

The factor Ke is determined based on Equation 10, that describes the range of primary plus secondary stresses for the load set pair. The nomenclature is given in the ASME Code and will not be repeated here. As shown in the piping equations, the peak stress range may be a function of pressure changes, change in moments, and changes in thermal conditions that occur during the two conditions of a load set pair. When evaluating strain rate effects, strain rates associated with pressure and thermal expansion moment changes are expected to be relatively low and may or may not occur simultaneously with the thermal effects. The strain rates due to thermal effects may be high or low, depending on the transient; high strain rates are expected to occur simultaneously with high stress range rapid thermal transients. Dynamic loading (e.g., due to an earthquake, if present) would be expected to have a very high strain rate.

30-5

In a piping analysis, the stress range is generally calculated based on the extreme range of pressure and moment for the two events for the load set pair. For the local thermal effects, the maximum contribution of each of the terms may be added directly, or time phasing may be considered. The stress indices are defined such that the maximum multiplier is used for each individual term in the expression for stress intensity. Thus, the reported total stress range is a conservative assessment of the potential stress intensity excursion between two load states. This approach has been shown to be conservative for use in design of piping components, allowing the many locations in piping systems to be economically evaluated using modern computer techniques.

Transient Components for Strain Rate Analysis Figure 1 depicts the stress time history between two states of a load set pair. State 1 is chosen as that event that causes relatively compressive thermal stresses when the thermal terms are combined. State 2 creates a state of stress that is tensile relative to State 1. Dynamic (seismic) stresses can increase the overall strain range and can be assumed in analysis to occur with one or both of the thermal stress extremes, usually concurrently with the most severe load set pair. For evaluating an effective environmental factor (Fen), only the rising strain portion of the transient that exceeds the threshold strain (th) has been evaluated since the integration is from the positive strain threshold to a larger value of strain. This is reasonable since it is the tensile straining that tends to rupture the material surface layer that protects it from the water environment. Evaluation of Fen The thermal stress time history in piping analysis may be determined based on a combination of the three thermal terms from Equation 11 of ASME Section III, NB-3650. Thus, for the inside surface of piping, where the most significant thermal effects are experienced, K ET1 ET2 S th = K e 3 K 3C3 E ab ( a Ta b Tb ) 1 ( 2)(1 ) In this equation, each of the thermal terms is a function of time. For simplicity, it is assumed that the effect of the elastic-plastic multiplier Ke is uniformly distributed over the total strain range. In reality, the excess plasticity would occur only after exceeding 3Sm, but the analysis would be quite complicated to assume other than the uniform distribution. Since the piping analysis equations combine stress ranges using absolute values, the sign of the third term is chosen as that which produces the maximum stress range between two event pairs. This thermal stress time history is determined for each of the two load set states. The strain rate time history is determined from the derivative of the thermal stress time history: d 1 = dt E d Sth dt

The inside surface temperature, which can potentially be used to assess environmental factors, can be estimated from the time history of temperatures:

30-6

Tsurf =

Ta + Tb T1 + T2 2 2

When evaluating the complete stress range between two thermal transients, the maximum thermal stress near the end of State 1 may not be equal to the minimum thermal stress near the beginning of State 2. This difference of thermal stress must be considered, and can be considered to have a very slow strain rate since States 1 and 2 may not necessarily be closely related, or even occurring at the same time. For example, State 1 could represent heatup and State 2 could represent cooldown, with other less severe transients occurring between them. The stress range due to other load changes (e.g., pressure and thermal expansion moments) can be determined from: So = 2Salt Sth Sdyn where: Sdyn = stress range due to dynamic loads, including effects of Ke Sth = stress range due to thermal effects, including effects of K e The stress intensity range due to dynamic loads may be determined separately (e.g., by comparing the stress range for the load set pair with and without the seismic event). In evaluating each load set and load set pair, State A is defined as that with the most compressive stress, or with an upward thermal transient. State B is defined as the other event. In addition, each of the states may or may not include a significant thermal transient. Also, the load set pair may or may not include an earthquake or other dynamic event. In Reference 1, any load set pair with seismic content was considered to exclude environmental effects. This assumes that the seismic strain range would be relatively large. If the seismic contribution is small, there may still be environmental effects for the load set pair. Thus, for any pair where the dynamic loading is a contributor, the Fen will be evaluated using a high strain rate for the dynamic strain range contribution that will minimize environmental effects for the dynamic contribution. The seismic stress range will be considered to be with State A and/or State B per the original non-environmental fatigue analysis. Thus, for determining environmental factors there are five strain contributions that are evaluated. 1 2 3 4 5 Associated with a dynamic event at the relatively compressive state, existing only if a dynamic load is defined for State A. Associated with a thermal transient with State A, existing only if State A contains a thermal transient. Associated with "slow" events and equal to the total strain range, minus the dynamic and thermal ranges associated with both states A and B. Similar to 2 except that it occurs with State B transient conditions. Similar to 1, except that it occurs with State B.

30-7

In determining Fen, only the portion of the strains above the strain amplitude threshold are considered, consistent with the approach proposed in Reference 7. No evaluations were conducted that evaluated the complete strain range in the integration to determine Fen although this might be an alternate approach that could be considered. Evaluation of Oxygen and Temperature Effects During normal power operation in pressurized water reactors, coolant oxygen concentration is typically below the threshold of detection of about 0.002 ppm. Since the components analyzed are all stainless steel, low oxygen is controlling in that it yields a higher environmental factor. Since at least one of the load states was always at normal operating conditions, environmental factors were based on low oxygen (< 0.05 ppm). Assessment of temperature affects on environmental factors is more complex and not so well defined. The question is: What temperature should one use in the fatigue equations to assess the values of Fen? Several choices existed: 1) Use instantaneous surface temperature in the integration of Fen, 2) Use maximum Fen over the fluid temperature range, or 3) Use an average Fen over the fluid temperature range. It is not clear if it is the fluid or metal temperature that produces the environmental effect, or if it is the instantaneous metal temperature or the temperature that existed for a long time before the transient occurred. This is one of the difficulties in translating the laboratory test data to a realistic fatigue evaluation. Four approaches were evaluated (Table 1) to show the sensitivity of the temperature assumption on the analysis outcome. Different approaches were applied to the dynamic strain ranges as compared to the quasi-steady portion of the ranges associated with each load set pair. In determining the environmental factor for the quasi steady non-transient portions of the range, the approaches to determine Fen were: Cases 1 and 3: Fen was evaluated over the complete range of temperature associated with the two associated load states. The maximum Fen factor over the range was conservatively chosen. Cases 2 and 4: An alternate means of determining the temperature effects was evaluated as suggested in Reference 7: 1) If the maximum temperature was below the temperature threshold, then an average F en was determined over the actual temperature range, by choosing a number of temperature steps over the interval and determining an integrated average. For the cases examined in this evaluation, F en in this case was constant and equal to exp(0.935) = 2.547. 2) If the maximum temperature was above the temperature threshold and the minimum temperature was below the threshold, then an average Fen was determined over the range between the maximum and the threshold, using an integrated average as above. 3) If the minimum temperature was above the temperature threshold, then an integrated average Fen was determined over the actual temperature range. For determination of the environmental factors for thermal transients, the actual surface temperature during the transient was used for Cases 1 and 2. As a more conservative alternative, the fluid temperature in the range leading to the highest value of Fen considering both load sets were used for Cases 3 and 4.

30-8

For any dynamic portions of the strain range, the temperature for the associated quasi-steady or thermal transient condition was used. Adjustment of Fen The current ASME Code fatigue curves contain nominal factors of 2 on stress and 20 on cycles below the developed mean curve. A portion of the factor of 20 on cycles was to account for moderate environmental effects. Thus, after computing Fen, for any load set pair evaluated using the ASME Code fatigue design curves, the value was adjusted to account for this factor. Thus, the final adjusted effective environmental factor is determined by: F*en = Fen (but 1)

For carbon and low alloy steels, = 4[2]. For stainless steels, = 2.0 [2]. The NRC, in recent interaction, indicated that the factor for carbon/low-alloy steels should be 3.0 and for stainless steels the factor should be 1.5 [6]. In the analyses described in this report, = 1.5 was used for stainless steel. For Alloy 600 materials the Fen was previously defined as a constant value of 1.49. It is reasonable to assume that the same environmental factor inherent in the ASME Code fatigue design curves for carbon/low-alloy steels and stainless steels is also inherent in the fatigue design curve for Alloy 600. Assuming that the factor of 1.5 also applies to Alloy 600, there would be no net effect of environment for Alloy 600 since Fen is a constant (1.49). Thus, environmental effects of any Inconel 600 welds or components were not evaluated. RESULTS OF ENVIRONMENTAL EVALUATION The reactor coolant system attached piping considered in this study was analyzed to the Class 1 requirements of the 1983 ASME Code [9]. For performing the environmental evaluation, the piping analysis was not fine tuned to reduce any conservatisms prior to addressing environmental effects. The analysis was conducted using standard techniques, making reasonable assumptions to assure that the usage factors and other Code limits could be met without consideration of environmental effects. Some representative locations with relative high computed cumulative fatigue usage were chosen to assess the effects of environment as shown in Table 2. Table 3 lists the transients analyzed, along with the transient numbers used in the fatigue tables. In some cases, transients were grouped to reduce analytical effort where the transients were not especially severe. For spray transients, S1, S2, etc. in a transient number refer to a pressurizer spray event during that transient. For the transients, up indicates increasing temperature transients (causing compressive thermal stress) and down indicates decreasing temperature transients (causing tensile thermal stress). Some transients had both up and down portions. In addition, there were specific stratification transients identified for each system, identified with only a letter designator. The environmental assessment was based on the fatigue tables taken from the Code fatigue analysis output. Load sets with no transient, or with a slow transient, or with stress always decreasing during the transient were treated as quasi-steady state. For load sets with thermal transients, beginning and end times were chosen to capture only increasing thermal stresses, thus defining times of min and max.

30-9

Table 4 summarizes the results, showing all four cases. The difference in the approaches for choosing Fen as a function of temperature were minor for the various cases. Tables 5 through 9 give detailed fatigue usage calculations, both with and without environmental effects, for the most conservative evaluation approach (Case 3). Details of determining F en is provided in Reference 10. COMMENTARY ON PIPING ANALYSIS AND ENVIRONMENTAL EFFECTS ASME Class 1 piping analysis is conducted to the rules in ASME Section III, NB-3650. A set of equations is provided that conservatively shows that the rules of NB-3200 are met. There are many conservatisms in the design by rule approach using the piping equations. There have been no identified cases where the process is not conservative, except for cases where the loading conditions were not known at the time of the initial plant design. The proposed environmental rules have been developed using laboratory tests simulating a light water reactor environment. In order to complete testing in a reasonable amount of time, the testing must be conducted continuously. The resulting stress(or strain) time history bears little resemblance to that which occurs in an operating nuclear plant, or the stress range pairing required by the ASME Code. The proposed environmental rules provide another level of conservatism into the piping design equations, and considerably add to the complexity and labor needed to design and analyze piping systems. Consider the fatigue analysis in Table 6. The earthquake loadings have been conservatively considered to occur simultaneously with the most severe thermal transients. The first transient pairing is that of the termination of an HPI injection with the initiation of an HPI injection. Certainly, these are closely related transients, but for the environmental evaluation, the combination includes first the termination and then the injection, which is backwards relative to reality. The next load set pair is end of heatup paired with termination of HPI injection following a rapid depressurization. In this case, the range of pressures is certainly opposite of the thermal stresses, and the two transients are not mechanistically connected. Further down the table, there is a combination of stratification and no stratification. For the stratification case, the events could be closely connected such that the compressive stress followed by tensile stress sequencing could occur. For the HPI Makeup piping, the transient pairing could be mechanistic in that one of the controlling combinations is a loss of makeup followed by reinitiation of makeup. In this case, the tensile transient would closely follow the compressive transient. This type of transient is generally not the general rule in reactor systems, however. What is found in most fatigue analyses is that there is no relationship between the transients such that there would be long periods of time between the two transients of a load set pair for a material passivating layer to re-establish. This long period of re-passivating between the extreme of the strain cycles did not exist in the fundamental test data from which the environmental factors were determined. The conservatism in the design analysis environmental approach was also demonstrated in industry evaluations of components at another PWR where actual plant data were evaluated using fatigue monitoring [2]. For the fatigue monitoring environmental evaluation, actual plant transient sequence and magnitudes were considered. The resulting multipliers on fatigue usage to account for environmental effects (without consideration of the factor ) were 1.4 to 1.6 [2]. Thus, using environmental factors in a design basis analysis is extremely conservative. CONCLUSIONS Increases in cumulative usage factor for piping fatigue analyses when environmental effects are considered in a design analysis were shown to be very significant in all cases, and ranged from a factor of

30-10

2.8 to 8.8. Cumulative usage factors using an environmental correction exceeded the ASME Code limits for some locations. It is noted that the methodology for analyzing piping systems in the ASME Section III, NB-3650, contains a number of individual conservative assumptions which contribute to additional conservatism in an environmental effects analysis. For example, there is no consideration of cycle sequence where compressive transients might mechanistically follow tensile transients in event pairing. The analysis and the proposed environmental fatigue curves also do not account for long periods of steady state stress conditions between transients that might mitigate some environmental effects. The extent of the increases in cumulative usage factor leads to a prediction of piping failures due to fatigue. The fact that design basis transient fatigue failures are not common indicates that there is a large amount of conservatism in the combination of the original fatigue calculations and in the formulas used to calculate environmental effects. REFERENCES 1. An Environmental Factor Approach to Account for Reactor Water Effects in Light Water Reactor Pressure Vessel and Piping Fatigue Evaluations, EPRI, Palo Alto, CA: 1995, TR105759. 2. Evaluation of Thermal Fatigue Effects on Systems Requiring Aging Management Review for License Renewal for the Calvert Cliffs Nuclear Power Plant, EPRI, Palo Alto, CA: 1997, TR-107515. 3. Environmental Fatigue Evaluations of Representative BWR Components, EPRI, Palo Alto, CA: 1998, TR-1079434. 4. Evaluation of Environmental Fatigue Effects for a Westinghouse Nuclear Power Plant, EPRI, Palo Alto, CA: 1998. 5. Evaluation of Environmental Thermal Fatigue Effects on Selected Components in a BWR Plant, EPRI, Palo Alto, CA: 1998, TR-110356. 6. Letter, Douglas J. Walters (NEI) to Christopher Grimes (NRC), Request for Additional Information on the Industry's Evaluation of Fatigue Effects for License Renewal, April 8, 1999. 7. Chopra, O. K., and Shack, W. J., Overview of Fatigue Crack Initiation in Carbon and Low Alloy Steels in Light Water Reactor Environments, Journal of Pressure Vessel Technology, Volume 121, February 1999. 8. NUREG/CR-5704, Effects of LWR Coolant Environments on Fatigue Design Curves of Austenitic Stainless Steels, US NRC, April 1999. 9. ASME Boiler and Pressure Vessel Code, Section III, 1983 Edition with no Addenda. 10. Effect of Environment on Fatigue Usage for Piping and Nozzles at Oconee Units 1,2, and 3, EPRI, Palo Alto, CA: 1999. TR-110120. 11. Mehta, H.S., "An Update on the EPRI/GE Environmental Fatigue Evaluation Methodology and its Applications", PVP Vol. 356, Probabilistic and Environmental Aspects of Fracture and Fatigue, ASME 1999, pp. 183-193.

30-11

Table 1. Cases for Evaluating Fen Variation with Temperature


Case 1 2 3 4 Non-Transient Loads maximum Fen over temperature range average Fen over temperature range maximum Fen over temperature range average Fen over temperature range Thermal Transients surface temperature at each time point surface temperature at each time point maximum Fen over input temperature range maximum Fen over input temperature range

Table 2. Locations for Environmental Assessment


System Core Flood HPI/Emergency Unit 1-3 1 Location 5 175A 116 160 305 135C 130 210A Description Nozzle safe-end to pipe weld Nozzle safe-end to pipe weld Pipe to valve weld Pipe to valve weld Pipe to valve weld Pipe to aux. spray tee weld Pipe to valve weld 4" pipe to reducer weld CUF 0.033 0.193 0.177 0.640 0.291 0.052 0.042 0.081

HPI/Makeup 1 HPI/Emergency 2&3 HPI/Makeup 2&3 Pressurizer Spray 1

Pressurizer Spray 2&3

30-12

Table 3. List of Transients and Transient Groups


Number 1A 1ACF 1BCF 1B 2A/2B 3 4 5 6 7 8A 8B 8C 8HPI 9 10 11 12 14 16 15 17A 17B 19A 20D 22A 22C Groups M down and M up Group S down Group S up Group T down Group T up Group S1 down Group S2 up Group S3 up Group S4 up Transient Description Heatup, may include hydrotest (at pressure indicated) Core Flood Check Valve Test During Heatup Cooldown with Core Flood Decay Heat Removal Start Cooldown (zero load if so indicated) Power Increase/Decrease Power Loading Power Unloading Step Load Increase Step Load Decrease Step Load Reduction Reactor Trip, Loss of Flow Reactor Trip, High Temp. Reactor Trip, High Press. Manual HPI Activation after Trip Rapid Depressurization Change of RCS Flow Control Rod Withdrawal Hydrotest Control Rod Drop Steam Line Failure Loss of Station Power Loss of Feedwater Turbine Bypass Feed and Bleed Operations Loss of Makeup HPI System Test Pressurizer Heat Loss Evaluation Transients 2A/2B, 3, 4, 5, 6, and 10 for HPI and Core Flood analyses Transients 7, 8A, 8B, 8C, 14, 15, and 17A for HPI and Core Flood analyses Transients 7, 8A, 8B, 14, and 15 for HPI and Core Flood analyses Transients 11 and 17B for HPI and Core Flood analyses Transients 8C, 11, 17A, and 17B for HPI and Core Flood analyses Transients 3, 5, and 10 for Pressurizer Spray analyses Transients 2, 3, 4, and 14 for Pressurizer Spray analyses Transients 7 and 8B for Pressurizer Spray analyses Transients 11, 16, 17A, and 17B for Pressurizer Spray analyses

30-13

Table 4. Cumulative Fatigue Usage Factors with and without Environmental Effects
System Core Flood HPI/Emerge ncy HPI/Makeup HPI/Emerge ncy HPI/Makeup Pressurizer Spray Pressurizer Spray Location 5 175A 116 160 305 130 210A Original 0.033 0.193 0.177 0.640 0.291 0.042 0.081 Case 1 0.092 (2.75) 0.844 (4.36) 1.562 (8.83) 2.131 (3.33) 2.313 (7.95) 0.165 (3.94) 0.508 (6.27) Case 2 0.092 (2.75) 0.756 (3.92) 1.560 (8.82) 1.963 (3.07) 2.311 (7.94) 0.148 (3.55) 0.425 (5.25) Case 3 0.092 (2.75) 0.904 (4.69) 1.563 (8.84) 2.443 (3.82) 2.316 (7.96) 0.208 (4.97) 0.601 (7.43) Case 4 0.092 (2.74) 0.812 (4.21) 1.528 (8.64) 2.275 (3.56) 2.215 (7.61) 0.191 (4.56) 0.519 (6.41)

Note: Numbers in parentheses are overall increase factors relative to no environmental effects

Table 5. Fatigue Results - Core Flood Location 5 (Case 3)


State A Trans. 1ACF up+OBE Trans. 1ACF up+OBE Trans. 1ACF up Group T up Trans. 1A P=3192 Group S up State B Trans. 1BCF+OBE Trans. 1BCF Trans. 1BCF Trans. 1ACF down Trans. 1ACF down Trans. 1ACF down Ke 1 1 1 1 1 1 Salt (ksi) 67.05 63.58 60.12 46.86 45.61 44.21 Cycles 2 1 357 211 1 148 Nallowed 8412 10330 13647 47092 54293 64392 Total = CUFair 0.00024 0.00010 0.02616 0.00448 0.00002 0.00230 0.033 F*en 3.533 3.708 3.017 1.698 1.698 1.698 Total = CUFen 0.00085 0.00037 0.07892 0.00761 0.00003 0.00391 0.092

Table 6. Fatigue Results - HPI/Emergency Location 175A (Case 3)


State A 8HPI up+OBE 8HPI up+OBE Trans. 8HPI up Trans. 1A, P=3192 Trans. 1A, P=2567 Trans. 1A, P=2252 Trans. 1A, P=2252 Trans. 16 up Trans. 1A, P=2252 Trans. 9 up Trans. 9 up Trans. 22A up Trans. 1A, P=2252 No Stratification No Stratification State B 8HPI dn + OBE Trans. 8HPI dn. Trans. 8HPI dn. Trans. 9 dn. Trans. 9 dn. Trans. 9 dn. Trans. 16 dn. Trans. 22A dn. Trans. 22A dn. Trans. 12 Trans. 1B Trans. 1B Trans. 1B Stratification A Stratification B Ke 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 Salt (ksi) 129.25 126.66 124.28 118.73 117.30 116.57 102.04 100.81 97.26 60.81 58.41 48.04 46.88 27.5 24.01 Cycles 2 1 67 1 14 25 1 1 39 5 35 40 280 22302 14868 Nallowed 785 833 881 1008 1055 1080 1769 1851 2115 12894 15747 41638 46994 Total = CUFair F*en 0.00255 3.021 0.00120 3.048 0.07605 2.909 0.00099 3.267 0.01327 3.156 0.02315 3.098 0.00057 3.367 0.00054 3.150 0.01844 3.199 0.00039 10.232 0.00222 10.232 0.00096 10.358 0.00596 10.206 0.03769 10.232 0.00887 1 0.193 Total = CUFen 0.00770 0.00366 0.22122 0.00324 0.04188 0.07172 0.00192 0.00170 0.05900 0.00399 0.02272 0.00994 0.06083 0.38565 0.00887 0.904

30-14

Table 7. Fatigue Results - HPI/Makeup Location 116 (Case 3)


State A 1A, P=3192+OBE 1A, P=2567+OBE 1A, P=2567+OBE Trans. 1A Trans. 1A Trans. 1A Trans. 1A Trans. 1A Trans. 1A Trans. 1A Trans. 1A Trans. 16 up Trans. 1A Trans. 8HPI up Group T up Trans. 19Bup Trans. 1B up No Stratification No Stratification State B 20D up+OBE 20D up+OBE Trans. 20D up Trans. 20D up Trans. 20D dn. Trans. 20D dn. Trans. 16 down Trans. 9 dn. Trans. 8HPI dn. Trans. 22A dn. Trans. 9 up Trans. 1A Trans. 19A dn. Trans. 1B dn., Zero Trans. 1B dn., Zero Trans. 1B dn., Zero Trans. 19A (dn.) Stratification A Stratification B Ke 2.11 1.96 1.73 1.51 1.27 1.19 1 1 1 1 1 1 1 1 1 1 1 1 1 Salt (ksi) 229.11 209.91 179.54 150.99 133.47 124.67 62.01 61.37 59.08 56.74 56.01 55.15 54.45 34.89 29.49 28.62 26.59 52.37 21.37 Cycles 1 1 1 7 5 5 1 40 70 40 40 1 148 70 170 120 630 3186 2124 Nallowed 147 189 299 498 714 872 11750 12391 15061 18518 19782 21417 22862 265321 756245 911810 1284179 27895 3248937 Total = CUFair 0.00680 0.00529 0.00334 0.01406 0.00700 0.00573 0.00009 0.00323 0.00465 0.00216 0.00202 0.00005 0.00647 0.00026 0.00022 0.00013 0.00049 0.11421 0.00065 0.177 F*en 9.722 9.708 10.022 10.006 5.461 5.361 1.698 1.698 1.698 1.698 1.698 1.698 1.698 1.698 1.698 1.698 1 10.232 1 Total = CUFen 0.06611 0.05136 0.03347 0.14069 0.03823 0.03072 0.00015 0.00549 0.00790 0.00367 0.00343 0.00009 0.01099 0.00044 0.00037 0.00022 0.00049 1.16863 0.00065 1.563

Table 8. Fatigue Results - HPI/Emergency Location 160 (Case 3)


State A 8HPI dn. + OBE Trans. 8HPI up Trans. 8HPI up Trans. 8HPI dn. Trans. 8HPI dn. Trans. 8HPI dn. Trans. 9 up Trans. 1A, P=3192 Trans. 1A, P=2567 Trans. 1A, P=2252 Trans. 1B, P = 0 Trans. 1A, P = 2252 No Stratification No Stratification No Stratification No Stratification State B 8HPI up+OBE 8HPI dn. + OBE Trans. 9 down Trans. 8HPI up Trans. 16 up Trans. 9 up Trans. 16 down Trans. 22A down Trans. 22A down Trans. 22A down Trans. 22A up Trans. 1B, P=0 Stratification A Stratification B Stratification C Stratification D Ke 2.428 2.301 2.238 2.181 2.151 2.046 1.762 1 1 1 1 1 1 1 1 1 Salt (ksi) 241.1 224.4 212 209 202 182.9 141.8 94.2 92.7 92.0 52.1 42.8 >27.5 <27.5 23.42 24.15 Cycles 2 1 40 27 1 39 1 1 14 25 40 320 28674 8496 1062 6372 Nallowed 123.42 149.32 173.63 180.32 197.38 265.28 572.85 2232.57 2371.3 2439.77 27532 76928 Total = CUFair 0.016204 0.006697 0.230374 0.149731 0.005066 0.147012 0.001746 0.000448 0.005904 0.010247 0.001453 0.00416 0.05606 0.00444 0.00048 0.00330 0.640 F*en 3.123 3.062 3.236 3.088 3.095 3.188 3.580 2.544 2.368 2.284 10.232 10.194 10.232 1 1 1 Total = CUFen 0.05061 0.02050 0.74543 0.46236 0.01568 0.46874 0.00625 0.00114 0.01398 0.02340 0.01487 0.04241 0.57368 0.00444 0.00048 0.00330 2.443

30-15

Table 9. Fatigue Results - Pressurizer Spray Location 210A (Case 3)


State A Trans. 1BS1 up+OBE Trans. 1BS4 up Trans. 1BS2 up Stratification A Stratification A Stratification A Trans. 1B Trans. 1A Stratification A Stratification B State B Trans. 8AS1+OBE Trans. 8AS1+OBE Trans. 8AS1 Trans. 1BS2 up Trans. 1BS4 up Trans. 1AS1 Trans. 1BS4 down Trans. 1AS2 Group S1 down Group S1 down Ke 3.16 2.46 1.94 1.6 1.6 1.43 1 1 1 1 Salt (ksi) 186.472 126.223 88.1531 70.4849 70.4849 60.7184 26.7159 26.5965 22.6826 17.8494 Cycles 2 1 77 13 267 80 270 199 1492 1800 Nallowed 268 841 3046 6989 6989 12991 1432151 1475441 4248321 2.1E+07 Total = CUFair 0.00746 0.00119 0.02528 0.00186 0.03820 0.00616 0.00019 0.00013 0.00035 0.00009 0.081 F*en 6.553 7.392 7.778 5.918 7.244 9.474 1 1 1 1 Total = CUFen 0.04889 0.00880 0.19664 0.01101 0.27674 0.05836 0.00019 0.00013 0.00035 0.00009 0.601

30-16

Figure 1. Transient Pair Strain Range

30-17

DESIGN BASIS ENVIRONMENTAL FATIGUE EVALUATION AT OCONEE


Authors: Arthur F. Deardorff* and William F. Weitze Structural Integrity Associates Stan T. Rosinski EPRI Tim Brown Duke Energy Corporation International Conference on Fatigue of Reactor Components Napa, California USA July 31 - August 2, 2000 *Presenting Author adeardor@structint.com

30-18

PROJECT OBJECTIVES

Perform a Modern Piping Stress Analysis Save All Transient Time Histories Develop Approach for Application of Environmental Fatigue Rules Assess Environmental Effects Using Design Transients

30-19

SCOPE OF ANALYSIS

Performed ASME Section III (NB-3600) Class 1 Analysis of Attached Piping


core flood pressurizer spray high pressure injection/makeup decay heat removal letdown loop drains

Last 3 Exempt From Cyclic Evaluation (NB-3630(d)(2) Remaining Affected by Significant Thermal Transients

30-20

EQUATIONS FOR ENVIRONMENTAL EFFECTS


Developed Environmental Factor (Fen) for Each Load Set Pair From Current Fatigue Analysis Based On NUREG/CR-5704 Laboratory Tests for Stainless Steel
Fen = exp (0.935 + * (T1 * T2 * O * ))

where

* T1* T2* O*

= = = =

strain rate function temperature function 1 temperature function 2 dissolved oxygen function

30-21

Fen (for DO < 0.05 ppm)

Dissolved Oxygen > 0.05 ppm


9 8 7 6 18 16 14 12

Dissolved Oxygen < 0.05 ppm

Fen

4 3 2 1 0 0 100 200 300 400 500 600 700

Fen

10 8 6 4 2 0 0 100 200 300 400 500 600 700

Temperature, F

Temperature, F

30-22

ASSESSMENT OF REAL TRANSIENTS

Used Effective Environmental Factors


Fen =
max th

Fen d max th

where:

= strain, relative to most compressive state of load set pair max = maximum relative strain th = threshold relative strain max - min = 2Salt/E

Fen

Adjusted to Account for Code Factor


Fen * = Fen / 1.5 for SS

30-23

ASSESSMENT OF REAL TRANSIENTS

30-24

INFORMATION AVAILABLE FROM PIPING ANALYSIS


Equation 11 of NB-3650
Sp = K1C1 D0P0 D + K 2C2 0 Mi + K 3C3Eab a Ta bTb 2t 2I 1 1 K 3E T1 E T2 + + 2(1 ) 1

Equation 14 of NB-3650
Salt = K e Sp 2

Ke > 1.0 if Equation 10 of NB-3650 > 3Sm

30-25

DETERMINATION OF TRANSIENT PARAMETER TIME HISTORIES


Thermal Stress Approximation
K ET1 ET2 S th = K e 3 K 3C3 Eab ( a Ta bTb ) 1 (2)(1 )

Strain Rate
d 1 = dt E d Sth dt

Estimate of Material Inside Surface Temperature


Tsurf = Ta + Tb T1 + T2 2 2
30-26

DETERMINATION OF STRAIN RATES

ASME Code Analysis Requires that Usage Factors be Calculated Based on Extreme Stress States
tensile state may/may not follow compressive state may be years between two events

So, Must Synthesize Load Set Pair Transient for Strain Rate Analysis

30-27

DETERMINATION OF STRAIN RATES

Thus, can perform Fen integration with respect to strain (if we know temperature and oxygen content)
30-28

EVALUATION OF DO AND TEMPERATURE


Laboratory Testing Based on "Uniform Conditions" Conditions Different Between Real Load Set Pairs

What Do We Put in Our Equation for Fen for T and DO?


30-29

CHOICE OF DISSOLVED OXYGEN FOR Fen


For PWR's it is Generally Easy
most severe transients have DO << .05 ppm

But, What if Transient Pair Includes One DO > 0.05 ppm and Other with DO < 0.05 ??
may be important for PWR surge line evaluations since reactors may be just coming out of outages

30-30

CHOICE OF TEMPERATURE FOR Fen


Many Choices Exist for Choosing the T to Use in Fen Equation Integration
use temperature before each transient? use temperature instantaneously? use same or different for each state? use most conservative temperature? use material or surface or fluid temperature?

Fen Varies Considerably With Temperature

30-31

EVALUATION OF TEMPERATURE ASSUMPTIONS


Several Cases Evaluated
Case 1 2 3 4 Non-Transient Loads maximum Fen over temperature range average Fen over temperature range maximum Fen over temperature range average Fen over temperature range Thermal Transients surface temperature at each time point surface temperature at each time point maximum Fen over input temperature range maximum Fen over input temperature range

Differences as High of a Factor of 2

30-32

SUMMARY OF RESULTS
Core Flood: Effective Fen* 2.75
most of transients < 200C

HPI Emergency Inj.: Effective Fen* 3.0 to 4.7


controlling transients had high strain rates

HPI Makeup: Effective Fen* 7.6 to 8.8


major contribution due to stratification transients (from plant data)

PZR Spray Piping: Effective Fen* 3.5 - 7.5


included both rapid and stratification transients

30-33

COMMENTARY ON PIPING ENVIRONMENTAL EFFECTS


Piping Analysis is by Itself a Conservative "Design-byRule" Approach, Overstating Actual Fatigue Results P + M + T1 + T2 + Ta + Tb Ke cycle sequence not considered Laboratory Data is not Prototypical of Long-Term Plant Operation Current Environmental "Equations" Add Complexity
very difficult to deal with without adding considerable conservatism may not be realistic

30-34

31
AN UPDATED METHOD TO EVALUATE REACTOR WATER EFFECTS ON FATIGUE LIFE FOR CARBON AND LOW ALLOY STEELS

Makoto HIGUCHI Ishikawajima-Harima Heavy Industries Co., Ltd. Isogo-ku, Yokohama, 2358501 JAPAN

31-1

31-3

31-4

31-5

31-6

31-7

31-8

31-9

31-10

31-11

31-12

31-13

31-14

31-15

31-16

31-17

31-18

31-19

31-20

31-21

31-22

31-23

31-24

31-25

31-26

32
ENVIRONMENTAL FATIGUE, CRACK GROWTH RATES IN TITANIUM STABILIZED STAINLESS STEEL
Jussi Solin, Pivi Karjalainen-Roikonen VTT Manufacturing Technology

32-1

32-3

32-4

32-5

32-6

32-7

32-8

32-9

32-10

32-11

32-12

32-13

32-14

32-15

32-16

CODES AND STANDARDS

33
STRESS INTENSIFICATION FACTORS
E. A. Wais Wais and Associates, Inc. R. Carter EPRI

33-1

33-3

33-4

33-5

33-6

33-7

33-8

33-9

33-10

33-11

33-12

33-13

33-14

33-15

33-16

33-17

33-18

33-19

33-20

33-21

33-22

33-23

33-24

33-25

33-26

33-27

33-28

33-29

33-30

33-31

33-32

33-33

34
FATIGUE EVALUATIONS USING ASME SECTION XI NON-MANDATORY APPENDIX L
S. R. Gosselin, P.E. Pacific Northwest National Laboratory 902 Battelle Blvd. Richland, WA 99352

34-1

FATIGUE EVALUATIONS USING ASME SECTION XI NON-MANDATORY APPENDIX L

S. R. Gosselin, P.E. Pacific Northwest National Laboratory 902 Battelle Blvd. Richland, WA 99352

34-3

FAITIGUE EVALUSTAIONS USING ASME SECTION XI NON-MANDATORY APPENDIX L

S. R. Gosselin, P.E. Pacific Northwest National Laboratory 902 Battelle Blvd. Richland, WA 99352

ABSTRACT This paper describes an EPRI and USNRC collaborative project to evaluate the current ASME Section XI Appendix L fatigue evaluation procedures and propose changes to address several deficiencies. As components continue to operate beyond fatigue usage factors that exceeding 1.0, their failure frequencies are expected to increase unless augmented inspection strategies are implemented. A probabilistic approach to address NDE performance and inspection frequency issues related to damage tolerance assessments and component fatigue life extension will be described. Probabilistic fracture mechanics calculations are presented to demonstrate that an augmented level of inservice inspection can ensure that the failure probability of fatigue critical components will not increase if operation is continued beyond usage factors permitted by the ASME Section III design code. ASME SECTION III FATIGUE DESIGN The discussion in this section was drawn from a presentation by Dr. William E. Cooper, Teledyne Engineering Services to the PVRC Workshop on Environmental Effects on Fatigue Performance in January 1992 (1). The essential points of his paper are summarized below: The basic intent of the ASME Section III Code has not changed since before its inception - to establish rules that address design requirements for new construction while providing reasonable assurance of reliable operation. These requirements were directed mainly at the Manufacturer, while the Owner was assigned the important responsibility of defining the operational conditions (i.e. anticipated service duty), which the Manufacturer must consider. These service conditions are defined for the Manufacturer in the Design Specification. Prior to World War II, the design of pressure retaining components was primarily based on selecting a thickness that satisfied pressure stress criteria of one-fifth ultimate strength. During the war, this nominal factor of safety was reduced from five to four as an emergency measure. As a result of the apparent success of this change, the Codes were revised to adopt the lower factor

34-4

of safety and efforts were initiated to determine if further reductions were practical. The Special Committee to Review Code Stress Basis (SCRCSB) was tasked to determine what would be required to reduce the nominal factor of safety from four to three. The results of the SCRCSB work, Design by Analysis, were published in Section III in 1963 and Division 2 of Section VIII in 1968. The Design by Analysis procedure included a number of related considerations; however, the purpose for adding fatigue as one of the failure modes was to assure that the reduction of the nominal factor of safety from four to three would not result in a decrease in reliability if the vessel were subjected to cyclic operating conditions. The fatigue design procedures were intended to provide confidence that the component could be placed in service safely, not necessarily to provide a valid measure of actual component service life. The cyclic loading conditions defined in the Owner's Design Specification were not intended to represent a commitment on how the vessel was to be operated. Neither did it mean that the Owner was intended to be completely oblivious to the manner in which the vessel was operated, only that the design transient definitions should provide useful information. For example, if an Owner were able to show that the Design Specification included a cyclic event more severe than actual operating loads, then it could be assumed that the vessel would not be subjected to an unevaluated condition. ASME SECTION XI PHILOSOPHY This section presents a philosophy, originally endorsed by the ASME Section XI Executive Committee, intended to guide future Code activities regarding fatigue and its impact on component serviceability (2). Once a nuclear plant begins commercial operation, subsequent components integrity assessments are performed to determine the fitness for service. The results of the fitness-for-service evaluation are then be used to assess component operability and establish a "fatigue operating basis". ASME Section XI has made significant strides in developing acceptability criteria for major components based on improved understanding of material performance in actual operating environments; for example, irradiation damage in vessels and stress-corrosion cracking in BWR piping. The Section XI approach for protecting against fatigue damage will utilize a similar philosophy. The key elements include: 1. Evaluation procedures and criteria that ensure adequate protection from pressure boundary failures at the end of a prescribed operating period. 2. Evaluation procedures that combine the results of analyses, inspections, and service experience. A cumulative fatigue usage factor calculated in accordance with the rules in ASME Section III may not, in itself, represent a limit on component serviceability.

34-5

3. Alternative procedures designed to complement, not replace, the cumulative fatigue usage factor approach contained in ASME Section III. This would be accomplished by providing guidelines to address service conditions and loading mechanisms not anticipated in the Owner's Design Specification or accounted for in the original Design Report. 4. Procedures that include a fracture mechanics based flaw tolerance approach and consider both crack initiation and propagation. In November 1991, the ASME Section XI Task Group on Operating Plant Fatigue Assessment was formed to study fatigue evaluation methods, develop procedural guidelines, and establish acceptance criteria that operating plants could use to assess component serviceability. The results of this work were included in ASME Section XI Non-Mandatory Appendix L (3). Of significance was the introduction of a damage tolerance procedure to assess the serviceability of fatigue sensitive components. In this procedure the component is first inspected to verify the absence of any relevant indication or flaw. Subsequent inspections are then based on deterministic fracture mechanics calculations and the time required for an assumed flaw to grow to unacceptable size. ASME SECTION XI APPENDIX L The fatigue assessment approach in Appendix L is based on satisfying one of two principles: 1) The component is designed and operated such that fatigue damage will be limited to an amount less than that required for crack initiation, or 2) the component is designed and operated in a manner that is damage tolerant. In the latter principle, damage tolerant means that it must be able to tolerate fatigue accumulation and crack growth without reducing the structural integrity below acceptable limits. Figure 1 (4) shows how these principles can interact when evaluating plant components suspected to have a fatigue concern. In the majority of cases, operating plant fatigue concerns are associated with the discovery of loading conditions not previously accounted for in the Design Report (5). Faced with this problem, utilities will usually attempt to establish component operability by "requalifying" the component design. For those plants designed to ASME Section III, the analyst simply adds the new loading condition(s) to the other design transients assumed in the Design Report and demonstrates that the revised design basis cumulative usage factor (CUF) remains below 1.0. 1 Depending on the significance of the loading condition in question, requalifying the as-built component to the original design and licensing requirements may be extremely difficult and in some cases may not be possible. Examples of this are can be seen in the industrys efforts to
1

For the many operating plants designed to earlier piping design codes (i.e., ANSI B31.7, ANSI B31.1, etc.)

without original Class 1 fatigue analyses, this approach to component fatigue qualification becomes significantly more difficult and expensive.

34-6

address operating plant fatigue concerns in BWR and PWR feedwater nozzles, PWR surge lines, PWR spray lines, etc. Consequently, it may be necessary to deviate from the traditional design qualification analysis to confirm component serviceability. In these cases, a flaw tolerance approach is used. This analysis examines the tendency for flaw growth of a postulate fatigue crack assumed to be present in the component. The subsequent inspection schedule would then be based on the time required for this hypothetical flaw to grow to an unacceptable size. Appendix L Flaw Tolerance Procedures The flaw tolerance approach is well suited for operating plants because fatigue concerns are typically limited to a few specific plant component locations. Also, with this approach future performance capabilities can be established independent of past loading histories. The procedure for performing fatigue flaw tolerance evaluations is contained in Article L-3000 of Appendix L (3). The procedural steps include: 1. Verify the absence of any relevant indication exceeding the applicable flaw acceptance standards in Table IWB-3410. 2. Postulate the existence of a hypothetical flaw according to the flaw model in L-3200. 3. Determine the stresses at the location of the postulated flaw. 4. Determine the postulated end of evaluation period flaw size and critical flaw sizes by using existing ASME Section XI the analytical procedures contained in Appendix A, Appendix C, or Appendix H as applicable. 5. Apply the appropriate flaw acceptance criteria contained in IWB-3600. 6. Continue operation and inspect the component at a frequency specified in L-3420. Appendix L Flaw Model The basis for the Appendix L flaw model is discussed in EPRI TR-104691, Operating Nuclear Power Plant Fatigue Assessments (6). Since the assumed absence of any actual flaws is based on inspection results, the flaw model (size and shape) must be consistent with the inspection capabilities. In other words, the hypothetical flaw assumed in the flaw tolerance assessment should represent the largest flaw expected to be missed during the initial inspection. The flaw model used in Appendix L is described in Table 1 (ASME Section XI 1995). The postulated flaw sizes were based primarily on judgment and limited quantitative information of ultrasonic inspection (UT) capabilities. At that time, limited performance demonstration data were available for thick walled reactor pressure vessels and were considered most applicable when the component section thickness was greater than 4 inches. Since the industrys Appendix VIII performance demonstration initiative (PDI) for piping was just starting, similar data was not available for thinner components. This was especially true for components whose section 34-7

thickness was below 1 inch. Consequently, conservative estimates were made regarding flaw detection capabilities for these components. The Appendix L flaw model was subsequently reviewed by the ASME Section XI Working Group on Procedure Qualification and Volumetric Inspection. In their opinion, the inner surface fatigue cracks described in Table 1 could be detected in service for most materials that comprise the reactor coolant system. However, the detection capability may not be appropriate for components containing dissimilar (bimetallic or trimetallic welds) and statically or centrifugally cast austenitic piping components, due to their challenging metallurgical characteristics (6). Table 1 Appendix L Flaw Model Section Thickness, t (in.) 0.5 1 2 3 4-12
Reference Flaw Depth, a/t (%) 30.0

20.0 15.0 11.7


10.0

Notes: 1. 6:1 minimum aspect ratio 2. For t<0.5 inches, the postulated flaw depth, a = 0.15 inches 3. For t>12 inches, the postulated flaw depth, a = 1.2 inches

Appendix L Augmented Inspections In Appendix L, flaw tolerance is measured in terms of the allowable operating period, P. As defined in the Appendix, P represents the time required for the hypothetical flaw to grow to a depth equal to the maximum allowable flaw size specified in ASME Section XI, IWB-3600. Successive inspections are then implemented at the location of concern according to the schedule shown in Table 2 (3). The time intervals between inspections are designed to ensure that the component will be re-examined before a growing hypothetical flaw that exceeds half the size allowed by ASME Section XI. Table 2 Appendix L Successive Inspection Schedule Allowable Operating Period, P > 20 years Successive Inspection Frequency End of each 10 year ISI Inspection Interval End of P/2 operating years or each operating cycle, whichever is greater

< 20 years

34-8

Appendix L Applications Most recent applications of Appendix L have been associated with license renewal applications. The license renewal rule requires that the licensee perpetuate the current design basis to the end of the renewal period (i.e., 60 years). With regard to fatigue, this means that the calculated CUF using ASME methods must be maintained less than 1.0. Emphasis is placed on the more fatigue sensitive areas in the plant that are selected based on anticipated loading conditions as well as industry service experience. These fatigue sensitive locations are not necessarily limited to Class 1 components and may include Class 2 or ANSI B31.1 piping components whose original design qualification was not based on a CUF analysis. These industry applications have raised concerns regarding the effectiveness of the current flaw tolerance procedures. In a report to the ASME Section XI Task Group on Operating Plant Fatigue Assessment (7), BG&E indicated that, in license renewal space, the potential for largescale replacements is real. Lines such as the surge line, main spray, auxiliary spray, and CVCS have the potential to exceed a CUF of 1.0 in a multitude of locations prior to the end of the extended design life. In many cases the Appendix L flaw tolerance procedures do not provide a sufficiently long inspection interval to allow long-term plant operation.2 As such, the only viable options to deal with this would be to recalculate the CUF using more refined inputs, or at worst, repair/replace the affected items. The general consensus within ASME Section XI Working Group Operating Plant Criteria and Task Group Operating plant Fatigue Assessments is that a stronger technical basis is needed for both the flaw growth models and the rules used to establish inspection frequencies. Most noteworthy are the following issues: 1. Limited Effectiveness of Appendix L - Insights into flaw detection capabilities gained from industry performance demonstration initiatives (PDI) suggest that, in some cases, the NDE capabilities of field teams performing piping inspections are better than assumed in Appendix L. In these cases, the flaw size assumptions in Appendix L may be overly conservative which render the flaw tolerance alternative ineffectual. This is especially true for components whose section thickness is below 1 inch. 2. Flaw Model Applicability - The Appendix L flaw tolerance procedure flaw size assumptions are common to both carbon steel and wrought austenitic stainless steel materials. The assumed flaw depths are presented as a function of section thickness only. No distinction is made with regard to differences in flaw detection and sizing capability for carbon and stainless

At BG&E, augmented inspection frequencies were found to be unacceptably short, less than one operating cycle,

when environmentally assisted fatigue crack growth curves were applied indiscriminately throughout the evaluation period for stainless steel locations in the CVCS and pressurizer surge line (8).

34-9

steel materials or other NDE performance factors such as inspection access, inspection procedure, or inspector qualification. 3. Technical Basis An improved technical basis is needed for both the flaw growth models and the rules used to establish inspection frequencies. The flaw model (i.e., size and shape) used in Appendix L should be closely tied to performance demonstration results and the inspection frequencies need to ensure that the likelihood of failure will not increase beyond that provided in the original design code of record. 4. Probabilistic Fracture Mechanics Approach - There is a need for probabilistic fracture mechanics evaluation methods and acceptance criteria that can be used to evaluate component serviceability and define augmented inspection frequencies. PROBABILISTIC ASSESSMENT OF APPENDIX L INSPECTION STRATIGIES In this section we show how a probabilistic approach can be applied to determine inspection frequencies that account for demonstrated NDE performance and to ensure that reliable piping performance is maintained throughout the components original or extended operating life (9). In the example described below, we assume that the weld location is subject to thermal fatigue. The example considers a stainless steel pipe (29 inch outside diameter by 2.5 inch wall) which is loaded at 5000 cycles per year such to give a CUF=1.0 after 20 years of operation given a weld root stress concentration factor of 3.0. This corresponds to a nominal alternating stress of 27.3 ksi and a peak alternating stress, at the weld root, of 81.9 ksi. The inspection frequency necessary to maintain the components failure probability at or below that associated with the fatigue limit specified in the ASME Section III design code (e.g., cumulative usage factor (CUF) must be less than 1.0) is then determined. The inspection frequencies are compared to those obtained from the deterministic procedures in Appendix L. Probabilistic Calculations Probabilistic fracture mechanics calculations are presented to demonstrate that an augmented level of inservice inspection can ensure that failure rates of fatigue critical components should not significantly increase as operation is continued beyond usage factors permitted by the design code. Uncertainties in flaw growth rates and in flaw detection were addressed by application of the probabilistic fracture mechanics code pc-PRAISE (10). Suitable inspection frequencies were established for a given flaw detection capability (probability of detection or POD curve) by adopting a goal for an acceptable piping failure probability (i.e. probability of through-wall crack per weld per year). Continued operation for calculated CUFs exceeding unity was considered to be acceptable only if additional inspections are performed. The additional inspections are required to maintain calculated failure rates at levels less than or equal to calculated failure rates before the usage factors became 1.0.

34-10

The pc-PRAISE model assumed semi-elliptical surface flaws with aspect ratios of 12 and 20, and a Paris law for fatigue crack growth having a mean rate corresponding to constants of C = 9.14E12 and m = 4. A simplified treatment of flaw initiation was assumed. At time = 0.0, very small inner surface cracks were assumed to be present, with depths uniformly distributed between 0.005 to 0.010 inch. The alternative inspection frequencies were limited to the case of no inspections and inspections every 2 or 4 years, with the inspection program being introduced after 20 years of operation. The reliability for the ultrasonic NDE was described by the error function type curves used by the pcPRAISE code to describe flaw detection. Two bounding curves were assumed for purposes of the demonstration calculations. The less effective NDE assumed a threshold detection capability (50% POD) for a 0.10t flaw (a * = 0.25 inch) whereas the more effective NDE had a 50% POD for a 0.05t flaw (0.125 inch). In each case the POD curve provided significantly better detection capabilities for flaws of greater depths, such that flaw depths 0.25 and 0.50 inch respectively, or about twice the threshold size, could be detected with a probability of better than 90%. Figure 2 shows the predicted cumulative probability of leak (through-wall crack) as a function of the operating time (0 to 40 years). At 20 years (when the calculated CUF becomes 1.0) the cumulative leak probability is about 1.0E-02, or one chance in one hundred that the weld would fail. If no inspections are performed, the cumulative failure probability curve continues to rise and with an increasing failure rate. All of the alternative inspection scenarios (combinations of POD and inspection frequency) reduce the calculated failure probabilities, but some scenarios reduce the failure probability much more than others. Figure 3 shows the calculated failure rates in terms of failures per weld per year of operation. The most effective inspection (a* = 0.125-inch) reduces the failure rate by about an order of magnitude compared to the alternative of no inspection. In this case the failure rates during the second 20years of operation are actually substantially lower than the corresponding rates during the first 20years of operation. Some of the other less rigorous inspections of Figure 3 are also sufficiently effective to maintain the calculate failure rates at or below the rate that exists at the time (20years) when the CUF attains the limiting value of unity. For example, Figure 3 indicates that an Appendix L inspection with a 4-year frequency and a* = 0.125-inch would meet the probabilistic criteria. The alternative of a 2-year frequency with a* = 0.25-inch would also meet the criteria. The most effective inspection (a* = 0.125 inch) reduces the failure rate by about an order of magnitude compared to the alternative of no inspection. In this case the failure rates during the second 20 years of operation are actually substantially lower than the corresponding rates during the first 20 years of operation. Therefore, in this extreme case where thermal fatigue loading is significantly high, a 2 to 4 year inspection frequency will maintain the components reliability at design basis levels.

34-11

Deterministic Appendix L Calculations Having shown with the probabilistic method that acceptable inspection frequencies range from 2 to 4 years, the deterministic criteria of Appendix L were then applied. The flaw model was again a semi-elliptical surface flaw with aspect ratios of 6, 12 and 20. The postulated flaws had initial depths of 0.326 inches and 0.250 inches. Fatigue crack growth was calculated using the ASME Section XI Appendix C curve for stainless steel at 550oF with no environmental correction. The Section XI end of inspection interval allowable flaw was 0.75t (1.875 inch). The results of fatigue crack growth calculations for the four cases considered are shown in Table 3. In the first case, the existing Appendix L flaw model (aspect ratio of 6.0 and initial flaw depth of 0.326 inches) was applied. The Appendix L flaw depth was also assumed in cases 2 and 3 and Table 3 Appendix L Inspection Requirements Assumed Reference Initial Flaw Size Case 1 2 3 4 a/t
13%

Allowable Operating Period Cycles 33,207 23,060 17,572 40,500 P 6.65 yrs 4.61 yrs 3.51 yrs 8.10 yrs

Inspection P/2 3.32 yrs 2.30 yrs 1.75 yrs 4.05 yrs

a 0.326 0.326 0.326 0.250

Aspect Ratio
6

13% 13% 10%

12 20 6

the aspect ratios were increased to 12 and 20 respectively. In the final case, the Appendix L assumed flaw depth was reduced from 0.326 inches to 0.250 inches in order to be consistent with observed detection capabilities (11) and those anticipated in pc-PRAISE (10 and 12). These results were the basis for the Appendix L inspection frequencies of Table 3, which range from 2 to 4 years and are in excellent agreement with the conclusions of the probabilistic methodology. EPRI-NRC REASEARCH ACTIVITES This section describes ongoing research activities sponsored by EPRI and the NRC to address the needs and concerns expressed above (13). The objectives of this project are: 1. Improve the understanding of demonstrated NDE performance capabilities. 2. Upgrade the POD modeling capabilities in the pc-PRAISE code. 3. Understand how NDE performance factors impact piping reliability and associated augmented inspection strategies.

34-12

4. Evaluate the current Appendix L flaw tolerance procedure and its ability to define augmented inspection strategies that can offset the increases in the failure frequencies that are expected to occur as components continue to operate after calculated usage factors exceed 1.0. 5. Develop procedures and guidelines for a probabilistic fracture mechanics approach to component fatigue life extension. PDI Performance Data Evaluation In this task experts from the EPRI NDE Center and Pacific Northwest National Laboratory (PNNL) will be consulted to estimate probability of detection (POD) curves for fatigue cracks based on industry performance demonstration data. A matrix of POD curves will be developed for various NDE performance categories. Each performance category will represent a combination of performance factors such as: material type (stainless steel or carbon steel), access (double-sided or single-sided), section thickness, procedure (automatic or manual), and inspector qualification. pc-PRAISE POD Model Evaluation and Upgrade This task will evaluate the current pc-PRAISE POD model against industry performance data. A menu of performance category POD curves and/or closed form functions will be prepared and incorporated into the pc-PRAISE probabilistic fracture mechanics code. Augmented Inspection Strategy Sensitivity Study In this task a matrix of probabilistic fracture mechanics (PFM) calculations will be performed to identify augmented inspection strategies that would result in a failure frequency (failures per year) at or below the failure frequency associated with the design basis fatigue limit (e.g., cumulative usage factor of 1.0). The calculations will consider high cycle and low cycle fatigue loading, air and reactor water environments, and high stress through-wall gradients versus uniform throughwall stress distributions. PFM models recently developed at PNNL will be adapted to address the issues associated with fatigue crack initiation and crack growth in combination with the benefits of inservice inspections (Simonen et al. 1999). Appendix L Flaw Tolerance Flaw Procedure Assessment This task will assess whether the reductions in failure frequencies due to the Appendix L program of inspections (for locations with usage factors greater than 1.0) can offset the increases in the failure frequencies that are expected to occur as components continue to operate after calculated usage factors exceed 1.0. Recommended changes to Appendix L will be identified which will ensure the following: 1. Flaw size assumptions consider various NDE performance factors and are consistent with performance demonstration results.

34-13

2. Augmented inspection frequencies will maintain component reliability at a level consistent with the components original design basis. Probabilistic Alternative Approach This task will develop a probabilistic alternative approach that may be used to address NDE performance and inspection frequency issues related to damage tolerance assessments and component fatigue life extension. Industry Benefits This program will establish inspection frequencies that are consistent with the service conditions and the demonstrated performance levels of the NDE methods being applied. Case studies have shown that when inspections are performed at appropriate frequencies with reliable NDE methods, justification for continued operation of components with calculated fatigue usage factors that exceed original design limits can be made. This can provide a direct benefit to those plants seeking license renewal. In many cases it is expected that using performance demonstration data to optimize postulated Appendix L flaw size assumptions will eliminate much of the unnecessary conservatism in these procedures and reduce the resulting augmented inspection frequencies. Also, by expanding Appendix L to consider a probabilistic fitness-for-service approach, special issues or concerns related to life extension beyond the original 40 years can also be addressed. The PDI program has generated a very large set of data on the reliability of NDE as performed on operating U.S. nuclear vessels and piping. This data is particularly valuable both because of the number of data points and because it is based on trials performed by the same inspection personnel, procedures, and UT equipment employed in field inspections and because the components, welds and flaws were intended to be represent realistic field conditions. There would be great benefits to industry by documenting the collective quantitative trends of the PDI work, namely: Industry could potentially take credit for the improved ISI to gain relief from present code and regulatory practices (i.e. the Appendix L improvements as described above, reduced inspection intervals for Section XI and augmented inservice inspection programs, plant life extension, etc.) At present the fracture mechanics analyses, including probabilistic methods, may be based on pessimistic estimates for NDE performance, or no credit can be taken for the impact of inspections on the structural integrity of components such as the RPV in the context of PTS risk Industry could justify reductions in testing requirements for performance demonstration efforts by demonstrating a high statistical confidence in collective NDE detection and sizing capability for the cohort of inspections teams that are performing field inspections.

34-14

Nevertheless within these restrictions the collective trends of the data could be characterized in considerable detail and at a level that could be used to the benefit of the industry that would enable structural integrity evaluations to be performed in a much more realistic manner than is now possible. Fracture mechanics computer codes can now include the benefits of inservice inspections at given intervals with specific techniques. Inputs are needed for probability of detection curves and in some cases the expected sizing errors. Such calculations have been performed as part of research efforts to support applications of risk informed inservice inspections. However, the uncertainties in estimated POD curves have been large, such that it has been difficult to take credit for ISI in terms of regulatory requirements. Any improvements in ISI reliability have essentially gone unrecognized with respect to use in flaw evaluations, establishing inspection intervals, plant specific PTS evaluations, etc. The PDI data would permit POD curves and sizing errors to be established with a sound technical basis that reflects the current demonstrated NDE capabilities of the nuclear industry. It is further noted that the current Section XI Appendix VIII tables for the minimum number of samples and passing scores for detection, false calls and sizing errors were based on statistical evaluations of round robin studies to estimate the performance capabilities of the inspection process and did not assume a qualification effort of the type encompassed by the PDI activities. Rather the current bases have assumed that each organization would have an independent performance demonstration activity and the statistical calculations did not take credit for the added knowledge and confidence gained from subjecting a large group of organizations to a common effort that would address the collective reliability of a population of inspections teams. Such an approach could reduce the need for each passing team to inspect as many samples or to demonstrate as high level of successful detection and sizing performance. In other words, having the inspection process quantified (personnel, equipment, and procedures) provides data necessary to check the Appendix VIII assumptions for performance demonstration and to make it possible to optimize the process to achieve the desired screening (pass/fail threshold) most economically. CONCLUSIONS The results for the above example show that inspection frequencies based on Appendix L evaluations are consistent with inspection frequencies derived from a probabilistic assessment. It is seen that inspections at appropriate frequencies with reliable NDE methods can justify continued operation for components with calculated fatigue usage factors that exceed original design limits. It is even possible with an aggressive inspection program to decrease failure frequencies during the later periods of plant life to the same levels that existed relatively early in life. It should be noted that the Appendix L approach was primarily developed for location with high fatigue for which there is only a perceived potential for cracking. The approach would not be applied in other cases for which the operating plant has actually experienced degradation (i.e.

34-15

detection of small cracks or leakage). In such cases, the preferred strategy would be one of repair/replacement, augmented inspections, and other mitigative actions. Future work will to expand the scope of the calculations to address other piping sizes, materials, and cyclic stress conditions. The goal will be to establish inspection frequencies that are consistent with the service conditions and the demonstrated performance levels of the NDE methods being applied. It is expected that, in many cases, the inspection frequencies required by the simplified rules of Appendix L are overly conservative. As shown in this paper, using performance demonstration data to optimize the Appendix L flaw size could eliminate much of this conservatism. In other cases, the calculations may indicate a need for improved NDE methods and/or for more frequent inspections. By expanding Appendix L to consider a probabilistic fitness-for-service approach, special issues or concerns related to life extension beyond the original 40 years can also be addressed. For example, the need to address potential fatigue cracking at outside surface locations, for which the NDE procedure will be based on dye penetrant and magnetic particle methods rather than UT methods. In the end, the analyst could have a choice of 3 alternative fatigue qualification methods, each having a sound technical basis that will ensure piping reliability is maintained throughout its desired service life. The flaw tolerance procedure contained in Appendix L is ineffectual, primarily due to very conservative flaw size assumptions. An improved technical basis is needed for the flaw models and rules use to establish inspection frequencies in Appendix L. It is expected that NDE performance data based on PDI will provide substantial information to develop appropriate flaw size assumptions and POD curves for use in structural integrity evaluations. The resultant products from this work will be especially useful for utilities that may have fatigue concerns or are planning to seek license renewal. ACKNOWLEDGMENTS The author wishes to acknowledge Dr. Steven Doctor and Dr. Fred Simonen of Pacific Northwest National Laboratory. The insights they provided with regard to the industry performance demonstration initiative and the development of Section XI Appendix VIII were incorporated directly into this paper. REFERENCES 1. Cooper, W. E., The Initial Scope and Intent of the Section III Fatigue Design Procedures, PVRC Workshop on Environmental Effects on Fatigue Performance, Clearwater Beach, FL, January 20, 1992, p. 1-6.

34-16

2. Gosselin, S. R., ASME Section XI Philosophy Related to Operating Nuclear Power Plant Fatigue Damage Protection, PVP Volume 313-2, International Pressure Vessel Piping Codes and Standards: Volume 2- Current Perspectives ASME 1995. 3. American Society of Mechanical Engineers, ASME Boiler and Pressure Vessel Code, Section XI, Non-mandatory Appendix L, Operating Plant Fatigue Assessments, 1995 Edition, New York, July 1995. 4. Gosselin, S. R., A. F. Deardorf, and D. W. Peltola, Fatigue Assessments in Operating Nuclear Power Plants, PVP-Vol. 286, pp. 3-18, Changing Priorities of Codes and Standards, American Society of Mechanical Engineers, New York, 1994. 5. EPRI TR-100252, ASME Section XI Task Group on Fatigue in Operating Plants, Metal Fatigue in Operating Nuclear Power Plants, April 1992. 6. EPRI TR-104691 Final Report Operating Nuclear Power Plant Fatigue Assessments, Palo Alto, CA, April 1995. 7. Connor, J. T., BG&E Report to ASME Section XI Task Group Meeting on Operating Plant Fatigue Assessments, August 1999. 8. EPRI TR-107515 Final Report, Evaluation of Thermal Fatigue Effects on Systems Requiring Aging Management Review for License Renewal for Calvert Cliffs Nuclear Power Plant, Palo Alto, CA, December 1997. 9. Gosselin, S. R. and Simonen, F. A., A Probabilistic Basis for Damage Tolerance Assessments and Component Fatigue Life Extension, PVP-Vol. 383, ASME, 1999. 10. Harris, D. O., and Dedhia, D. D., Theoretical and Users Manual for pc-PRAISE, A Probabilistic Fracture Mechanics Computer Code for Piping Reliability Analysis, NUREG/CR-5864, U. S. Nuclear Regulatory Commission, Washington, D.C. 11. Heasler, P.G., and Doctor, S. R., A comparative Analysis of Round Robin Studies, Presentation at 1st International Conference on NDE in Relation to Structural Integrity for Nuclear Pressurized Components, October 1998, Amsterdam, The Netherlands. 12. Simonen, F. A., and Khaleel, M. A., A Model for Predicting Vessel Failure Probabilities Due to Fatigue Crack Growth, PVP-Vol. 304, Fatigue and Fracture Mechanics in Pressure Vessels and Piping, pp. 401-416, American Society of Mechanical Engineers, New York, 1995. 13. Carter, R. G., Mayfield, M., Gosselin, S. R., and Simonen, F. A., An Evaluation of Flaw Tolerance Procedures Used for Component Fatigue Life Qualification, 8 th International Conference on Nuclear Engineering, April 2000.

34-17

14. Simonen, F. A., Phan, H. K., Harris, D. O., Dedhia, D., Kalinousky D. N., and Shaukat, S. K., Evaluation of Environmental Effects on Fatigue Life of Piping, Presented at 27 th Water Reactor Safety Information Meeting, Bethesda, MD, October 25-27, 1999.

34-18

Operating Plant Fatigue Concern

Are Actual Load Cycles < Design?

Yes

No

Section III Fatigue CUF Evaluation

CUF<1.0 ?

Yes

No

Appendix L Flaw Tolerance Evaluation

Continue Operation Inspect at Normal Section XI Frequency

Fitness for Continued Service ?

Yes

No

Repair or Replace

Continue Operation Inspect at Increased Frequency

Figure 1: Operating Plant Fatigue Assessments

34-19

5.E-02

C:\PRAISE96\APPENL\APL1.XLS December 1998

Fatigue Usage Factor > 1.0 Inspections as per Appendix L

4.E-02

Cumulative Probability of Leak

Baseline Case No Additional ISI 3.E-02

ISI @ 4 Yr a* = 0.25 Inch

ISI @ 4 Yr a* = 0.125 Inch

2.E-02

ISI @ 2 Yr a* = 0.25 Inch

1.E-02
ISI @ 4 Yr a* = 0.125 Inch

ISI @ 2 Yr a* = 0.125 Inch

0.E+00 0 5 10 15 20 25 30 35 40 45

Time, Years

Figure 2: Cumulative Leak Probability

1.5E-03
C:\PRAISE96\APPENL\APL1.XL S

Fatigue Usage Factor > 1.0 Inspections as per Appendix L

Leak Frequency, Leaks per Weld per Year

Baseline Case No Additional ISI

1.0E-03 ISI @ 4 Yr a* = 0.25 Inch "Acceptable" Leak Frequency 5.0E-04 ISI @ 4 Yr a* = 0.125

ISI @ 2 Yr a* = 0.25 Inch ISI @ 2 Yr a* = 0.125 Inch

0.0E+00 0 5 10 15 20 25 30 35 40 45 50

Time, Years

Figure 3: Leak Frequency

34-20

35
AN UPDATE ON THE CONSIDERATION OF REACTOR WATER EFFECTS IN CODE FATIGUE INITIATION EVALUATIONS FOR PRESSURE VESSELS AND PIPING

Hardayal S. Mehta GE Nuclear Energy San Jose, California

35-1

AN UPDATE ON THE CONSIDERATION OF REACTOR WATER EFFECTS IN CODE FATIGUE INITIATION EVALUATIONS FOR PRESSURE VESSELS AND PIPING

Hardayal S. Mehta GE Nuclear Energy San Jose, California

ABSTRACT Reactor water environmental fatigue effects are currently a subject of active discussion among the industry, the standards making bodies and the regulatory bodies in U.S. and Japan. The PVRC has recently recommended to the BNCS a procedure that can potentially be implemented into the ASME Code. The objective of this paper is to provide details of the procedure and discuss the rationale for some of the modifications incorporated into the procedure prior to the recommendation to BNCS. Also, a brief discussion of the current Regulatory and ASME Code activities in progress on this subject is provided.

The Steering Committee on Cyclic Life and Environmental Effects (CLEE) of the Pressure Vessel Research Council (PVRC) has reviewed the EPRI/GE environmental fatigue evaluation methodology and has endorsed it as a preferred approach to evaluate the environmental fatigue effects in any ASME Code fatigue evaluations [Yukawa, 1998]. Discussions in the various Task Groups of the CLEE last year resulted in several modifications to the methodology and are described next.

INTRODUCTION Mehta and Gosselin (1995, 1998) provide the details of the proposed EPRI/GE methodology to account for the reactor water environmental fatigue effects in the ASME Code fatigue evaluations. The methodology essentially consists of applying an environmental correction factor, Fen, that is computed using equations involving environmental parameters, to each load set pair that does not meet the threshold criteria. Argonne National Laboratory (ANL) proposed statistical relationships were used to develop the mathematical expressions for Fen. The materials covered are carbon, low alloy and stainless steels. Mehta and Gosselin (1995) also included the format for a proposed non-mandatory appendix for potential implementation into the ASME Code. Mehta (1998a, 1998b) provided an example of the application of the methodology to representative BWR locations. An update of the methodology including the impact of the recent ANL statistical relationships, is described by Mehta (1999). Based on their extensive research work, the Japanese researchers also have independently proposed a somewhat different approach for the calculation of a Fen [e.g., Higuchi, 1999] for carbon and low alloy steels.

RECENT PVRC ACTIONS Last year, the CLEE had a detailed discussion on the guidelines to be used in developing the fatigue life environmental correction factor for potential ASME Code implementation [WRC, 1999]. Items agreed on in the discussion include the following: (i) The basis for the calculation of the fatigue life correction factor (Fen) will be the room temperature air S-N curve which means that Fen = ratio of fatigue life in room temperature air to life in the service temperature water. (ii) Include the concept of an effective Fen for the purpose of maintaining the existing Code margins with respect to the mean water curves. This can be done by defining an effective Fen = Fen/Z where Z = 3 for carbon and low alloy steels and Z = 1.5 for austenitic stainless steels. At present, Z=1.5 will be applied to both cast and wrought stainless steels. Also, the value of effective F en must be 1.0. (iii) To minimize confusion in Code applications, the basis and rationale for effective Fen calculations would be separated from the procedures for the calculation of basis Fen factors. (iv) The following values of threshold strain amplitudes were discussed and accepted: For carbon and low alloy steels 0.07% to 0.08% with a ramp function between the two values; for wrought and cast austenitic steels 0.1% to 0.11% with a ramp between the two values.

35-3

(v) The temperature threshold for wrought and cast austenitic stainless steels was discussed and it was agreed to replace the present threshold = 200C (derived from ANL correlations) by a ramp function from 180C to 220C. (vi) The criterion and threshold values for austenitic stainless steels were agreed on with a recognition of the scarcity of data. After the meeting, Mr. Higuchi communicated that the Japanese project on environmental fatigue would review the available data in Japan and send the results as soon as possible. The proposed methodology of Mehta and Gosselin (1995) in the form of a non-mandatory appendix was modified to reflect the preceding discussion and is included here as Appendix A. The PVRC forwarded to the Board on Nuclear Codes and Standards (BNCS) the proposed non-mandatory appendix text along with the suggestions for consideration and potential Code implementation [WRC, 1999].

Potential Section XI Implementation The most logical place for enabling words in Section XI is Paragraph IWB-3740 which pertains to operating plant fatigue assessment. For example, the IWB-3740(a) could be modified as follows: (a) Appendix L provides procedures that may be used to assess the effects of thermal and mechanical fatigue concerns on component acceptability for continued service. When the environmental effects on the fatigue analysis are considered significant, such effects may be accounted for by using the methods described in Appendix XX. The reference to Appendix XX in the above sentence could be a non-mandatory appendix in Section III.

POTENTIAL CODE IMPLEMENTATION APPROACHES This section presents the suggested changes in the appropriate articles of ASME Section III and XI to provide stress analysts with enabling words to include environmental effects in the Code fatigue evaluations. The evaluation procedure itself could be first developed as a Code Case and in the future can potentially be incorporated in the form of a non-mandatory appendix to the Code.

Potential Implementation as Section XI Code Case This could be in the form of an inquiry worded as follows: Inquiry: Section XI, Division 1, IWB-3740 has provision for assessment of the effects of thermal and mechanical fatigue concerns on component acceptability for continued service. What procedures may be used to account for the reactor water environment effects in such assessments? Reply: It is the opinion of the Committee that the thermal and mechanical fatigue concerns from the reactor water environment effects may be evaluated in accordance with the following requirements (Appendix A).

Potential Section III Implementation For the evaluation of new designs, the environmental fatigue procedures would be part of Section III. The enabling words could be added in Paragraphs NB-3200 and NB-3600.

NB-3200. The procedure for fatigue analysis in NB-3200 is contained in NB-3224.4(e), Procedure for Analysis for Cyclic Loading. Paragraph (5) of NB-3224.4 (e) stipulates six steps for calculating the cumulative fatigue damage when there are two or more types of stress cycles which produce significant stresses. Add the following step to NB-3224.4(e): Step 7: When the environmental effects on the fatigue life are considered significant, such effects may be accounted for by using the methods described in Appendix XX.

RECENT REGULATORY ACTIONS The following excerpts from a recent letter written by the Advisory Committee on Reactor Safeguards (ACRS) to the US NRC [Powers, 1999] provides a good summary of the current regulatory thinking in relation to environmental fatigue effects:

Recommendations We agree with the staffs proposal that GSI-190 be resolved without any additional regulatory requirements. The staff should ensure that utilities requesting license renewal consider the management of environmentally assisted fatigue in their aging management programs.

NB-3600. Paragraph NB-3610 specifies the general requirements of piping design. A sub-paragraph NB-3614 worded as follows may be added to provide enabling words: NB-3614 Environmental Effects. When the environmental effects on the fatigue analysis required by NB-3650 are considered significant, such effects may be accounted for by using the methods described in Appendix XX. The other location that needs to be modified is Subparagraph 3653.8 which might be added as follows: NB-3653.8 Consideration of Environmental Effects. When the environmental effects are considered significant, the cumulative fatigue damage shall be that calculated using the procedures of Appendix XX.

Background The effects of fatigue for the 40-year initial reactor license period were studied and resolved under GSI-78, Monitoring of Fatigue Transient Limits for Reactor Coolant System, and GSI-166, Adequacy of Fatigue Life of Metal Components. The staff concluded that risk from fatigue failure of components in the reactor coolant pressure boundary was very small for 40-year plant life. In our March 14, 1996 letter, we agreed with the staffs conclusion. GSI-190 was established to address the residual concerns of GSI78 and GSI-166 regarding the environmental effects of fatigue on pressure boundary components for 60-years of plant operation. The scope of GSI-190 included design-basis fatigue transients, studying the

35-4

probability of fatigue failure and its effects on core damage frequency (CDF) of selected metal components for 60-year plant life.

Discussion Resolution of GSI-190 was based on the results of an NRCsponsored study performed by the Pacific Northwest National Laboratory (PNNL). In that study, PNNL examined design-basis fatigue transients and the probability of fatigue failure of selected metal components for 60-year plant life and the resulting effects on CDF. The PNNL study showed that some components have cumulative probabilities of crack initiation and through-wall growth that approach unity within the 40- to 60-year period. The maximum failure rate (through-wall cracks per year) was in the range of 10-2 per year, and those failures were associated with high cumulative usage factor locations and components with thinner walls, i.e., pipes more vulnerable to through-wall cracks. There was only a modest increase in the frequency of through-wall cracks in major reactor coolant system components having thicker walls. In most cases, the leakage from these through-wall cracks is small and not likely to lead to core damage. Therefore, the projected increased frequency in through-wall cracks between 40- and 60-years of plant life does not significantly increase CDF. Based on the low contributions to CDF, we agree with the proposed resolution of GSI-190. Environmentally assisted fatigue degradation should be addressed in aging management programs developed for license renewal. Minimization of leakage is important for operational safety, occupational doses, and for continued economic viability of the plants. The PNNL probabilistic study the ACRS referred to was conducted by Simonen, et al. (1999). Recently, the NRC wrote a letter to the BNCS [Craig, 1999] requesting ASME action to address issues related to the effects of the reactor water environment on the reduction of fatigue life of light water reactor components.

(Oconee Nuclear Station) have essentially proposed to use the PVRCpreferred approach to evaluate the environmental fatigue effects for the license renewal terms [Tuckman, 1999; Cruse, 1999]. Given that a significant number of plant owners are expected to apply for the license renewal of their units, the industry is currently working with the license renewal committee of the EPRI and the Nuclear Energy Institute (NEI) to develop an approach to this issue that does not impose a significant burden while assuring safe plant operation.

ASME CODE ACTIVITY In the Section XI Code space, the environmental fatigue issue is currently being addressed by the Task Group on Operating Plant Fatigue Assessments reporting to the Working Group on Operating Plant Criteria. The PVRC-endorsed approach is under discussion in the Task Group. The responsibility for the Code fatigue curves in the Section III space is with the Subgroup on Design and Subgroup on Materials, Fabrication and Examination. However, the main responsibility for the Code S-N curves and the fatigue design methods is with the Subgroup on Fatigue Strength reporting to the Subcommittee on Design. The Operating Plant Fatigue Assessments Task Group plans to hold joint meetings with the Subgroup on Fatigue Strength to coordinate the activities related to the environmental fatigue issue.

DISCUSSION A key difference between the PVRC-preferred approach outlined in the Appendix and the Japanese approach [e.g., Higuchi, 1999] is the fact that the former uses a working definition of moderate environmental effects that are assumed to have already been accounted for in the factor of 20 on cycles used in developing the original Code S-N curves. Therefore, the calculated values of Fen are divided by a factor (Z factor in the Appendix) to obtain an effective value of Fen. Some of the supporting arguments in the favor of an effective Fen approach are the following: A recent paper by Chopra and Shack (1999) provides a good rationale for taking some credit out of the factor of 4 assigned to surface finish, atmosphere, etc. They state, Because carbon and lowalloy steels and austenitic SSs develop a corrosion scale in LWR environments, the effect of surface finish may not be significant, i.e., the effects of surface roughness are included in environmentally assisted decrease in fatigue life in LWR coolant environments. In water, the sub-factor on life to account for surface finish effects may be as low as 1.5 or may be eliminated completely; a factor of 1.5 on strain and 7 on cycles is adequate to account for the uncertainties that arise from material and loading variability. Therefore, the factor of 20 on life that is used in developing the design fatigue curves includes, as a safety margin, a factor of 3 or 4 on life that may be used to account for the effects of environment on the fatigue lives of these steels. Cooper (1992) estimated an average factor of 3 on fatigue initiation with respect to the design curve in the PVRC tests on low alloy steel vessels tested with water at room temperature. The analytical expressions for Fen give values greater than 1 even when the parameters are below the threshold. For example, in the case of stainless steel, the Fen equation (Equation 3 in Appendix) predicts a value of 2.55 even at room temperature. Similarly, the value is 2.47 for low alloy steels. Ideally, the Fen values should approach 1 as the threshold values are reached. Development of a sound technical basis for the effective Fen approach is currently under consideration by the CLEE.

INDUSTRY ACTIVITIES The industry concern with the environmental fatigue effects has mostly been in the context of the plant license renewal. Baltimore Gas & Electric (Calvert Cliffs Nuclear Power Plant) and Duke Energy

SUMMARY A PVRC-preferred procedure to address reactor water environmental effects in pressure vessel and piping fatigue evaluations, was described in this paper. This subject is currently under active discussion with the regulatory agencies, the ASME Code and the industry. The paper briefly described the activities of these groups in relation to this subject.

35-5

ACKNOWLEDGEMENT The author would like to acknowledge many helpful discussions with and encouragement received from Drs. Sumio Yukawa and W.A. Van Der Sluys.

NONMANDATORY APPENDIX XX

FATIGUE EVALUATIONS INCLUDING ENVIRONMENTAL EFFECTS

REFERENCES Chopra, O.K. and Shack, W.J., 1999, Methods for Incorporating Effects of LWR Coolant Environment into ASME Code Fatigue Evaluations, ASME PVP-Volume 386, pp. 171-181. Cooper, W.E., 1992, The Initial Scope and Intent of the Section III Fatigue Design Procedures, PVRC Workshop on Environmental Effects on Fatigue Performance, Clearwater Beach, FL, January 20. Craig, J.W., Director, Division of Engineering Technology, office of Nuclear Regulatory Research, NRC, Letter to J.H. Ferguson, Chairman, BNCS, December 1, 1999. Cruse, C.S., Baltimore Gas and Electric Company, Letter to USNRC Document Control Desk, Confirmatory Item 3.2.3.3-1, July 2, 1999. Higuchi, M., 1999, Fatigue Curves and Fatigue Design Criteria for Carbon and Low Alloy Steels in High-Temperature Water, ASME PVP-Volume 386, pp. 161-169. Hollinger, G.L., Executive Director, PVRC, Letter to J.H. Ferguson, Chairman, BNCS, October 31, 1999. Mehta, H.S. and Gosselin, S.R., 1995, An Environmental Factor Approach to Account for Reactor Water Effects in Light Water Reactor Pressure vessel and Piping Fatigue Evaluations, EPRI Report No. TR105759. Also, ASME PVP-Volume 323 (1996). Mehta, H.S. and Gosselin, S.R., 1998, Environmental Factor Approach to Account for Water Effects in Pressure Vessel and Piping Fatigue Evaluations, Nuclear Engineering and Design Journal, Volume 181, pp. 175-197. Mehta, H.S., 1998a, Environmental Fatigue Evaluations of Representative BWR Components, EPRI Report No. TR-107943. Mehta, H.S., 1998b Application of EPRI/GE Environmental Factor Approach to Representative BWR Pressure Vessel and Piping Fatigue Evaluations, ASME PVP-Vol. 360, pp. 413-425. Mehta, H.S., 1999, An Update on the EPRI/GE Environmental Fatigue Evaluation Methodology and Its Applications, ASME PVPVolume 386, pp. 183-193. Powers, D.A., Chairman ACRS, Letter to William D. Travers, Executive Director for Operations, US NRC, Dated December 10, 1999,Subject: Proposed Resolution of GSI-190. Simonen, F.A., et al., 1999, Evaluation of Environmental Fatigue Effects on Fatigue Life of Piping, Presented at the 27th Water Reactor Safety Information Meeting, October 25-27. Tuckman, M.S., Duke Energy Corporation, Letter to USNRC Document Control Desk, response to SER Open Item 4.2.3-2, October 15, 1999. Welding Research Council Progress Reports, 1999, Volume LIX, No. 5/6, May/June, pp. 33-36. Yukawa, S., 1998, PVRC Progress Report to BNCS, October 9, 1998, Atlanta, Georgia, Welding Research Council Progress Reports, Volume LIII, No. 9/10, Sept./Oct.

ARTICLE X-1000

SCOPE This Appendix provides methods for performing fatigue usage factor evaluations of reactor coolant system and primary pressure boundary components when the effects of reactor water on fatigue initiation life are judged to be significant.

X-1100 ENVIRONMENTAL FATIGUE CORRECTION The evaluation method uses as its input the partial fatigue usage factors U1, U2, U3, .....Un, determined in Class I fatigue evaluations. In Class I design by analysis procedure, the partial fatigue usage factors are calculated for each type of stress cycle in paragraph NB3222.4(e)(5). For Class I piping products designed using NB-3600 procedure, Paragraph NB-3653 provides the procedure for the calculation of partial fatigue usage factors for each of the load set pairs. The cumulative fatigue usage factor, Uen, considering the environmental effects is calculated as the following: Uen = U1Fen, 1 + U2Fen, 2 + U3Fen, 3 ... UiFen, i ....+ UnFen, n where, Fen,i is the effective environmental fatigue correction factor for the ith stress cycle (NB-3200) or load set pair (NB-3600).

X-1200 ENVIRONMENTAL FACTOR DEFINITION X-1210 The nominal values of environmental fatigue correction factors are to be calculated using the expressions below.

Carbon Steel Fen,nom = [exp (0.559 - 0.101S*T*O**)] Low Alloy Steel Fen,nom = [exp (0.903 - 0.101S*T*O**)] (2) (1)

Stainless Steels (wrought and cast) Fen,nom = exp [0.935 - T*O**)] (3)

X-1260 The effective environmental fatigue correction factor, Fen, is obtained by dividing the nominal value calculated in X-1210 with a material-specific factor which accounts for moderate

35-6

environmental fatigue effects already included in the S-N curves of Figures I-9.1 and I-9.2. Fen = Fen,nom/Z, but no less than 1.0

Where, Z = 3.0 for carbon and low alloy steels and 1.5 for wrought and cast stainless steels.

X-1300 EVALUATION PROCEDURES For some types of stress cycles or load set pairs any one or more than one environmental parameters are below the threshold value for significant environmental fatigue effects. The value of the environmental fatigue correction factor, Fen for such types of stress cycles or load set pairs shall be equal to 1.0. Article X-2000 provides procedure for threshold criteria evaluation. The procedures for the evaluation of Fen factors for design by analysis and for Class I piping products fatigue evaluations are provided in X-3000.

= Elapsed time between the start of temperature transient and the time when Tt is reached, seconds tT,th = Elapsed time between the start of decreasing temperature transient and the time when metal surface in contact with fluid reaches threshold temperature, seconds Uen = Cumulative fatigue usage factor including the environmental effects Ui = Cumulative fatigue usage factor for load set pair i obtained by using Code fatigue curves i = Strain range for load set pair i, % = Strain rate, %/second * = Transformed strain rate *() = Transformed strain rate at elapsed time equal to

tt

ARTICLE X-2000 ENVIRONMENTAL FATIGUE THRESHOLD CONSIDERATIONS

X-1400 NOMENCLATURE The symbols adopted in this Appendix are defined as follows: E = Youngs Modulus, psi Fen = Effective environmental correction factor applied to fatigue usage calculated using Code fatigue curves Fen() = Environmental correction factor calculated at a specific instant in time, . Fen,int = Environmental correction factor based on integrated approach. DO = Dissolved oxygen content of water (ppm) O* = Transformed oxygen content S = Sulfur content of carbon and low-alloy steels, weight % S* = Transformed sulfur content Salt = Alternating stress amplitude, psi Srange = Range of stress intensity associated with a transient cycle, psi T = Temperature (C) T* = Transformed temperature. Ta = Average temperature on side a during a temperature transient Tb = Average temperature on side b during a temperature transient Tc = Sum of |Ta - Tb|, |T1| and |T2| for temperature transient producing compressive stresses at the component surface in contact with fluid Tm = Metal temperature during a temperature transient at surface in contact with fluid Tt = Sum of |Ta - Tb|, |T1| and |T2| for temperature transient producing tensile stresses at the component surface in contact with water T1 = Linear temperature gradient through a component wall during a temperature transient T2 = Nonlinear temperature gradient through a component wall during a temperature transient

X-2000 SCOPE This Article provides procedure for screening out types of stress cycles or load set pairs for which any one or more than one environmental parameters are below the threshold value for significant environmental fatigue effects. The value of the environmental fatigue correction factor, Fen for such types of stress cycles or load set pairs shall be equal to 1.0.

X-2100 STRAIN AMPLITUDE THRESHOLDS X-2110 The strain amplitude threshold for carbon and low alloy steels is 0.07%. Fen values shall be used at strain amplitudes equal to or exceeding 0.08%. A linear interpolation may be used to calculate Fen values for strain amplitudes between 0.07% and 0.08%. X-2120 The strain amplitude threshold for wrought and cast stainless steels is 0.10%. Fen values shall be used at strain amplitudes equal to or exceeding 0.11%. A linear interpolation may be used to calculate Fen values for strain amplitudes between 0.10% and 0.11%. X-2130 Calculate the strain amplitude, i associated with a type of stress cycle or load set pair i by multiplying the alternating stress intensity Salt i by 100 and dividing by the modulus of elasticity E. The value of E shall be obtained from the applicable design fatigue curves of Figs. I-9.0. X-2140 If the value of i calculated in X-2130 for a load set pair is less than or equal to appropriate value from X-2110 or X-2120, that load set pair satisfies the threshold criterion and the value of F en i is 1.0. No further evaluation with respect to other threshold values need be made for this load set pair. X-2200 STRAIN RATE THRESHOLD The strain rate threshold is 1.0%/second for carbon and low alloy steels, and 0.4%/second for wrought and cast stainless steels. A load set pair involving only the seismic loading satisfies the strain rate threshold criterion for strain rate and the value of Fen i is 1.0. No

35-7

further evaluation with respect to other threshold values need be made for this type of stress cycle or load set pair. If the strain rate associated with the tensile stress load set for any other load set pair exceeds the threshold value, Fen is 1.0 for that load set pair.

X-2300 TEMPERATURE THRESHOLD X-2310 The temperature threshold for carbon and low alloy steels is 150C. X-2320 The temperature threshold for wrought and cast stainless steels is 180C. X-2330 Define the effective temperature, T associated with a type of stress cycle or load set pair i as equal to the higher of the highest temperatures in the two transients or load sets constituting the type of stress cycle or load set pair. X-2340 If the temperature calculated in step (b) is less than or equal to the threshold value, the stress cycle or load set pair satisfies the threshold criterion for temperature and the value of Fen i is 1.0.

where, Srange i is the stress difference range for cyclei as determined in NB-3224.4(e)(5) and the tmax is the time in seconds when the stress difference reaches a maximum from the start of the temperature transient. This calculation is performed only for the step down temperature transient or other tensile stress producing cycle in the stress cycles constituting a pair. X-3212 The transformed strain rate * for carbon and low alloy steels is obtained as the following: * = 0 * = ln() * = ln(0.001) ( > 1%/sec) (0.001 < < 1%/sec) ( < 0.001%/sec)

X-3213 The transformed strain rate * for stainless steels is obtained as the following: * = 0 * = ln(/0.4) * = ln(0.0004/0.4) ( > 0.4%/sec) (0.0004 < < 0.4%/sec) ( < 0.0004%/sec)

X-2400 DISSOLVED OXYGEN THRESHOLD This is applicable only to carbon and low alloy steels. (a) Define the effective dissolved oxygen content, DO ass ociated with a type of stress cycle or load set pair i as equal to the higher of the highest oxygen content in the two transients or load sets constituting the type of stress cycle or load set pair. (b) If the value of DO determined in step (a) for a type of stress cycle or load set pair is less than or equal to 0.05 ppm, that type of stress cycle or load set pair satisfies the threshold criterion and the value of Fen i is 1.0.

X-3220 Determination of Transformed Temperature X-3221 The temperature, T associated with a stress cycle i is equal to the higher of the highest metal temperatures in the two transients constituting the stress cycle or load set pair. X-3222 The transformed temperature T* for carbon and low alloy steels is obtained as the following: T* = 0.0 T* = T-150 (T< 150C) (T> 150C)

ARTICLE X-3000 ENVIRONMENTAL FACTOR EVALUATION

X-3223 The transformed temperatures T* for stainless steels are obtained as the following: T* = 0.0 T* = (T-180)/40 T* = 1.0 (T<180C) (180C<T<220C) (T>220C)

X-3100 SCOPE This Article provides procedure for calculating the Fen factors for types of stress cycles (NB-3200) or load set pairs (NB-3600). Only the types of stress cycles or load set pairs that do not meet the threshold criteria of X-2000 need to be considered for F en calculation.

X-3200 EVALUATION PROCEDURE FOR DESIGN BY ANALYSIS

X-3230 Determination of Transformed DO X-3231 For carbon and low alloy steels, the effective dissolved oxygen content, DO associated with a load set pair i is equal to the higher of the highest oxygen level in the two transients constituting the load set. The transformed DO, O* is obtained as follows: O* = 0 (DO<0.05 ppm) O* = ln(DO/0.04) (0.05 ppm < DO < 0.5 ppm) O* = ln(12.5) (DO > 0.5 ppm) X-3232 For wrought stainless steels, the effective dissolved oxygen content, DO associated with a load set pair i is equal to the lower of the oxygen level in the two transients constituting the load set. The transformed DO, O* is obtained as follows:

X-3210 Determination of Transformed Strain Rate

X-3211 The strain rate (%/sec) for a stress cycle is determined as the following: = Srange i 100/Etmax

35-8

O* = 0.260 O* = 0.172 X-3233

(DO < 0.05 ppm) (DO > 0.05 ppm)

X-3620 Determination of Transformed Temperatures The transformed temperatures shall be obtained as described in X-3220.

For cast stainless steels, O* = 0.260 X-3630 Determination of Transformed DO The transformed DO shall be obtained as described in X-3230.

X-3240 Determination of Transformed Sulfur for Carbon & Low Alloy Steels The sulfur content S in terms of weight percent might be obtained from the certified material test report or an equivalent source. If the sulfur content is unknown, then its value shall be assumed as 0.015%. The transformed sulfur, S* is obtained as the following: S* = S S* = 0.015 (0<S<0.015 wt%) (S>0.015 wt%)

X-3640 Determination of Transformed Sulfur for Carbon and Low Alloy Steels The transformed sulfur shall be obtained as described in X-3240.

X-3650 Determination of Fen The environmental correction factor F en i shall be calculated using equations given in X-1200.

X-3250 Determination of Fen The environmental correction factor F en i for a type of stress cycle and the cumulative fatigue usage factor shall be calculated using equations given in X-1200.

X-3660 Determination of Fen Based on Integrated Approach When the results of detailed transient analyses are available to predict strain rate, such results may be used to reduce conservatisms in the calculated values of Fen. The following expression or equivalent shall be used: tT,th Fen,int = (1/tT,th) 0 [Fen()]d The preceding value of Fen may be used in lieu of the Fen value calculated in X-3650. Fen() is the appropriate environmental factor derived from X-1200, with time dependent properties/factors for the time in the transient where the temperature exceeds the threshold value.

X-3260 Determination of Fen Based on Damage Approach Procedure similar to that described in X-3660 may be used to remove some of the conservatism built into the Fen i determined in X-3250.

X-3600 EVALUATION PROCEDURE FOR PIPING The procedures in this Article use the input information and the partial fatigue usage results from the NB-3650 fatigue evaluation. The example of specific load set information needed is: internal pressure, the three moment components, |Ta-Tb|, T1 and T2. When the detailed results of one-dimensional transient heat transfer analyses are available in the form of time history of |Ta-Tb|, T1 and T2, such results may be used to reduce conservatisms in the calculated values of environmental correction factor.

X-3610 Determination of Strain Rate The strain rate (%/sec) for a load set pair i is determined as the following: i = 200Salt i [Tt/(Tt + Tc )]/(Ett) where, Salt i is the alternating stress intensity for load set pair i calculated in NB-3653.3. This calculation is performed only for the step down temperature transient in a load set pair. The transformed strain rate i* shall be obtained as described in X-3210.

35-9

36
CODES AND THERMAL FATIGUE: STATUS AND ON-GOING DEVELOPMENT
C. Faidy EDF-SEPTEN

36-1

36-3

36-4

36-5

36-6

36-7

VIBRATION/HIGH CYCLE FATIGUE

37
VIBRATION FATIGUE TESTING OF SOCKET WELDS, PHASE II
Paul Hirschberg, Peter C. Riccardella Structural Integrity Associates Michael Sullivan, Jon Schletz Pacific Gas & Electric Co. Robert Carter EPRI

37-1

VIBRATION FATIGUE TESTING OF SOCKET WELDS, PHASE II

Paul Hirschberg Peter C. Riccardella Structural Integrity Associates 3315 Almaden Expressway, Suite 24 San Jose, CA 95118-1557

Michael Sullivan Jon Schletz Pacific Gas & Electric Co. 3400 Crow Canyon Rd. San Ramon, CA 94583

Robert Carter EPRI Project Manager 1300 Harris Boulevard Charlotte, NC 28262

ABSTRACT This paper describes the results of the second phase of an EPRI sponsored program to perform high cycle fatigue testing of socket welds in order to quantify the effects of various factors upon the fatigue strength. Analytical results had demonstrated that the socket weld leg size configuration can have an important effect on its high cycle fatigue resistance, with longer legs along the pipe side of the weld greatly increasing its predicted fatigue resistance. Other potentially important factors influencing fatigue life include residual stress, weld root and toe condition, pipe size, axial and radial gaps, and materials of construction. The second phase of the program tested 27 additional socket weld specimens of various designs by bolting them to a vibration shaker table and shaking them near their resonant frequencies to produce the desired stress amplitudes and cycles. Another objective of the second phase of testing was to evaluate various methods of in-situ modification or repair of socket welds, which could be used as alternatives to replacement with butt welds. The results of the program are presented which include comparisons of the various socket weld designs with standard Code socket welds, butt welds, ASME mean failure data, and recent test data published by Higuchi et al. Fatigue Strength Reduction Factors are calculated based on the testing results. INTRODUCTION Failures of small bore piping connections continue to occur frequently in U.S. nuclear power plants, resulting in degraded plant systems and unscheduled plant downtime. Prior research [1] has indicated that the majority of such failures are caused by vibration fatigue of socket welds. In order to better understand and characterize this phenomenon, investigations have been performed in the U.S. [1,2,3] and overseas [4,5,6]. Analytical results reported in Reference [3] have demonstrated that the socket weld leg size configuration can have an important effect on its high cycle fatigue resistance, with longer legs along the pipe side of the weld greatly increasing its predicted fatigue resistance. Other factors which potentially influence socket weld fatigue life are residual stress, weld root and toe condition, loading mode, pipe size, axial and radial gaps, and materials of construction. The test program described in this paper was initiated in 1997 under EPRI sponsorship to study the importance of these factors. A large number of socket weld samples were vibration-fatigue tested to failure on a high frequency shaker table. The objectives of the testing were to improve the industrys understanding and characterization of the high cycle fatigue resistance of socket welds and to develop appropriate fatigue strength reduction factors for such welds reflecting the effects of those factors listed above which prove to be significant. The ultimate goal of this research is to develop recommended design and fabrication practices that can be used to enhance socket weld fatigue resistance in vibrationsensitive locations, as well as to provide guidelines for screening out and preventing vibration-fatigue failures in existing welds. The first phase [7], completed in 1998, investigated the variation in high cycle fatigue resistance as a function of weld leg length, pipe diameter, and piping material. The effect of an additional weld pass, post-weld heat treatment, and eliminating the ASME Code-required axial gap were also studied. The test setup and results have been described in a previous paper [8]. The second phase investigated remedial actions for existing socket welds, in order to determine their effectiveness in extending the useful life of socket welded joints, in lieu of making expensive modifications. The effect of varying toe conditions was

37-3

also studied in Phase II, and additional data at higher loads was collected to further quantify the benefit of increased weld leg length. This paper describes the second phase of the test program, and provides the cumulative results for both phases. BACKGROUND AND APPROACH The probability of fatigue failure at a weld is a function of many considerations: type of weld, weld profile, material, weld size, wall thickness, heat treatment, grinding after welding, etc. Socket welds are commonly used in small bore piping (2 and under) joints due to their ease of assembly, but their fatigue strength is generally considered to be less than that of a butt weld. The fatigue strength reduction factor (FSRF) is a means of quantifying the effect of a local structural discontinuity on the fatigue strength of the joint. It is defined as the fatigue strength of the component without the discontinuity divided by the fatigue strength with the discontinuity. The FSRF is usually determined by test, and is generally calculated by dividing the endurance limit of the ASME mean failure curve for polished bar specimens, by the endurance limit of the joint in question. The endurance limit is used because the FSRF varies with fatigue life; the value at the endurance limit is bounding. For practical reasons, the endurance limit is often defined as the fatigue strength at some large, predefined number of cycles, such as 2x107. Article NB-3600 of Section III, which provides rules for the design of Class 1 piping, accounts for the variation in fatigue strength between different welds and fittings by multiplying the calculated alternating stress by stress indices, before entering the fatigue curves. For primary plus secondary bending moment loading, the applicable stress index is the product of C2 and K2. C2K2 is approximately equivalent to the FSRF, although there are some differences in derivation relating to testing requirements and statistical treatment of the data. For socket welds (1998 Edition of the Code), C2 = 1.93 for a Code minimum weld according to the rules of the fabrication section (NB-4427), and the product C2K2 is 3.86, or about an FSRF of 4 for a socket weld. PARAMETERS TESTED While the ASME piping design Code provides stress indices for socket welds, these represent the predicted fatigue strength of standard weld designs. This test program included welds of the standard Code design as well as several variations in weld design. The following is an overview of the parameters that were tested. Weld Leg Size Analysis described in Reference 3, and testing reported in Reference 10, have identified variations in socket weld design that can significantly affect the fatigue life of the joint under high cycle loading. One of the most important factors is the weld leg size. Extending the weld leg length along the pipe side of the weld by a factor of two over the Code minimum of 1.09 tn (where tn is the nominal pipe wall thickness) has resulted in a significant improvement in fatigue resistance. In this testing program, the effect of weld size on fatigue resistance was studied by testing samples fabricated with oversized legs on the pipe side, and comparing them to control samples at nominal Code dimensions. Testing was also done to study how conventional small bore butt welded fittings compare to new 2 x 1 welds, and to 1 x 1 welds modified to 2 x 1 weld dimensions, under comparable loading conditions. The butt welded specimens consisted of pipes welded to standard weld neck flanges, which is similar to a typical method used in power plants for upgrading socket welded fittings when vibration problems have occurred. Weld Modification and Repair An important purpose of the testing program was to study repair or modification actions that could be taken to improve the fatigue resistance of an existing socket welded joint. Repair concepts for leaking socket welds that would allow plants to continue operating until the next outage before implementing a permanent repair were also tested. Weld overlay repairs were applied to four of the welds that developed leaks in the first phase of testing. The leaks were peened with water still in the pipe, and a seal weld was applied when the leak stopped. The overlay design applied sufficient additional reinforcement to cover

37-4

the possibility of either a toe or a root failure having occurred. The design is shown in Figure 3. The weld was applied using a shielded metal arc welding process to simulate the most likely welding process that would be used for a temporary repair in a power plant. The pipe side toe of the weld did not have a blending radius, producing a sharp transition, as would likely be found in a field repair. The repaired welds were tested at their original load levels to determine if they could survive until the next outage, or convenient time for a permanent repair. Specimens chosen for these tests included carbon and stainless steel welds that had experienced root and toe failures. Another remedial approach tested was to modify existing standard Code welds to a 2 x 1 leg length configuration after having been fatigue cycled in their original configuration but before any obvious cracking or fatigue damage was observed. This is illustrated in Figure 2. Testing compared the degree of improvement for the modified welds relative to new 1 x 1 and 2 x 1 welds. Weld Profile and Toe Condition The profile of the weld and the condition of the weld toe can have a significant effect on socket weld fatigue life. The ideal toe condition is for the weld to blend smoothly with the pipe with no discontinuity or undercut. Premature toe failures are generally the result of a discontinuity or small flaw in the toe, which propagates through the base metal. Grinding the weld can help by removing small surface imperfections, which could be sites for crack initiation. Conversely, heavy abusive grinding can worsen the situation by leaving a cold worked condition and tensile residual stresses. In the first phase of testing, several specimens exhibited toe failures prior to the expected number of cycles. When toe failures occurred, they were generally associated with minor welding discontinuities at the toe, which would normally be acceptable by Code workmanship standards. In the second phase, samples with polished toes, as-welded toes, and intentionally poor toes were tested to quantify the effect of the toe condition. One of the butt welds was also fabricated with a toe defect to see if this is a concern for butt welds. Residual Stress The residual stress in the weld can have an important effect on fatigue life in the high cycle regime. Residual stress acts as a mean stress, which reduces the allowable number of cycles at a particular alternating stress. This is more important in the high cycle region than in low cycle fatigue, where the effects of plasticity act to relieve the mean stress. The residual stress in a socket weld can be altered by varying the welding technique, such as by adding a final cover weld pass. Last Pass Improvement refers to a technique in which a normal Code socket weld is improved by adding a last pass on the pipe side of the weld, which changes the residual stress in the weld root to compressive and extends the leg length along the pipe somewhat. Tests were conducted to determine the amount of benefit gained by this process. Post-Weld Heat Treatment Post-weld heat treatment (PWHT) is another method of reducing the residual stress in a weld. The ASME Code requires PWHT for most ferritic steel welds, but exempts socket welds and austenitic steels. One of the reasons is that heating austenitic steels can sensitize them to intergranular stress corrosion cracking (IGSCC). However, in view of the importance of residual stresses in high cycle fatigue life, applying PWHT to socket welds in vibration sensitive locations that are not exposed to an IGSCC conducive environment can be a viable means of reducing residual stress. Axial Gap The ASME Code requires that an axial gap of 1/16 be provided between the pipe end and the socket. The reason is that if the pipe carries hot fluid, differential thermal expansion between the pipe and the fitting may add significant stress to the weld if there is no gap. Also, without the gap, shrinkage of the fillet weld could produce residual stresses in the weld, pipe, and fitting wall. However, some recent testing [4] has indicated that absence of the gap has a negligible effect on fatigue life.

37-5

Pipe Size The ASME Code design fatigue curves do not reflect any differences in fatigue strength as a function of pipe diameter. However, the Phase 1 testing showed that pipe had superior fatigue strength to the 2 pipe. Piping Material Although the endurance limit of carbon steel is lower than that of stainless, previous test data has indicated that some of the fatigue strength reduction factors for stainless steel are higher. While the majority of the samples tested in this program were stainless steel, one-third of the 2 specimens were carbon steel. The purpose was to note whether any of the modifications to socket weld design are more or less effective in carbon steel than in stainless. TESTING ARRANGEMENT The test setup was previously described in detail in [8]. The testing method used specimens vertically cantilevered on a shaker table. All specimens were fabricated from Schedule 80 piping and compatible components. Loading amplitudes were selected based on fatigue data available in the literature, with the target of generating failures in approximately 106 to 107 cycles. As a large number of runouts (tests ended without failure) were obtained in the first phase of testing, the stress levels in the second phase of testing were increased. Sets of nine socket weld specimens were bolted to the shaker table and shaken simultaneously near their resonant frequencies to produce the desired stress amplitudes and cycles. A cantilevered specimen of the type illustrated in Figure 1 was used, with the test weld being the weld at the lower end of the specimen between the pipe and the flange used to bolt the specimen to the table. Different load amplitudes were applied to different samples in the same test by fine tuning the specimen natural frequencies relative to the shaker table excitation frequency. The shaker table typically ran at 100-110 Hz and the test specimens had natural frequencies that were nominally 4-8 Hz lower than the excitation frequency. By adding or subtracting small masses, such as nuts and washers, the frequencies of the test specimens were moved enough off resonance to adjust each individual response acceleration. The flange configurations were modified to produce socket weld details typical of the socket welded fittings used on small bore piping in nuclear plants (tees, elbows, weldolets, couplings, etc.). The specimens were pressurized with air to a moderate pressure of approximately 50 psig. When depressurization occurred, indicating a failure, the specimen was removed from the table at the next convenient test stoppage, and the testing was then resumed with only the remaining specimens. This test method is directly comparable with the plant loading mechanism of concern (vibration fatigue), whereas conventional fatigue testing techniques (rotating beam or four point bending) can have considerable variability with respect to each other [5,6], and possibly with respect to in-plant vibration. The stress in the test sample was determined from the measured acceleration response and test frequency, using beam formulas superimposing the concentrated weight and distributed weight effects. Nominal pipe wall but actual lengths and weights were used. It was observed that when the crack formed, the natural frequency of the specimen reduced, causing the acceleration response and the resulting stress to decline (this usually affected only the last 3-5% of the cycles). The stress reported in the test results is an average over all of the cycles. TEST RESULTS Table 1 is a summary of the results of the Phase II tests. The Phase I results have been reported previously in [8]. However, the results of both phases of testing are plotted in Figures 4-8. Trend curves from socket weld fatigue testing reported in [4,5,6] are also shown on the plots, labeled Higuchi Curves. The ASME mean failure curves for polished bar specimens are also shown for comparison purposes. When specimens exhibited toe failures (solid points in the figures) they tended to fail somewhat prematurely, relative to the more common root failures.

37-6

Table 1 - Phase II Test Results Test Series 4 Repairs and Mods. - 2 SS and CS Specimen Sa(ksi) Nf Comments 1 Mod. to 2x1 SS 17.4 2.57E+07 Runout 2 Mod. to 2x1 SS 17.6 5.05E+06 Toe Failure 3 Prev. Runout SS 16.3 1.33E+07 Root Failure 4 Mod. to 2x1 CS 14.7 2.57E+07 Runout 5 Mod. to 2x1 CS 11.1 2.57E+07 Runout 6 Weld Repair CS 8.9 2.57E+07 Runout 7 Weld Repair CS 8.0 2.57E+07 Runout 8 Weld Repair SS 11.6 1.10E+07 Failure at Orig. Toe 9 Weld Repair SS 11.0 2.57E+07 Runout Test Series 5 Toe Conditions - 2 SS Specimen Sa(ksi) Nf Comments 1 Smooth Toe 24.2 1.26E+06 Toe Failure 2 Smooth Toe 23.6 1.04E+06 Toe Failure 3 Poor Toe 18.2 1.62E+06 Toe Failure 4 Poor Toe 20.5 3.80E+05 Toe Failure 5 Polished Toe 21.4 3.24E+05 Toe Failure 6 Polished Toe 22.5 9.10E+05 Root Failure 7 Smooth Last Pass 22.8 3.56E+05 Toe Failure 8 Smooth Last Pass 23.3 1.62E+06 Root Failure 9 Smooth Last Pass 22.0 1.96E+06 Root Failure Test Series 6 2x1 vs. Butt Weld - 2 Specimen Sa(ksi) Nf Comments 1 Butt Weld SS 23.0 1.22E+06 I. D. Root Failure 2 Butt Weld SS 22.3 6.61E+05 I. D. Root Failure 3 Butt Weld CS 15.3 2.19E+06 Toe Failure 4 Butt Weld CS 16.8 1.97E+06 Toe Failure 5 Butt SS Undercut 22.6 1.00E+06 Toe Failure 6 2x1 SS 24.0 1.04E+06 Toe Failure 7 - 2x1 SS 23.5 2.11E+07 Runout 8 2x1 CS 16.5 2.38E+06 Root Failure 9 2x1 CS 16.9 3.75E+06 Root Failure In general, failures at a lower number of cycles and/or higher stress tended to originate at the toe while the higher cycle / lower stress failures tended to occur at the root. The following sections describe in detail the test results for each of the parameters studied. Effect of Weld Size Figures 4 and 5 show the results of testing of standard Code size welds (1 x 1) and enhanced, 2 x 1 welds, which have a leg dimension along the pipe of twice that of the Code minimum leg dimension. Also shown for comparison are the test results for the butt welds. Figure 4 shows the 2 stainless steel data; Figure 5 shows the 2 carbon steel data. In general, the nominal Code dimension (1 x 1) specimens yielded data somewhat above the corresponding Higuchi Curve. The failures at the lower numbers of cycles were toe failures and those at higher cycles were root failures (open points in the figures). The two butt weld failures at low cycles in this figure originated in the inside diameter of the pipe, at the edge of the weld root.

37-7

The 2 x 1 specimens were significantly stronger than the standard welds. All exhibited runouts at the lower stress levels, even though tested at stress amplitudes 30% higher than those applied to the standard Code specimens. At the higher stress levels, in the stainless steel tests, one was a runout at 23.5 ksi, while the other failed at only 1 x 106 cycles, and 24 ksi. The latter specimen was a toe failure, however, and performed about the same as the standard Code specimens tested at this stress level, which also exhibited toe failures. In the carbon steel tests at higher stress levels, both of the 2 x 1 welds performed significantly better than the standard welds. The butt welds were tested for the purpose of comparison against the 2 x 1 welds. However, the butt welds did not perform as well as expected. In the stainless steel tests, they were about as strong as the standard Code socket welds, while in the carbon steel tests, they were somewhat better than the standard welds, but not quite as good as the 2 x 1 welds. The butt weld results can be explained by the fact that the tests were comparing a socket welded flange to a butt welded, weld neck flange, and not to a pipe-to-pipe butt weld. The weld neck flange acts as a tapered transition, and the ASME Code stress indices reflect a reduced fatigue strength for such a joint. For the as-welded standard socket weld (1998 Code), C2K2 = 3.86; for the as-welded butt welded transition, C2K2 = 3.78. For an as-welded pipe-to-pipe butt weld, C2K2 = 2.57. Thus the stress indices for the butt welded transition are only slightly better than that of a socket weld. Socket welds are rarely used in pipe-to-pipe connections but rather at elbows, branch connections, or other fittings which would result in thickness transitions even if replaced with butt welds. Therefore the practice of replacing socket welds with butt welds to fix vibration fatigue problems appears to be of little value, if the butt weld is to a fitting with a geometric transition. Two of the stainless steel butt welds failed by a crack that originated at the inside diameter of the pipe, at the edge of the weld root. The crack grew perpendicular to the pipe axis, through the weld metal. Further examination of the weld root indicated the presence of a number of discontinuities, which may have served as crack initiation sites. The other stainless steel butt weld had an undercut intentionally placed at the toe; although the specimen failed at the toe, the fatigue strength was equivalent to the other butt welds. It was concluded from these tests that the 2 x 1 welds offer a significant improvement in fatigue strength over standard Code welds. They also provide a greater strength improvement than replacement with butt welded fittings, in addition to easier fitup and construction. Toe Condition The results of testing the three different toe conditions are shown in Figure 6. The poor toe conditions produced failures at or below the Higuchi data curve. As expected, they were toe failures. It is interesting to note that the two polished toe specimens did not show any improvement over the smooth, as-welded pieces. One of the polished toes failed at the toe right on the Higuchi trend line, which was below all of the standard, as-welded samples. Examination of the fracture surface under magnification showed multiple crack initiation sites along the scratch marks generated by the polishing. It is concluded from these tests that while a poor toe condition definitely reduces the fatigue life, polishing the toe does not seem to improve it. Possibly using a finer grit sandpaper and polishing in the axial direction would have produced better results. Four of the five specimens with a final weld pass at the toe performed somewhat better than the standard welds, with one failing early near the Higuchi curve. Weld Process Enhancements Figure 7 compares the results of the weld process enhancements for 2 stainless socket welds. Figure 8 shows the same for the carbon steel tests. The enhancements were: adding a last pass at the weld toe on the pipe side to improve the residual stress; post-weld heat treatment of the weld; and eliminating the Code required axial gap. The testing results for these specimens were previously reported in [8]. The last pass improved specimens yielded somewhat mixed results. In general, where premature failures occurred in last pass improved specimens, they were due to toe failures, indicating that the last pass welding may have left a discontinuity or stress raiser at the toe. Three of the five last pass samples performed better than the standard socket welds. Post weld heat treatment appears to increase the fatigue life of the

37-8

standard Code specimens. One of the stainless steel runouts from Phase I was retested at a higher load in test series 4. Despite having withstood 23 million cycles at 12 ksi, it lasted another 13 million cycles at 16 ksi. The ASME Code required gap appears to have no effect on high cycle fatigue resistance. Modified and Repaired Welds Two stainless steel and two carbon steel 1x1 socket welds that had been tested to runout in the second or third test series had additional weld metal applied to them to make them 2x1 welds. The modified welds were then retested at loads that were approximately 50% higher than in the previous tests. The purpose was to determine whether an existing Code standard weld can be increased to a 2x1 weld in situ to improve its fatigue strength, and how it would compare against a new 2x1 weld. The results are shown on Figures 4 (stainless) and 5 (carbon steel). One of the stainless welds was a runout at 17.4 ksi. The other failed at the toe after 5 million cycles at 17.6 ksi. The first specimen was a former Code standard weld, and the second was a post-weld heat treated specimen. It should be noted that the latter weld was sectioned, and examinations indicated that the original weld did not have any indications of cracking prior to the buildup. Both of the carbon steel welds were runouts, the first being a former post-weld heat treated specimen tested at 14.7 ksi, and the second, a no-gap specimen tested at 11 ksi. The results clearly show an improvement in fatigue strength over standard Code welds. As for comparing the built up 2x1 welds with new 2x1 welds, most of the welds of both categories were runouts, so a definitive comparison was not possible. However, the built up welds appear to be at least as good as the new 2x1 welds. The results of the tests of the weld overlay repaired welds are shown on Figures 7 and 8. One of the stainless steel repaired welds lasted about 11 million cycles at 11.6 ksi. The original weld had been a lastpass specimen that had failed at the toe at 10 million cycles under the same load. This time, the weld failed not at the new toe, but at the continuation of the original weld crack. Although the overlay did not arrest the original toe crack, it was still successful in restoring all of the fatigue life of the original weld. The second stainless specimen had originally been a Code standard weld that had failed at the root at 10 million cycles at 10.5 ksi. In this test, it was a runout at a stress of 11 ksi. The weld overlay was successful in arresting the original crack. The repaired weld was thus better than the original. In the carbon steel tests, both repaired welds were runouts. They had both originally been standard Code welds that had failed at the root after 7 million and 10 million cycles at 8 ksi. The repaired welds withstood 26 million cycles at 8 and 9 ksi respectively and did not show any evidence of crack initiation. It can be concluded from these tests that the weld overlay repair process was not only successful in restoring the original fatigue strength of the specimen, but it actually improved the welds fatigue resistance. The weld overlay design used was more effective in arresting root cracks than toe cracks. Fatigue Strength Reduction Factors Fatigue strength reduction factors (FSRFs) were calculated as a means for quantifying the test results of the various socket weld designs. The FSRFs are approximate, as they are based on estimates of the endurance limit, a number of tests that were runouts, and a limited number of test points. Table 2 is a summary of the FSRFs for each of the categories of tests, based on the ratio of the endurance limit from the ASME Code mean failure curve to the apparent endurance limit from the testing results. Since the testing was terminated at approximately 2 x 107 cycles, the alternating stress corresponding to this number of cycles was taken as the endurance limit for the purposes of this calculation. For the 2 stainless steel specimens, the standard Code socket welds had an FSRF of approximately 3.4. The 2 x 1 leg size specimens improved upon this to a value of 2.3, versus 2.9 for the butt welded specimens. Similarly, for the carbon steel specimens, the standard weld FSRF was 3.5. The 2 x 1 welds reduced this value to 2.0, versus 2.4 for the butt welds. An apparent break in this trend is that, as observed previously in Reference 5, standard Code specimens showed less of an improvement in going to the 2 x 1 weld geometry. This is due to the fact that the standard welds in this pipe size have a higher ratio of weld section modulus to pipe section modulus than in 2 pipe and are consequently stronger. It appears that it is reasonable to use an FSRF of 3.9 for standard Code socket welds, and an FSRF of 2.6 for 2 x 1 leg size socket welds in vibration fatigue applications, independent of pipe size. This is consistent with the

37-9

current ASME Section III Code requirement of C2K2 = 3.9 for standard socket welds and the lower bound of 2.6 for larger leg length welds. This applies only to welds with no root defects or lack of fusion. Testing of the samples with toe defects in this program resulted in an FSRF of about 5. The EPRI Handbook recommends a FSRF of 8.0 for Poor welds, which is intended to encompass welds with root or toe defects. A comparison of the various weld treatments indicates that post-weld heat treatment had the most consistent benefit, with an FSRF of 2.5 for 2 stainless, 2.5 for carbon steel, and 2.6 for stainless. The last pass welds had FSRFs of about 3, and the no-gap welds about 3.5. The modification of the previously tested 1 x 1 welds, by building up the weld on the pipe side to a 2 x 1 profile, produced welds that were as good as or better than new 2 x 1 welds. For stainless steel, the FSRF for the built-up 2 x 1 welds was 2.2, versus 2.3 for new 2 x 1 welds and 3.4 for new 1 x 1 welds. For carbon steel, the FSRFs were 1.8 for the built-up 2 x 1, 2.0 for the new 2 x 1, and 3.5 for the Code standard welds. The weld overlay repairs were also successful in improving the fatigue strength of leaking, standard Code welds. The stainless steel welds that were repaired had an FSRF of 3.2, which is better than new standard welds, at 3.4. The results were even better in the carbon steel specimens, with an FSRF of 2.8 for the repaired welds versus 3.5 for new Code welds. Table 2 - Fatigue Strength Reduction Factors Category 2 Stainless Steel Tests Code Standard (1 x 1) 2x1 1 x 1 Built-up to 2 x 1 Butt Weld Weld Overlay Repair PWHT Last Pass No Gap Polished Toe Smooth As-Welded Toe Smooth Last Pass Toe Poor Toe 2 Carbon Steel Tests Code Standard (1 x 1) 2x1 1 x 1 Built-up to 2 x 1 Butt Weld Weld Overlay Repair PWHT No Gap Stainless Steel Tests Code Standard (1 x 1) 2x1 PWHT Last Pass No Gap No. Tested 2 5 2 3 2 2 2 2 2 2 3 2 2 5 2 2 2 1 1 2 3 1 2 1 Avg. FSRF 3.4 2.3 2.2 2.9 3.2 2.5 3.1 3.7 3.6 2.8 2.9 4.8 3.5 2.0 1.8 2.4 2.8 2.5 3.1 2.4 2.2 2.6 2.3 2.4

37-10

CONCLUSIONS AND RECOMMENDATIONS On the basis of the testing, it is concluded that socket welds with a 2 to 1 weld leg configuration (weld leg along the pipe side of the weld equal to twice the Code required weld leg dimension) offer a significant high cycle fatigue improvement over standard ASME Code socket welds. This weld design offers a superior improvement in fatigue resistance than does replacement of socket welded fittings with butt-welded fittings. Since vibration fatigue of socket welds has been a significant industry problem, it is recommended that this improved configuration be used for socket welds in vibration critical locations. The majority of the test failures occurred due to cracks that initiated at weld roots. However, toe initiated failures occurred in tests at higher stress levels that were premature in comparison with identical tests in which root failures prevailed. Therefore, care must be taken with socket welds of any design to avoid metallurgical or geometric discontinuities at the toes of the welds (such as undercut or non-smooth transitions). Such discontinuities promote a tendency for toe failures which greatly reduces fatigue life. Tests of welds with intentionally poor toes clearly demonstrated the early failures that such discontinuities can produce. Polishing the toes did not provide any benefit, causing cracking originating at scratch marks. Because of the importance of the toe condition, the last pass improvement process (in which a final pass is added to the pipe side toe of a standard Code weld) cannot be given an unqualified recommendation at this time. The last pass improved specimens had a tendency to develop toe failures. Other conclusions drawn from this program are that the code required axial gap in socket welds (1/16) appears to have little or no effect on high cycle vibration fatigue resistance (thermal expansion effects were not part of the test), and that post weld heat treatment appears to increase the fatigue resistance of standard Code specimens. Although post-weld heat treatment consistently showed improved results, it has the downside of potentially sensitizing austenitic welds for IGSCC in certain environments. Therefore, PWHT is not recommended for situations where IGSCC may be a damage mechanism. The test data supports the use of the ASME Section III stress indices for standard Code welds (currently C2K2 = 3.9). It also indicates that this factor may be reduced to two-thirds that value (2.6) for 2 to 1 leg size welds. (The current Code lower bound of C2K2 = 2.6 is based on weld leg length, but it is a function of the shortest leg length.) Both of these values are appropriate only if the weld roots are free of defects such as lack of fusion or lack of penetration. Testing of modification and repair concepts indicated that these approaches were successful in improving the strength of already installed socket welds without having to replace them with butt welds. Code standard 1x1 welds that were built up to a 2x1 profile performed as well as new 2x1 welds. Weld overlay repairs of leaking standard welds not only provided enough fatigue resistance for the welds to last to the next outage, but actually improved their fatigue strength to better than new standard Code welds. The weld overlay process was somewhat more successful repairing root failures than toe failures. REFERENCES 1. EPRI 1994, EPRI Fatigue Management Handbook, EPRI TR-104534-V1, -V2, -V3. 2. Smith, J. K., 1996, Vibrational Fatigue Failures in Short Cantilevered Piping with Socket-Welding Fittings, ASME PVP 338-1. 3. EPRI 1997, Vibration Fatigue of Small Bore Socket-Welded Pipe Joints, EPRI TR-107455. 4. Higuchi, M. et al, 1995, Fatigue Strength of Socket Welded Pipe Joints, ASME PVP 313. 5. Higuchi, M. et al, 1996, A study on Fatigue Strength Reduction Factor for Small Diameter Socket Welded Pipe Joints, ASME PVP 338-1. 6. Higuchi, M. et al, 1996, Effects of Weld Defects at Root on Rotating Bending Fatigue Strength of Small Diameter Socket Welded Pipe Joints, ASME PVP 338-1. 7. EPRI 1998, Vibration Fatigue Testing of Socket Welds, EPRI TR-111188. 8. Riccardella, P. C. et al, 1998, Vibration Fatigue Testing of Socket Welds, ASME PVP 360. 9. ASME Boiler and Pressure Vessel Code, Section III, Subsection NB, 1998 Edition. 10. WRC Bulletin 432, 1998, Fatigue Strength Reduction Factors and Stress Concentration Factors for Welds in Pressure Vessels and Piping, Welding Research Council.

37-11

Figure 1. Test Setup

Figure 2. 1 x 1 Weld Built Up to 2 x 1

37-12

Figure 3. Weld Overlay Design

EPRI SOCKET WELD VIBRATION TESTS 2" Stainless Steel Welds 2x1 vs. Code Standard Welds
100

Solid Symbols are Toe Failures

Runout

Sa(ksi)

10

Higuchi Curve ASME Mean Failure Code 1 x1 2x1 Welds 1x1 Built up to 2x1 Butt Welds

1 1.00E+05

1.00E+06 N(Cycles)

1.00E+07

1.00E+08

Figure 4. Weld Size, Stainless Steel

37-13

EPRI SOCKET WELD VIBRATION TESTS 2" Carbon Steel Welds 2x1 vs. Code Standard Welds
Solid Symbols are Toe Failures 100.00

Runouts

Sa (ksi)

10.00

Higuchi Curve ASME Mean Failure Code 1x1 2x1 Welds 1x1 Built up to 2x1 Butt Welds

1.00 1.00E+05

1.00E+06 N (Cycles)

1.00E+07

1.00E+08

Figure 5. Weld Size, Carbon Steel

EPRI SOCKET WELD VIBRATION TESTS 2" Stainless Steel Welds Toe Conditions
100

Solid Symbols are Toe Failures

Runout

Sa(ksi)

10

Higuchi Curve ASME Mean Failure Code Smooth As Welded Polished Toe Poor Toe Last Pass

1 1.00E+05

1.00E+06 N(Cycles)

1.00E+07

1.00E+08

Figure 6. Toe Conditions

37-14

EPRI SOCKET WELD VIBRATION TESTS 2" Stainless Steel Welds Weld Process Enhancements
100

Solid Symbols are Toe Failures

Runout

Sa(ksi)

10

Higuchi Curve ASME Mean Failure Code Standard PWHT Retested PWHT Runout No Gap Last Pass Weld Overlay Repaired

1 1.00E+05

1.00E+06 N(Cycles)

1.00E+07

1.00E+08

Figure 7. Weld Enhancements, Stainless Steel

EPRI SOCKET WELD VIBRATION TESTS 2" Carbon Steel Welds Weld Process Enhancements
Solid Symbols are Toe Failures 100.00

Sa (ksi)

Runouts

10.00

Higuchi Curve ASME Mean Failure Code Standard PWHT No Gap Weld Overlay Repaired

1.00 1.00E+05

1.00E+06 N (Cycles)

1.00E+07

1.00E+08

Figure 8. Weld Enhancements, Carbon Steel

37-15

VIBRATION FATIGUE TESTING OF SOCKET WELDS, Phase II


Paul Hirschberg, Pete Riccardella Structural Integrity Associates Mike Sullivan, Jon Schletz Pacific Gas & Electric Co. Robert Carter EPRI July 2000

37-16

Background
Leaks of small bore socket welds due to vibration are a common problem in nuclear plants Can cause unscheduled outages, resulting in significant cost impact EPRI sponsored a research program to test improvements in socket weld design Phase I completed 1998, Phase II in 1999 (this paper)

37-17

Purpose
Test different socket weld designs on a high frequency shaker table Determine how variations in socket weld design affect fatigue life Develop fatigue strength reduction factors Develop guidelines for avoiding fatigue failures in vibration sensitive locations Test in-situ repair options

37-18

Scope of Testing - Phase I


Increased leg length (2 x 1) 2" and 3/4" NPS Carbon steel and stainless steel Additional weld pass at toe Post weld heat treatment Eliminating axial gap

37-19

Scope of Testing - Phase II


Increased leg length (2 x 1) at higher loads
Compared against standard Code welds Compared against butt welds Carbon steel and stainless steel

Toe condition
Smooth / as-welded toe Polished toe Poor toe condition Smooth additional weld pass

37-20

Scope of Testing - Phase II, contd


Modified welds
Increase previously cycled standard welds to 2 x 1 Carbon steel and stainless steel

Repaired welds
Repaired leaking welds with weld overlay Carbon steel and stainless steel

All phase II tests were 2" NPS Sch 80

37-21

Test Setup
Specimens vertically cantilevered on shake table
Similar to vents and drains

Amplitude of vibration set to cause failure in 10E6 -10E7 cycles Nine specimens tested at a time Individual load amplitude varied by tuning specimen natural frequency vs. table frequency Specimens pressurized with air; failure identified as loss of air pressure (through-wall leak) Applied stress determined from beam formulas using measured accelerations and frequencies, specimen length and weight
37-22

Signal from Accelerometer

Pressure Gauge

Direction Of Oscillations Mass (M)

SCH 80 Pipe

Note: M & L will be adjusted to yield desired natural frequency (~ 100 Hz)

Socket Welding Flange Modified from Standard Flange

Test Weld

Custom Drilled Plate on Shaker Table


98039r0

Test Specimen Detail


37-23

Test Apparatus
(Six of Nine Specimens Shown - Test Control Computer on Right)
37-24

1 x 1 Weld Built up to 2 x 1

37-25

Weld Overlay Design

37-26

Weld Overlay Repair (B2-2SS-2)

37-27

Results - Weld Size


2 x 1 welds much stronger than standard Code welds 2 x 1 welds stronger than butt welds Root failures at high cycles, lower amplitude; toe failures at lower cycles, higher amplitude Two butt welds failed at low cycles due to ID crack along side weld root

37-28

EPRI SOCKET WELD VIBRATION TESTS 2" Stainless Steel Welds 2x1 vs. Code Standard Welds
100

Solid Symbols are Toe Failures

Runout

Sa(ksi)

10

Higuchi Curve ASME Mean Failure Code 1 x1 2x1 Welds 1x1 Built up to 2x1 Butt Welds

1 1.00E+05

1.00E+06 N(Cycles)

1.00E+07

1.00E+08

2" (5.08 cm) Stainless Steel Weld Sizes


37-29

EPRI SOCKET WELD VIBRATION TESTS 2" Carbon Steel Welds 2x1 vs. Code Standard Welds
Solid Symbols are Toe Failures 100.00

Runouts

Sa (ksi)

10.00

Higuchi Curve ASME Mean Failure Code 1x1 2x1 Welds 1x1 Built up to 2x1 Butt Welds

1.00 1.00E+05

1.00E+06 N (Cycles)

1.00E+07

1.00E+08

2" (5.08 cm) Carbon Steel Weld Sizes


37-30

Butt Weld Failure

37-31

Results - Toe Condition


Poor toe condition resulted in premature toe failures Three additional weld pass specimens lasted longer than standard welds, two failed earlier at the toe Polished toe specimens failed earlier than standard welds

37-32

EPRI SOCKET WELD VIBRATION TESTS 2" Stainless Steel Welds Toe Conditions
100

Solid Symbols are Toe Failures

Runout

Sa(ksi)

10

Higuchi Curve ASME Mean Failure Code Smooth As Welded Polished Toe Poor Toe Last Pass

1 1.00E+05

1.00E+06 N(Cycles)

1.00E+07

1.00E+08

Toe Conditions

37-33

Polished Toe Failure

37-34

Weld Process Enhancements


Additional weld pass improves fatigue life as long as toe blends smoothly Post weld heat treatment improves fatigue life significantly; may cause sensitization or heat treatment may be difficult to achieve Eliminating axial gap has no consistent effect temperature effects not tested

37-35

EPRI SOCKET WELD VIBRATION TESTS 2" Stainless Steel Welds Weld Process Enhancements
100

Solid Symbols are Toe Failures

Runout

Sa(ksi)

10

Higuchi Curve ASME Mean Failure Code Standard PWHT Retested PWHT Runout No Gap Last Pass Weld Overlay Repaired

1 1.00E+05

1.00E+06 N(Cycles)

1.00E+07

1.00E+08

2" (5.08 cm) Stainless Steel Weld Enhancements


37-36

EPRI SOCKET WELD VIBRATION TESTS 2" Carbon Steel Welds Weld Process Enhancements
Solid Symbols are Toe Failures 100.00

Sa (ksi)

Runouts

10.00

Higuchi Curve ASME Mean Failure Code Standard PWHT No Gap Weld Overlay Repaired

1.00 1.00E+05

1.00E+06 N (Cycles)

1.00E+07

1.00E+08

2" (5.08 cm) Carbon Steel Weld Enhancements


37-37

Size Comparison
Code standard 3/4" socket welds have longer fatigue life than standard 2" welds Improved weld configurations show less of an improvement in the 3/4" size

37-38

Modified and Repaired Welds


Three of four previously cycled socket welds that were built up to 2 x 1 would not fail Built up welds clearly superior to new standard welds, as good as new 2 x 1 welds Weld overlay repair at a minimum restored full life of weld; three of four tests did not produce failure Weld overlay repair more successful arresting root cracks; toe crack continued to propagate

37-39

Weld Overlay of Root Crack

37-40

Weld Overlay of Toe Crack

37-41

Fatigue Strength Reduction Factors


Method to estimate relative benefit of various weld design improvements Calculated by dividing endurance limit stress of polished bar specimen by endurance limit stress of weld Roughly similar to ASME Code stress indices C 2K 2

37-42

Fatigue Strength Reduction Factors


Category 2 Stainless Steel Tests Code Standard (1 x 1) 2x1 1 x 1 Built-up to 2 x 1 Butt Weld Weld Overlay Repair PWHT Last Pass No Gap Polished Toe Smooth As-Welded Toe Smooth Last Pass Toe Poor Toe 2 Carbon Steel Tests Code Standard (1 x 1) 2x1 1 x 1 Built-up to 2 x 1 Butt Weld Weld Overlay Repair PWHT No Gap Stainless Steel Tests Code Standard (1 x 1) 2x1 PWHT Last Pass No Gap 37-43 2 3 1 2 1 2.4 2.2 2.6 2.3 2.4 2 5 2 2 2 1 1 3.5 2.0 1.8 2.4 2.8 2.5 3.1 2 5 2 3 2 2 2 2 2 2 3 2 3.4 2.3 2.2 2.9 3.2 2.5 3.1 3.7 3.6 2.8 2.9 4.8 No. Tested Avg. FSRF

Conclusions
2 x 1 welds offer significant improvement in fatigue strength over Code standard welds Replacement of socket welds with butt welds provides little improvement Important to avoid discontinuities in weld toe to prolong fatigue life Polishing the weld toe did not provide any benefit Additional pass provided benefit but was subject to toe initiated cracks Post weld heat treatment helps, but may sensitize austenitic material

37-44

Conclusions, contd
Eliminating axial gap had no effect Previously cycled standard welds built up to 2 x 1 were as good as new 2 x 1 welds Weld overlay repaired welds were better than new standard welds Weld overlay repairs were more successful with root cracks than toe cracks

37-45

You might also like