Anda di halaman 1dari 8

Proceedings of the 2012 9th International Pipeline Conference IPC2012 September 24-28, 2012, Calgary, Alberta, Canada

IPC2012-90518

HOW NEW VINTAGE LINE-PIPE STEEL FRACTURE PROPERTIES DIFFER FROM OLD VINTAGE LINE-PIPE STEELS
G. Wilkowski, D-J. Shim, Y. Hioe, S. Kalyanam, and F. Brust Engineering Mechanics Corporation of Columbus 3518 Riverside Drive - Suite 202 Columbus, OH 43221 U.S.A.

ABSTRACT Newer vintage line-pipe steels, even for lower grades (i.e., X60 to X70) have much different fracture behavior than older line-pipe steels. These differences significantly affect the fracture control aspects for both brittle fracture and ductile fracture of new pipelines. Perhaps one of the most significant effects is with brittle fracture control for new line-pipe steels. From past work brittle fracture control was achieved through the specification of the drop-weight-tear test (DWTT) in API 5L3. With the very high Charpy energy materials that are being made today, brittle fracture will not easily initiate from the pressed notch of the standard DWTT specimen, whereas for older line-pipe steels that was the normal behavior. This behavior is now referred to as Abnormal Fracture Appearance (AFA). More recent work shows a more disturbing trend that one can get 100-percent shear area in the standard pressed-notch DWTT specimen, but the material is really susceptible to brittle fracture. This is a related phenomenon due to the high fracture initiation energy in the standard DWTT specimen that we call Abnormal Fracture Behavior (AFB). This paper discusses modified DWTT procedures and some full-scale results. The differences in the actual behavior versus the standard DWTT can be significant. Modifications to the API 5L3 test procedure are needed. The second aspect deals with empirical fracture control for unstable ductile fractures based on older line-pipe steel tests initially from tests 30-years ago. As higher-grade line-pipe steels have been developed, a few additional full-scale burst tests have shown that correction factors on the Charpy energy values are needed as the grade increases. Those correction factors from the newer burst tests were subsequently found to be related to relationship of the Charpy energy values to the DWTT energy values, where the DWTT has better similitude than the Charpy test for fracture behavior (other than the transition 1

temperature issue noted above). Once on the upper-shelf, recent data suggest that what was once thought to be a grade correction factor may really be due to steel manufacturing process changes with time that affect even new low-grade steels. Correction factors comparable to that for X100 steels have been indicated to be needed for even X65 grade steels. Hence the past empirical equations in Codes and Standards like B31.8 will significantly under-predict the actual values needed for most new line-pipe steels. BRITTLE FRACTURE CONTROL CONSIDERATIONS The concern of brittle fracture control in gas pipelines was one of the main motivators for the start of the PRCI (originally the AGA NG-18 project) going back to the early 1960s. Brittle fractures had occurred in pipelines for distances of up to 10 miles in several cases in the late 1940s and early 1950s. Brittle fracture control in line-pipe steels was addressed in detail in an NG-18 (PRCI) report by W. A. Maxey using results from fullscale and vessel burst tests for materials made prior to 1970 [1]. The pipes in that study were early vintage thin-walled pipes having low Charpy upper-shelf energy values. These NG-18 efforts involved the development of the Drop-Weight Tear Test (DWTT) with validation from 99 pipes in full-scale burst tests that was adopted into API, ASTM, ISO and other international standards. It was a simple, inexpensive, and reliable test at that time. In 1975, a Canadian Arctic Gas Study Limited (CAGSL) research project was conducted to determine the fracture initiation and propagation behavior of experimental arctic pipes. The propagation behavior work involved conducting DWTT and West Jefferson burst tests (short vessels partially filled with gas) [2]. For one of the materials used in the pipe test, the standard pressed-notch (PN) DWTT results predicted that there would be ductile fracture at -30F or warmer temperatures. Copyright 2012 by ASME

However, when the 48-inch diameter by 1.25-inch thick X65 pipe test was conducted at -30F, the pipe shattered into many pieces with 5 to 10 percent shear area (SA), see Figure 1. In reexamining the 3 tested DWTT specimens, see Figure 2, it was found that two of them did not break, and the third one had an abnormal fracture appearance. The abnormal fracture appearance in the broken specimen consisted of 100% SA from the start of the pressed notch for about 60% of the fracture ligament, and then a virtually instantaneous transition to ~10% SA for the remaining part of the fracture. Only the back half of the specimen had the same SA percent as in the full-scale test. When eliminating the back and front fracture area for one wall thickness, the resulting API SA% rated values from the PNDWTT was 80% while it was believed that the other two specimens effectively had 100% SA even though they did not fully break (started with 100% shear area). This material had a very high Charpy upper-shelf energy [292J (215 ft-lb)] compared to the other materials at that time. This result caused quite a concern back in the mid-1970s, and these data were used for follow-on NG-18 (currently PRCI) efforts [3,4]. The Japanese pipe manufacturers proposed various modified notch DWTT specimens (i.e., precompressed notches, TIG spot welds, fatigue precracked notches, chevron shaped notches, etc.), some example references are [5] and [6]. For the NG-18/Battelle efforts, Wilkowski proposed deposited brittle weld metal notch specimens and another one called a staticprecracked notch. The brittle weld notch specimen was undependable since, in some cases, the small defects in the fusion-line region caused the entire weld to pop out of the specimen during impact testing. Additionally, it was considered too costly to make the welds for every specimen. Hence, the static-precracked notch was an alternative after studying all the other proposed notches. Some additional pipe tests were conducted, and although from the NG-18 work it was suggested that the brittle deposited weld or static-precracked (SPC) DWTT specimen was the most appropriate procedure, the API committee decided to use a chevron-shaped notch if abnormal fracture behavior occurred. The chevron notch is shown in the upper part of Figure 3. At that time, one of the concerns of the static-precracked notch was that because the specimen was quasi-statically preloaded to just past maximum load prior to the test to eliminate the plastic initiation energy, the specimen might be overly deformed after the static-precracking , see Figure 4(a). Instrumented data with high-speed video were not available during impact testing at that time, so it was not realized that the load-displacement data from a SPC-DWTT specimen matches the PN-DWTT (if shifted to account for the plastic initiation energy), see Figure 4(b) (taken from [7]). Hence, that deformation is part of the natural testing deformation behavior even under impact loading with the PN-DWTT. Recent Trends with New Line-Pipe Steels Back in 1975, the production of high Charpy upper-shelf energy materials was obtained by quenched-and-tempered steels 2

that were not frequently used for line-pipe applications. Currently, most pipe mills have improved their steel-making capabilities so that they can routinely make steel with Charpy upper-shelf energy values of 300 J (220 ft-lb) and many times up to 500 J (369 ft-lb). This is possible by making the steels much cleaner (lower inclusion content) than the older vintage steels. As a result, the initiation toughness increases considerably, but occurrence of abnormal fracture behavior in the PN-DWTT is becoming more common. Also, since 1975, there have been some data where the chevron-notch DWTT specimen actually gave a lower transition temperature than the PN-DWTT, which is the opposite trend than intended, see Figure 5. The basic problem with the chevron-notch DWTT specimen is that although it reduced the required total energy, the crack did not initiate in a brittle manner from the notch as in the pressed notch for the older line-pipe steels. This stems from the fact that as the material thickness decreases, the brittle-to-ductile transition temperature decreases; and at the point of the chevron-notch the thickness reduces to zero. Hence brittle fracture initiation that was the prerequisite for measuring the material arrest ability in the PN-DWTT specimen for older line-pipe steels cannot be achieved with the chevron-notch, see bottom photo in Figure 3. Because of this and many other examples, a working group in API is considering the elimination of the chevron-notch from the 5L3 standard. Figure 6 shows some additional samples of abnormal fracture behavior in DWTT specimens. This abnormal fracture behavior is now occurring in controlled-rolled steels as well as quenched-and-tempered steels. Abnormal fracture behavior occurs not only in PN-DWTT specimens, but also in Charpy V-notch and dynamic tear tests, see Figure 7. (Note the PN-DWTT specimen at the bottom of Figure 7 was from the same pipe shown in Figure 1, and again showed abnormal fracture behavior, but with a different length of ductile tearing prior to transitioning to the ~10% shear area.) Current experience shows that the length of ductile fracture before tripping into cleavage fracture is highly variable. Recent results from work conducted at Emc2 on examining variability in PN-DWTT tests in high Charpy energy line-pipe steel raised another concern, when one specimen had 100% SA but another PN-DWTT had ~20% shear area. Two DWTT specimens were reconstituted from each of the original specimens (one original specimen had 20% shear area and the other 100% shear area) and one of the reconstituted specimens was tested with a PN-DWTT and the other reconstituted specimen with the SPC-DWTT for each original specimen, see Figure 8. These reconstituted specimens showed consistent uniform shear lips in all SPC-DWTT specimens, but inconsistent behavior in the PN-DWTT specimens. So it is possible to have 100% SA in some PN-DWTT specimens for high Charpy energy materials, but the real fracture mode would be a transitional or brittle fracture. An additional aspect of interest is that over the last 4 years, Emc2 has found that it is possible to modify the DWTT specimen with a back-slot in a manner that gives upper-shelf Copyright 2012 by ASME

Load, kN

fracture speeds more typical of full-scale behavior [8,9]. The typical PN-DWTT specimen exhibits ductile (100% shear) fracture speeds of 30 to 40 ft/sec, while the modified backslotted (MBS) DWTT specimen (5-inch width rather than 3-inch width) has ductile fracture speeds up to 400 ft/sec which is comparable to the lower-bound ductile fracture speeds in fullscale pipe burst tests, see Figure 9. A static-precracked notch is also used in the MBS-DWTT specimen to eliminate the plastic energy required for crack initiation during the impact test. This specimen is now being used in the transition temperature region and is giving consistent results with the SPC-DWTT specimen for a number of materials. Additionally, the back-slotted region eliminates fracture in the material region where there has been compressive plastic strains from bending in the specimen shown in Figure 6d. Ideally a modified DWTT specimen like this might give more consistent shear lip formation for the entire fracture path compared to the very transient and inconsistent fracture behavior in the current API PN-DWTT specimen. The success of the SPC and MBS-DWTT specimens may be due to the fact that they are getting cleavage fracture from the precracked notches in the newer tougher steels, which was a key aspect in the success of the original PN-DWTT on the older lower-toughness line-pipe steels. In fact, all of the graphs and charts in API 5L3 show how to rate the shear area percent with cleavage fracture starting from the pressed notch. There are additional proprietary full-scale pipe burst tests that also show the PN-DWTT specimen is non-conservative in predicting the transition temperature by -25C for the high Charpy energy steel pipe used in that case. However, a good study with pipe burst tests and various alternative DWTT specimens is still needed for modification of API 5L3 for pipes having various thicknesses and Charpy energy level steels to make sure the new alternative specimen provides a general solution. Finally it should be noted that the Brittle Fracture Arrest in Pipeline criterion in PRCI report L51436 [1] needs to be modified for these materials. Even the Charpy specimen is affected by abnormal fracture behavior and the large plastic initiation energy in the Charpy specimen of the newer steels needs to be accounted for in the dynamic crack propagation behavior.

Figure 2 DWTT tests from pipe used in burst test in Figure 1

Figure 3 Chevron notch used as alternative to pressed notch in API 5L3


200 180 160 140 120 100 80 60 40 20 0 0 2 4 6 8 10 Displacement, mm 1.25% load drop

Figure 1 Results of 1975 burst test on 48-inch diameter by 1.25-inch thick X65 pipe at -30F

(a) Static precracking load-displacement record done prior to impact testing

Copyright 2012 by ASME

200 150 100 50

PN-DWTTs

Load, kN

SPC-DWTT

0 0 20 40 60 80 Displacement, mm

(c) 2010 thick-walled line-pipe steel (c) 2010 thick-walled linepipe material

(b) Impact testing records Figure 4 Instrumented DWTT data at a temperature where upper-shelf behavior occurred and showing that the SPC-DWTT curve can be shifted to give the same propagation behavior as in the PN-DWTT
100
CN-DWTT

80

PN-DWTT

SA%

60 40 20 0 -40 -30 -20 -10 0

(d)(d) 2008 Line-pipe material 2008 Linepipe material Figure 6 Several examples of abnormal fracture behavior in PN-DWTT specimens

Temperature, C

Figure 5 Comparison of chevron-notch and standard API pressed-notch DWTT results

(a) A different 1973 high Charpy upper-shelf material that had abnormal fracture in DWTT specimens even when machined down.

Figure 7 Abnormal fracture behavior in Charpy, dynamic tear tests, and PN-DWTT specimens from same pipe steel having Charpy upper-shelf of 220 ft-lb (298J) same pipe as in Figures 1 and 2.

(b) 2009 X65 linepipe material

(b) 2009 X65 line pipe material

Copyright 2012 by ASME

Figure 8 Reconstituted PN and SPC DWTT samples from original PN-DWTT specimens. All tests from the same pipe and at the same temperature.

Figure 9 Comparison of standard PN-DWTT and MBSDWTT specimens (back-slot depth is variable) DIFFERENCES IN UPPER-SHELF TOUGHNESS High energy pipelines like natural gas or liquid CO2 pipelines are susceptible to unstable ductile fracture under certain conditions. Arrest of a ductile propagating fracture has long been determined by the Battelle Two-Curve (BTC) analysis method [10]. The analysis procedure was first developed in the early 1970s and X52, X60 and X65 steels were used in the initial development of the BTC method. As time progressed, full-scale burst tests were conducted on newer grades of steels. X70, X80, X100 and X120 grade steel burst tests have been conducted over subsequent decades. From these burst tests, it has been empirically determined that as the grade level increased there needed to be some correction factor applied to the Charpy energy value. It was generally determined that a correction factor of 1.4 was needed for X80 pipes. Later, burst tests on

X100 pipe showed that a correction factor of 1.7 was needed for the X100 pipe material in those tests. In 2006, Wilkowski published a paper [11] that binned all the pipe in full-scale burst tests by grade level and statistically determined the correction factor to predict full-scale arrest needed on the Charpy energy for the BTC analyses. In addition, some proposed correction factors on Charpy energy values higher than 100J (Leis [12] and Wilkowski [13] relationship) were also examined. It was not surprising that the correction factors were different for; the original BTC, the Leis correction to the BTC, and the Wilkowski correction to the BTC; and all others increased as the grade levels increased, see Figure 10. The US DOT is currently considering use of the grade-correction factor from Ref. [11] and also in Ref [14] be used for pipeline design submittals. What was surprising from Ref. [11] was that when the DWTT energy was used instead of the Charpy energy, no grade correction factors were needed for the DWTT energy as the grade level changed for the pipes in the full-scale burst test database. Apparently, the Charpy specimen impact energy was changing relative to the full-thickness DWTT specimen. Figure 11 shows how the DWTT-to-Charpy energy values changed by grade level for the pipes in the full-scale database. The DWTT specimen is a much larger specimen for dynamic fracture behavior, where the specimen has enough of a fracture path to develop a steady-state fracture behavior [15] and also agrees in appearance with pipe burst test fracture behavior [16] (except for the newer very high Charpy energy steels discussed earlier with abnormal fracture behavior), so in hind sight this result should not have been surprising. Additionally in Ref. [11], it was found that there was a correlation between the statistical BTC correction factor needed for the Charpy arrest energy and the slope of the Charpy to DWTT energy/area (E/A) values. The low-grade steels involved in the past burst tests behaved similarly to work in Reference [17] published in 1977. Recent Trends with New Low-Strength Steels In the last few years, Emc2 has been examining recent vintage X60 to X70 steels for various pipeline projects. In plotting the DWTT and Charpy upper-shelf energy/area (E/A) values, surprisingly, the majority had a similar DWTT/Charpy E/A slope as the higher grade X100 steels; see some results in Figure 13. In fact, for one of these new steels there was also severe splitting (separations) on the fracture surface of the DWTT even at +10C. However, that steel had very mild splitting in the Charpy tests at that same temperature, where such splitting behavior frequently occurs due to thickness differences of the DWTT and Charpy specimens [16]. Since the splitting behavior reduces the energy absorbed [16], the DWTT E/A was lower than the Charpy E/A for this case. As a result, an estimated correction factor of 2.2 was needed for that steel, see Figure 13. The explanation for the older vintage low-grade steels behaving differently than the new low-grade steels is assumed to 5 Copyright 2012 by ASME

PN-DWTT E/A, ft-lb/in 2

be due to the change in steel-making procedures, in particular the cleanliness of current versus older vintage steels. The lowergrade pipes in the full-scale database were tested about 30 years ago using older manufacturing methods that gave consistently higher DWTT E/A values, while the X80 and X100 steels in the burst tests were fabricated only 5 to 15 years ago with newer procedures. In examining the DWTT/Charpy data for different grades of steel in the full-scale database, it can be seen that even at a particular grade level some pipes behaved like older vintage materials, while other pipes behaved like the newer X100 steels. The change in steelmaking was gradually occurring and appeared to be variable with different pipe mills. Today, it is starting to look like more mills are producing lower-grade steels that have similar fracture behavior to the high-grade steels. The good news is that current X60 to X70 steels have much higher Charpy energy values than older vintage steels and yield much greater initiation toughness; however, for ductile fracture arrest these new low-grade steels may also need correction factors on the Charpy by a factor of 1.7 or even higher if separations occur in the DWTT specimen, which could be a problem for some designs.
2.0

9,000

PN-DWTT E/A, ft-lb/in 2

8,000 7,000 6,000 5,000 4,000 3,000 2,000 1,000 0 0 500

Original 1977 Wilkowski equation for lower-grade steels (E/A)DWTT = 3(E/A)Charpy + 300, in-lb/in2 Linear regression of all X70 data (E/A)DWTT = 2.366(E/A)Charpy + 300, in-lb/in2

Trend with higher grade steels

1,000

1,500
2

2,000

Charpy E/A, ft-lb/in

(b) X70 grade pipes in past full-scale burst tests


9,000 8,000 7,000 6,000 5,000 4,000 3,000 2,000 1,000 0
Trend with some higher grade steels Original 1977 Wilkowski equation for lower-grade steels (E/A)DWTT = 3(E/A)Charpy + 300, in-lb/in2

Linear regression of all X80 data (E/A)DWTT = 2.53(E/A)Charpy + 300, in-lb/in2

Statistical correction factor from full-scale tests

1.8 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 60 70 80 90 100 110 120 DWTT energy based criterion
Original Battelle Equations - minimum Charpy energy Leis 2000 - minimum Charpy energy Wilkowski 1977 - minimum Charpy energy Wilkowski 2000 - minimum Charpy energy Wilkowski 2000 - minimum DWTT energy

Charpy energy based criteria

500

1,000

1,500

2,000

Charpy E/A, ft-lb/in2

(c) X80 grade pipes in past full-scale burst tests


9,000
All X100 data

8,000

JIP data Original 1977 Wilkowski equation for lower-grade steels (E/A)DWTT = 3(E/A)Charpy + 300, in-lb/in2

PN-DWTT E/A, ft-lb/in 2

Grade, ksi

7,000 6,000 5,000 4,000 3,000 2,000 1,000

Figure 10 Grade effect of correction factors on BTC Charpy energy determined from full-scale pipe burst test database as determined in [11]
9,000

PN-DWTT E/A, ft-lb/in

8,000 7,000 6,000 5,000 4,000 3,000 2,000 1,000 0 0 500

Original 1977 Wilkowski equation for lower-grade steels (E/A) DWTT = 3(E/A) Charpy Linear regression of all X60 data (E/A) DWTT = 2.94(E/A) Charpy

+ 300, in-lb/in

Linear regression through X100 data (E/A)DWTT = 1.82(E/A)Charpy + 300, in-lb/in2

0 0 500

+ 300, in-lb/in2

Charpy E/A, ft-lb/in2

1,000

1,500

2,000

2,500

(d) X100 grade pipes in past full-scale burst tests Figure 11 Change in DWTT to Charpy energy for pipes in the full-scale pipe burst test database by grade level from [11]
Trend with higher grade steels

1,000

Charpy E/A, ft-lb/in

1,500

2,000

(a) X60 grade pipes in past full-scale burst tests

Copyright 2012 by ASME

specimen. Based on recent findings, a material is susceptible to abnormal fracture behavior when: 1. 2. The Charpy upper-shelf energy is higher than 200 J. There is 100% shear area extending from the notch with a sudden change to a much lower SA% later along the fracture path (AFA). There is a sudden change in the PN-DWTT SA% from low values (>20%) to 100% shear area (AFB) at two close temperatures. One PN-DWTT specimen shows low shear area (>20%) and the other shows 100% shear area for duplicate PN-DWTT specimens (from the same pipe) tested at the same temperature (AFB).

3.

4.

Figure 12 Original 1977 relationship of DWTT versus Charpy energy per fracture area for steels at that time that were not in full-scale burst tests [3]
6,000 Mill A X70 - 2010 Mill B X70 - 2010 Mill I X70 - 2010 Mill W X65 - 2011 Mill N X70 - 2011 Mill J X60 - 2004

If any of these behaviors occurs, then alternative testing of the candidate materials with the static-precracked or modified backslot DWTT specimens (which are non-standard tests at this time but seem to be the best alternatives) is suggested. There are some on-going efforts to conduct pipe burst tests and compare to modified DWTT tests, but it will take additional efforts to ensure the results are general enough to be applicable. This will take several years before there is consensus on how to change API 5L3. The second aspect on the fracture behavior deals with ductile fracture propagation resistance when on the upper-shelf of the material. Many, but not all, newer low-grade steel pipes have indications that they need the same correction factor on the BTC Charpy energy as for X100 steels. The way to check this is to compare the Charpy and DWTT energy per fracture area (E/A) values as shown in Figure 13. Steels that exhibit separations on the fracture surface may require a higher correction factor on the Charpy energy than what has been observed from X100 pipe burst tests. Hence the past empirical equations in Codes and Standards like B31.8 may significantly under-predict the actual minimum Charpy energy values needed for ductile fracture arrest needed for most new line-pipe steels. It is also being proposed that API 5L3 start to include the measurement of DWTT energy in the standard in the future. Most large pipe producers already measure the energy values, some with sophisticated instrumentation. The use of the DWTT energy will help to alleviate the uncertainties of the correction factors needed on Charpy energy for ductile fracture arrest. REFERENCES [1] Maxey, W. A., and others, Brittle Fracture Arrest in Gas Pipelines, A.G.A. Report L51433, April 1983. [2] Wilkowski, G. M., Maxey, W. A., and Eiber, R. J., Examination of Critical Flaw Sizes Predictions and Fracture Propagation Transition Temperatures in

DWTT E/A, ft-lb/in

New low-grade steels behaving like old vintage steels

2
4,000

2,000

Original 1977 relationship for low grade X100 steel steels relationship

Large separations on DWTT but not Charpy fracture surfaces

New low-grade steels behaving like X100 steels

0 0 500

Figure 13 Recent Emc2 DWTT and Charpy energy/area results on new X60 to X70 steels CONCLUSIONS This paper demonstrates that there are two significant differences between fracture behavior in older-vintage low-grade steels and newer low-grade steels ( X70 grades). Of these two aspects, the abnormal fracture behavior is perhaps the most troublesome since brittle fracture is generally a higher concern than unstable ductile fracture propagation. Also, it is quite possible to observe 100% shear area in the standard PN-DWTT specimen whereas the pipe may still be susceptible to brittle fracture. This is a greater concern because API 5L3 (and comparable international DWTT standards) does not have a good alternative test to use (chevron-notch DWTT is to be eliminated as an option as discussed earlier). Consequently, pipeline designers need to be informed so that materials can be tested in a different manner than just the standard PN-DWTT

Charpy E/A, ft-lb/in2

1,000

1,500

2,000

2,500

Copyright 2012 by ASME

Experimental Arctic Line Pipe, Battelle report to Northern Engineering Services, June 30, 1976. (Not available to general public) [3] Wilkowski, G. M., Problems in Using the Charpy, Dynamic Tear Test, and Drop Weight Tear Test for High Toughness Quenched and Tempered Steels, Presentation 9, A.G.A.EPRG Line Pipe Research Seminar III, November 15-16, 1978. [4] Wilkowski, G. M. and Eiber, R. J. Problems in Using the Charpy & DWTT for High Toughness Q&T Steels, What Does Charpy Energy Really Tell Us? published by ASM, ISBN 0-87170-027-1, February 26-March 2, 1978, pp. 201226. [5] Yamura, J., et al., Some Experiments on Brittle Fracture Arrestability of As-Rolled and Quenched and Tempered Steel for Line Pipes, Sumitomo Metal Industry Ltd., Osaka, Japan, March 1975. [6] Kashimura, H. Ogasawara, M., and Mimura, H., Modifications and Analytical Study of DWTT for Tougher Materials, 1975 Material Science Symposium by ASM Material Science Division and AIME, The Metallurgical Society, Cincinnati, OH, Nov 11-13, 1975. [7] Wilkowski, et al. Recent Development on Determining Steady-State Dynamic Ductile Fracture Toughness from Impact Tests, Proceedings of 3rd International Pipeline Technology Conference, Brugge, Belgium, May 21-24, 2000, Volume 1, pp. 359-386. [8] Wilkowski, G. M., Shim, D. -J., and Brust F. W., Rudland, D. L., and Duan, D.-M., Evaluation of Fracture Speed on Ductile Fracture Resistance, 2009 Pipeline Technology Conference, Belgium, October 2009. [9] Shim, D.-J., Wilkowski, G. M., Duan, D.-M., and Zhou, J., Effect of Fracture Speed on Ductile Fracture Resistance, IPC2010-31634, Proceedings of the 8th International Pipeline Conference, September 27 - October 1, 2010, Calgary, Alberta, Canada. [10] Maxey, W., Kiefner, J. F., and Eiber, R. J., Ductile Fracture Arrest in Gas Pipelines, A.G.A. catalogue number L32176, May 1976. [11] Wilkowski, G., Rudland, D., Xu, H., and Sanderson, N., Effect of Grade on Ductile Fracture Arrest Criteria for Gas Pipelines, paper #IPC2006-10350, 2006 International Pipeline Conference. [12] Leis, B.N., Predicting Fracture Arrest Based on a Relationship Between Charpy Vee-Notch Toughness and Dynamic Crack-Propagation Resistance, Proceedings of 3rd International Pipeline Technology Conference, Pipeline Technology, Volume I, R. Denys (editor), Brugge, Belgium, May 21-24, 2000, Volume 1, pp 407-420. [13] Wilkowski, G., Rudland, D., and Wang, Y. Y., Recent Development On Determining Steady-State Dynamic Ductile Fracture Toughness From Impact Tests, Proceedings of 3rd International Pipeline Technology

[14]

[15]

[16]

[17]

Conference, Brugge, Belgium, May 21-24, 2000, Volume 1, pp. 359-386. Rudland, D., Shim, D.-J., Xu, H., Rider, D., Mincer, P., Shoemaker, D., and Wilkowski, G., First Major Improvements to the Two-Curve Ductile Fracture Model, Emc2 Report to U.S. DOT and PRCI, May (2007). Wilkowski, G. M., Rudland, D. L., Wang, Y.-Y., Horsley, D., Glover, A., and Rothwell, B., Determination Of The Region Of Steady-State Crack Growth From Impact Tests, in Proceedings of IPC-2002 Conference, October 2002, Calgary, Alberta, Canada. Wilkowski, G. M., and Eiber, R. J., Ductile Fracture Propagation Resistance for Rising Shelf Controlled-Rolled Steels, What Does Charpy Energy Really Tell Us? published by ASM, ISBN 0-87170-027-1, February 26March 2, 1978, pp. 108-132. Wilkowski, G. M., Maxey, W. A., and Eiber, R. J., Use of a Brittle Notch DWTT Specimen to Predict Fracture Characteristics of Line Pipe Steels, ASME 1977 Energy Technology Conference, Houston, Texas, Paper 77-Pet-21, September 18-22, 1977.

Copyright 2012 by ASME

Anda mungkin juga menyukai