Sei sulla pagina 1di 9

Materials Transactions, Vol. 47, No. 2 (2006) pp.

409 to 417 #2006 The Japan Institute of Metals

High-Cycle Fatigue Behavior of Type 316L Stainless Steel


Jiunn-Yuan Huang* , Ji-Jung Yeh, Sheng-Long Jeng, Charn-Ying Chen and Roang-Ching Kuo
Institute of Nuclear Energy Research (INER), P. O. Box 3-14, 1000 Wenhua Road, Chiaan Village, Lungtan, Taiwan 325, R. O. China.
High-cycle fatigue tests were conducted to investigate the eects of temperature, stress ratio (R), specimen orientation, welding and specimen size on the fatigue behavior of type 316L stainless steel. The high-cycle fatigue test results indicated that the fatigue limits signicantly decreased when the stress ratio (R) decreased. The corresponding fatigue limits were reduced to lower values when tests were conducted at 300 C, compared to those obtained at room temperature. The fatigue behavior and fatigue limits of standard and subsize specimens were observed to be consistent at both room temperature and 300 C. The constant life diagram was established from the SN curves acquired. The fatigue limit strongly depended on the materials strength, which was a function of specimen orientation, test temperature, and welding processes. The dimension of the fatigue damaged area on a fracture surface increased as the stress ratio decreased. In the case of R 1:0, the fatigue damaged region extended over the whole fracture surface. The subgrain boundaries after high-cycle fatigue tests were clearly demonstrated by their diraction patterns, which were related to the dynamic recovery of multiple dislocations. (Received September 27, 2005; Accepted December 9, 2005; Published February 15, 2006) Keywords: high-cycle fatigue, stress ratio, subsize, constant life diagram, subgrain boundary

1.

Introduction

Type 316L stainless steel is one of the important structural materials used for the in-core components and pressure boundaries of light water reactors (LWR). Ensuring structure integrity of these components is crucial to the operational safety of nuclear power plants. Welds are made in reactor internal systems to provide the advantages of a continuous metal joint. It has been reported that failures usually occur at the degradation of a component or components to which the weld is axed.1) The largest number of weld failures is attributed to either the high-cycle fatigue caused by owinduced vibration or due to fatigue in general. In addition to stress corrosion cracking and radiation damage, fatigue is one of the dominant mechanisms responsible for the degradation of pressure boundaries and many in-core components.25) The fatigue damage is induced by cyclic variations of pressure (stress), temperature and ow induced vibration, thereby reducing the lifetime of structures and components.68) In order to understand the eects of stress ratio and environment temperature on the high-cycle fatigue behavior of type 316L stainless steel and its weld, the fatigue limits and the fracture modes were investigated at room temperature and 300 C. The subsize specimen911) was also studied in this work. Comparisons were made between the data generated from subsize and standard specimens to examine the eect of specimen size on the fatigue behavior. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) were used to study the microstructural evolution of SS316L steel and to characterize their fatigue resistant behavior. 2. Experimental Procedures

type 316L stainless steel were given in Tables 1 and 2. A solution of 10 vol% oxalic acid was used as the electrolytic etching solution. 2.2 Welding procedures The weld specimen design, comprising two beveled test plates with a V groove of 75 , a root opening gap of 2.0 mm and root face of 1.6 mm, is shown in Fig. 1. Welding was performed by a manual GTAW (Gas Tungsten Arc Welding) butt weld using ller metal ER316L with a diameter 1.6 mm for the rst two passes and 3.2 mm for the subsequent ones. Twelve weld passes were performed in four layers. The weld current was set at 60 amperes for the rst two passes and then raised up to 8090 amperes for the rest passes. 2.3 High-cycle fatigue test Standard and subsize fatigue specimens of plate type were designed to the specications of ASTM E 466,12) as shown in Fig. 2. The dimensions of a subsize specimen, thickness of 2.6 mm and gage length of 10.4 mm, were in the lower limit of the specications. The overall length of a standard specimen was 190 mm and that of the subsize 67 mm. Before fatigue testing, all specimens were polished as per the recommendations of ASTM E 466. High-cycle fatigue tests, at room temperature and 300 C, were performed on a 100 KN MTS 810 close-loop servohydraulic machine under a sinusoidal load control, at stress ratios (R min =max ) of 0.8, 0.2, 0:2 and 1:0, and at a frequency of 20 Hz. For the tests at 300 C, specimens were heated to 300 C and then maintained for 2 h to homogenize the temperature distribution. The standard specimens were loaded in the rolling direction, and the subsize specimens either in a direction parallel or transverse to the rolling direction. Fatigue tests were stopped when specimens broke or the fatigue cycles reached 107 cycles. 2.4 Scanning electron microscopic (SEM) examination The fracture surfaces were examined by scanning electron

2.1 Materials The chemical compositions and mechanical properties of


*Corresponding

author, E-mail: jyhuang@iner.gov.tw

410

J.-Y. Huang, J.-J. Yeh, S.-L. Jeng, C.-Y. Chen and R.-C. Kuo Table 1 Chemical compositions of type 316L stainless steel.

Designation C wt% 0.02 Si 0.3 Mn 1.46 P <0:03

Composition (mass%) S <0:02 Ni 12.3 Mo 2.28 Cr 17.10 Fe Bal.

Table 2 Tensile properties of type 316L stainless steel. Designation As-received (Rolling) As-received (Transverse) Weld Temperature ( C) 25 300 25 300 25 300 Ultimate tensile strength, UTS (MPa) 526 405 544 432 538 441 Yield strength 0.2% oset YS (MPa) 245 150 245 145 306 288 Total elongation (%) 65 53 63 39 30 25 Uniform elongation (%) 50 32 48 33 23 15.7

Fig. 1 A sketch of weld specimen design.

Rolling direction
L1 W1 L2 R L3 T Rolling Direction W2

Fig. 3

Optical micrographs of type 316L stainless steel.

Designations Large (Standard specimen) Subsize specimen

Reduced T W1 L1 W2 L2 L3 R W1 L1 W2 R cross section (mm) (mm) (mm) (mm) (mm) (mm) (mm) T W1 W1 W1 area (TW1) 5 2.6 10 30 30 40 190 85 41.6 2 2 3 2 3 2 8.5 8 50 mm2 13.52 mm2 19.4 mm2 645 mm2 (preferable)

5.2 10.4 10.4

14 67

2.54 ASTM E466 (as specifications possible as)

2 2 6 3

1.5

Fig. 2 Dimensions of standard and subsize high-cycle fatigue specimens as per ASTM E 466 specications.

microscopy to characterize the fracture mode, crack initiation site and the evolution of the fatigue striation structure. 2.5 Transmission electron microscopic (TEM) examination To examine the fatigue eects on the evolution of the dislocation structure on the surface and bulk of fatigue-tested

specimens, TEM samples were cut from dierent locations along the as-received and fatigued specimens with their sample surfaces normal and parallel to the loading axis, respectively. In order to prevent the nal plastic deformation from complicating the fatigue structure, the sampling position should be kept away from the necking region, but as close to the nal fracture site as possible. TEM samples were polished with ne emery paper to a thickness of 0.1 mm, then chemically thinned with a double-jet polishing machine in 10 vol% perchloric acid (HClO4 ) and 90 vol% methanol (CH3 OH) at 30 C. The current was approximately controlled at 0.1 A and voltage at 10 V. The transmission electron microscope used in this study is a JOEL 2000FX. 3. 3.1 Results and Discussion Metallographic features of the as-received and welded specimens There was no clear disparity in microstructural features

High-Cycle Fatigue Behavior of Type 316L Stainless Steel


600

411
SN316OT.grf

Weld
Maximum Stress, Smax /MPa
500

R=0.8
run-out run-out run-out

R= 0.2 R=0.8
run-out run-out run-out

Base metal Base metal

400
R= 0.2
run-out run-out run-out

300

Standard Specimen Soild symbol at room Temperature Open symbol at 300C

R= -0.2
run-out

run-out

R= -0.2
run-out

Fig. 4

Optical micrograph of the type 316L stainless steel weld specimen.

200

R= -1.0

run-out run-out

R= -1.0

run-out run-out

100 1.0E+0 1.0E+1 1.0E+2 1.0E+3 1.0E+4 1.0E+5 1.0E+6 1.0E+7 1.0E+8
316HV

300.00

Fatigue Life, N /Cycles

cap root

Weld root
250.00

Fig. 6 SN curves for standard specimens of type 316L stainless steel tested under dierent stress ratios at room temperature and 300 C.

Weld cap

Weld

Hardness (Hv, 300g)

200.00

150.00

Weld

100.00 -12.00 -8.00 -4.00 0.00 4.00 8.00 12.00

Distance along the center of weld, d/ mm

Fig. 5

Micro-hardness measurement along the weld root and weld cap.

observed along the rolling, transverse and short transverse direction, as demonstrated in the optical micrographs of Fig. 3. Flow lines induced by hot rolling were observed to be along the rolling direction. Given in Fig. 4 is an example of the optical micrograph of a weld specimen. Dendritic structures were observed and prevalent in the weld. In general, dendrites grew in the direction parallel to the heat transfer direction. So dendrites were observed to lie in a direction vertical to the interface between the weld and base metal, as shown in Fig. 4. The heat-aected zone is not distinguishable. Based on the Vickers micro-hardness measurements, the melted zone is harder than the base metal, as illustrated in Fig. 5. And further the hardness values of the weld root are relatively higher than those for the weld cap. 3.2 High-cycle fatigue test The fatigue limits for type 316L stainless steel subjected to high-cycle fatigue tests were summarized in Table 3. The
Table 3 Designation Stress ratio Room temperature 300 C 0.8 519.8 408.3

fatigue limits were mainly a function of stress ratio and test temperature. The fatigue limits were observed to increase with increasing the stress ratio. When the stress ratio was set at 1:0, the fatigue limit decreased to a minimum value. On the other hand, the corresponding fatigue limits at 300 C were lower than those tested at room temperature. An important nding was that the fatigue limits for the subsize specimens were consistent with those for the standard specimens at both room temperature and 300 C. Figure 6 shows the SN curves for standard specimens tested under dierent stress ratios at room temperature and 300 C. In this gure, there were 4 sets of SN curves generated for each test temperature. The corresponding stress ratios were at 0.8, 0.2, 0:2, and 1:0, respectively. Each S N curve is comprised of a slanting line and a horizontal line, which depicts the relationship between the fatigue life, log(N), and maximum stress, Smax , as also given in Fig. 6. The horizontal line is the level of the fatigue limit. From the gure, it can be seen the fatigue limit for R 0:8 was the largest, and the one for R 1:0 was the smallest irrespective of the test temperature. The results indicated the cyclic actions of tension and compression inicted the severest fatigue damage. On the other hand, the cyclic actions of tension and tension exerted a smaller stress amplitude, so the fatigue damage was moderate. At 300 C, the fatigue life was relatively shorter and the fatigue limit lower. From the gure, it can be seen that the fatigue limit at a specic temperature was completely determined by the R value, but that the fatigue life was determined by both R and Smax . The high test temperature (300 C) was detrimental to the highcycle fatigue properties of type 316L stainless steel. Comparing the SN results for the subsize specimens with those for the standard specimens indicates that both speci-

The fatigue limits for type 316L stainless steel subjected to high-cycle fatigue tests. Standard (rolling) 0.2 0:2 269.7 220.7
7

Subsize (rolling) 1:0 183.9 122.6 0.8 519.8 404.6 0.2 404.5 N/A 0:2 269.7 220.7 1:0 183.9 N/A

Subsize (transverse) 0:2 318.7 245.2

392.7 318.8

The unit for fatigue limit (the maximum stress for N = 10 cycles): MPa

412
(a)
600
R=0.8 log(N)=110.644177-0.199111 (Smax)

J.-Y. Huang, J.-J. Yeh, S.-L. Jeng, C.-Y. Chen and R.-C. Kuo
600
Subsize Specimen at R = -0.2 SN316STr

SN316SS.grf

run-out

Maximum Stress, Smax /MPa

500

Transverse Direction at Room Temperature Rolling Direction at Room Temperature Transverse Direction at 300 C Rolling Direction at 300 C

Maximum Stress, Smax /MPa

500
log(N)=9.25577876-0.008557 (Smax)

run-out run-out run-out

400

400

At room temperature Solid symbol : standard specimen Open symbol : subsize specimen

R=0.2

run-out run-out

run-out

300
run-out run-out

300

log(N)=8.641271-0.010003 (Smax)

R=-0.2

run-out run-out run-out

200
200
log(N)=14.092104-0.041524 (Smax) R=-1

run-out

run-out run-out run-out

100 1.0E+1 1.0E+2 1.0E+3 1.0E+4 1.0E+5 1.0E+6 1.0E+7 1.0E+8


1.0E+8

100 1.0E+1 1.0E+2 1.0E+3 1.0E+4 1.0E+5 1.0E+6 1.0E+7

Fatigue Life, N /Cycles

Fatigue Life, N/Cycles

(b)
500
SN316OST

Fig. 9 SN curves for SS316L subsize specimens with dierent orientations at room temperature and 300 C.

Maximum Stress, Smax /MPa

R= 0.8

400

log(N)=800.916667-1.944444 (Smax) R= 0.2

run-out run-out

300

At 300C Soild symbol: standard specimen Open symbol: subsize specimen

log(N)=13.084738-0.020667 (Smax)

run-out

(a)
600
SN316w1

R= -0.2 log(N)=8.678040-0.012747 (Smax)


Weld at R = 0.2 R=0.2

Maximum Stress, Smax /MPa

run-out

500

Weld at R = -0.2 Base Metal at R = 0.2 Base Metal at R = -0.2 log(N)=8.096182-0.05717 (Smax)

200
R= -1.0 log(N)=14.055435-0.060210 (Smax)

400
R= 0.2

run-out

run-out run-out run-out


log(N)=15.231225-0.029238 (Smax)

100 1.0E+0 1.0E+1 1.0E+2 1.0E+3 1.0E+4 1.0E+5 1.0E+6 1.0E+7 1.0E+8

Fatigue Life, N/Cycles

300

run-out

run-out run-out

Fig. 7 SN curves for standard and subsize SS316L specimens with dierent stress ratios at (a) room temperature and (b) 300 C.

200

100 1.0E+1 1.0E+2 1.0E+3 1.0E+4 1.0E+5 1.0E+6 1.0E+7 1.0E+8

Fatigue Life, N/Cycles

(b)
600
Weld at R = 0.2 Weld at R = -0.2 SN316wt1

600
R=-0.2 R=0 R=+0.2 R=+0.8
(424.4,530.6) (99.5,497.3) (421.7,527.1) (415.8,519.8)

Maximum Stress, Smax /MPa

(a)
550 500
les 5 cyc 10 N=
(83.1,415.7)

500

Base Meatl at R = 0.2 Base Metal at R= -0.2

55 50 0

400

Welded R=0.2 Base Metal R=0.2 log(N)=13.868428-0.022320 (Smax)

Maximum Stress, Smax /MPa

450 400 350 300 250 200 150 100 50


35 R 0

N=

5x

cy 10
5

cle

s
ue Lim it)

=-

(78.5,392.7)

(C

(-72.8,364.0)

N=

7 cyc 10

les

(F

g ati

45 t es ) 0 0 40

run-out run-out
log(N)=13.084738-0.020667 (Smax) log(N)=10.818176-0.018989 (Smax)

300

Welded R= 0.2 Base Metal R= 0.2

om

ple

te

30

ev

(-58.8,294.1)

er

R
(-53.9,269.7)

=+

(T

en

sil

35

run-out

sa

l)
(-219.0,219.0)

25

Yield Point
25 20 0 15 0

0 Sa

0 a 30 MP , Sm

200

run-out
log(N)=8.678040-0.012747 (Smax)

20 ,M 0 Pa

(-202.1,202.1) (-183.9,183.9)

15

0 10 0 50 50 0

10

45

316L stainless steel, ASTM E-466 standard, axial load tests, test performed in air at room temperature, SU =526MPa, SY =245MPa, frequency = 20 Hz (with fan cooling in the case of R = -1.0, N=105 cycles)

100 1.0E+1 1.0E+2 1.0E+3 1.0E+4 1.0E+5 1.0E+6 1.0E+7 1.0E+8

Fatigue Life, N/Cycles

0 -400 -350 -300 -250 -200 -150 -100 -50

50 100 150 200 250 300 350 400 450 500 550 600

Minimum Stress, Smin /MPa

Fig. 10 SN curves for weldment and base metal specimens of SS316L stainless steel at (a) room temperature, and (b) 300 C.

(b)
600
R=-0.2 R=0 R=+0.2 R=+0.8

550 500
50 45 40 0 0 0

55

450
(327.5,409.3)

400
1 =R

350 300 250 200 150 100 50

35

30

(-57.7,288.5)

25

(-46.7,233.7) (-44.1,220.7)

(327.2,409.0) 5 cycles (326.6,408.3) N=10 es 5 ) cycl imit) st X10 ue L (71.5,357.4) N=5 Te atig (F e 7 cles sil 0 cy n 0 e 35 N=1 (T 1 (63.8,318.8) =+ R 0 Pa 30 , M Sm 0 25 (78.2,391.2)

20

0 15 0 10 0 50 50 10 0 15 0

20

(-150.4,150.4)

(-138.8,138.8) (-122.6,122.6)

316L stainless steel, ASTM E-466 standard, axial load tests, test performed in air at 300C, frequency = 20 Hz

45

0 -400 -350 -300 -250 -200 -150 -100 -50

50 100 150 200 250 300 350 400 450 500 550 600

Minimum Stress , Smin /MPa

Fig. 8 Constant life diagrams for subsize specimens of type 316L stainless steel subjected to high-cycle fatigue tests at (a) room temperature, and (b) 300 C.

mens have similar SN curves and the same fatigue limits. An example is given in Figs. 7(a) and (b) to compare the SN curves for standard and subsize specimens tested at room temperature and 300 C, respectively. It can be concluded that the high-cycle fatigue properties of standard and subsize specimens were equivalent at either test temperature. Therefore, the subsize specimen could be employed as a substitute to the standard specimen. The SN curves obtained and their corresponding fatigue life equations provide a database for evaluating the high-cycle fatigue properties of type 316L stainless steel components. The constant life diagrams for high-cycle fatigue tests at room temperature and 300 C, as shown in Figs. 8(a) and (b), were derived from Figs. 6 and 7. There are 2 sets of

Maximum Stress , Smax /MPa

te ple om (C l) sa er ev R
Sa Pa ,M

High-Cycle Fatigue Behavior of Type 316L Stainless Steel

413

Fatigue damaged region Fatigue damaged region

Fatigue damaged region

Smax= 490 MPa

Smax= 441 MPa

Smax= 404 MPa

(a)

(b)

(c)

Fig. 11 SEM micrographs showing the fatigue damaged region gradually expanded when the maximum applied stress (Smax ) was successively decreased at the R value of 0.2.

Ductile rupture region

Fatigue damage region

Crack initiation

(a) R= - 0.2

(b) R= - 1.0

Fig. 12 SEM micrographs showing the fatigue damaged region extended over the whole fracture surface with R ratio decreasing from (a) R 0:2 to (b) R 1.

coordinates in each constant life diagram. The basic coordinates are the maximum applied stress (Smax ) and minimum applied stress (Smin ). The coordinates inclined at an angle of 45 are the stress amplitude (Sa ) and mean stress (Sm ). Mean stress eect could be obtained from the constant life diagram. The stress amplitude drops sharply when mean stress is greater than 200 MPa. Any points with a combination of Smax , Smin , Sa and Sm below a certain constant fatigue life curve, no high-cycle fatigue failure is expected to occur before the predicted life. Thus the established constant life diagrams can serve as a guide for safe operation and design of relevant in-core components, so as to prevent the unexpected failure due to the high-cycle fatigue. Transverse subsize specimens, with the loading axis perpendicular to the rolling direction, were also tested. A comparison of the results, as illustrated in Fig. 9, shows the transverse specimens exhibit slightly higher fatigue limits than those specimens loaded in a direction parallel to the rolling direction. The welded specimens using the ER316L as ller were tested under high-cycle fatigue loading conditions at room

temperature and 300 C. As shown in Figs. 10(a) and (b), the welded specimens exhibit slightly higher fatigue limits than those of base metal specimens. It can be concluded that the fatigue limit strongly depends on the materials strength, as shown in Figs. 6, 9, 10 and Table 2. The higher the materials strength, the higher the fatigue limit was observed. 3.3 Fractographic features of high-cycle fatigued specimens The fatigue damaged region gradually expanded when the maximum applied stress (Smax ) was successively decreased under the R value of 0.2, as illustrated in Fig. 11. It could be that the ligament area could sustain the lower maximum applied stress during the fatigue crack propagation. Thus a fatigue specimen under a lower applied maximum stress led to a larger fatigue damage region and smaller nal rupture plastic region. The eect of stress ratio on the fracture features is exemplied in Fig. 12. The more negative stress ratio could yield a larger fatigue damaged region. In the case of R 1:0, the fatigue damaged region extended over the whole fracture surface. The specimen was fractured mainly

414

J.-Y. Huang, J.-J. Yeh, S.-L. Jeng, C.-Y. Chen and R.-C. Kuo

Fig. 13 Multiple initiation sites observed at R 0:2.

by a fatigue rupture mode. The fatigue striations were the prevalent features. In the case of R 0:2, part of the fracture surface was featured by ductile rupture. The specimens appeared to have a smaller plastic region when tested at R 0:2, relative to those at R 0:2, Figs. 12(a) and 11. This nding is in good agreement with the above-mentioned conclusion that the lower applied maximum stress leads to a larger fatigue damage region and smaller plastic region. The fatigue crack initiated from the specimen surface. As is often the case with the high-cycle fatigued specimens, a main initiation site was observed to be responsible for fatigue rupture. But in some cases, as demonstrated in Fig. 13, there were more than one initiation sites observed. The multiple initiation sites were observed with the severe fatigue condition of R  0:2. The reason could be that the lower applied maximum stress yield a larger fatigue damage region, which could induce more crack initiation sites. In other words, the higher applied maximum stress would result in nal rupture as soon as the fatigue initiation is formed. The spacing between fatigue striations was strongly dependent on the value of R, as demonstrated in Fig. 14. The more negative the stress ratio, the wider the spacing between the fatigue striations. That could be accounted for by the observation that the larger stress range led to the higher crack growth rate. So the striation spacing at the equivalent crack length was observed to be larger for the more negative stress ratios. Most of the type 316L stainless steel specimens subjected to high-cycle fatigue failed at the base metal, some at the weld, illustrated in Table 4. That could be attributed to the material strength of the weld greater than that of base metal. But some failure occurred at the weld, which could be due to the micro-defect in the weld. These micro-defects were beyond the detection limit of X-ray inspection. Figure 15(a) shows the micro-defect observed at the fatigue initiation site. The striations were the prevalent features for the weld fracture surface, as shown in Fig. 15(b). No dendritic structures were observed at the weld because the dendrite orientation was roughly perpendicular to the fracture surface. Itow reported that interdendritic cracking morphology could be observed on the fracture surface of compact tension

(a)

(b)
Fig. 14 SEM micrographs showing the fatigue striations spacing is determined by the value of R, (a) R 0:2, (b) R 1:0.

High-Cycle Fatigue Behavior of Type 316L Stainless Steel Table 4 Test temperature ( C) Failure locations of weld specimens for type 316L stainless steel subjected to high-cycle fatigue tests. Specimen No. WHF-7 25 WHF-8 WHF-11 WHF-15 WHF-14 WHF-17 25 WHF-18 WHF-19 WHF-20 TWHF-4 TWHF-3 300 TWHF-2 TWHF-1 TWHF-8 TWHF-6 300 TWHF-5 TWHF-7 0:2 0.2 0:2 0.2 R ratio Applied max. stress 416.8 441.4 441.4 453.6 490.4 294.2 318.7 343.2 343.2 343.2 343.2 367.8 392.3 392.3 245.2 269.7 294.2 Fatigue life >107 370,480 373,776 325,952 196,096 >10
7

415

Failure location Run-out Base metal Base metal Base metal Base metal Run-out Weld Weld Weld Run-out Base metal Base metal Weld Weld Run-out Base metal Base metal

818,560 131,936 182,656 >10


7

351,872 456,256 125,846 133,216 >107 497,664 170,496

specimen when the crack propagates parallel to the alloy 182 weld dendrite direction.13) Dislocation structure of high-cycle fatigued specimens A number of researchers1420) have observed a strong dependence of dislocation structures on fatigue behavior. Figure 16 shows the typical TEM images of surface and cross sectional layers from the as-received and fatigued specimens. As a consequence of work-hardening by plate rolling, the dislocation density of the surface is much higher than that of the bulk, as shown in Figs. 16(a) and (b). In the cross sectional layer, planar defects [see Fig. 16(b)] exhibiting fringe contrast are generally described as stacking faults or microtwins due to their similar contrast. Microtwins, which are closely structurally related to stacking faults, has been fully discussed by Chen and Stobbs.21) Such thin twins were observed to be distorted and associated with dislocations after fatigue tests, as shown in Figs. 16(d) (f). It was also noted that in the surface layers of the fatigue-tested specimens at room temperature, dislocations were observed to arrange themselves on {111} slip planes, as illustrated in Fig. 16(c). At 300 C the dynamic recovery occurred, as illustrated in Fig. 16(e). The dislocations migrated from their slip planes into subgrain boundaries (or cell walls). That would lower the average strain energy associated with the dislocations22) and thus tends to lower the work-hardening produced by the fatigue test. These subgrain boundaries were clearly demonstrated by their diraction patterns, as the insets in Fig. 17 showing low-angle boundaries. The subgrain for each grain is near to [011] zone under the same imaging condition. The dynamic recovery in type 316L stainless steel is believed to be related to the test temperature and dislocation density. So the failure mechanism for the type 316L stainless steel at 300 C might be inuenced by the dynamic recovery. 3.4

Micro-defect

Initiation site
(a)

(b)

Fig. 15 SEM micrographs showing (a) micro-defects were observed at the fatigue initiation site, (b) striations were prevalent features for the fracture surface of weld specimens subjected to applied max. stress of 343.2 MPa and R ratio of 0:2 at room temperature.

416

J.-Y. Huang, J.-J. Yeh, S.-L. Jeng, C.-Y. Chen and R.-C. Kuo

Surface layers

Cross section

R= -1

Fig. 16 TEM images of Type 316L showing signicant changes of microstructures between surface layers (a, c, e) and cross sectional layers (b, d, f). (a, b) as-received specimens, (c, d) fatigued specimens tested with t 392 MPa at room temperature, and (e, f) fatigued specimens tested with t 319 MPa at T 300 C. Micrographs were taken with the beam direction near to [011]. Scale bar in (c) applied to all images as well.

4.

Conclusions

(1) The fatigue limit of type 316L stainless steel strongly depended on the materials strength, which was a function of specimen orientation, test temperature, and welding processes. (2) The high-cycle fatigue properties of subsize specimens as dened in this report were the same as those for standard specimens at both room temperature and 300 C. Therefore, it could be employed as a substitute to the standard specimen. (3) The dimension of the fatigue damaged region of a fracture surface increased as the stress ratio decreased. In the case of R 1:0, the fatigue damaged region extended over the whole fracture surface. The fatigue striations were the prevalent features. (4) The subgrain boundaries in the fatigue specimens were clearly demonstrated by their diraction patterns,

Fig. 17 TEM image in Fig. 16(e) showing subgrain boundaries as illustrated by the inset diraction pattern for each subgrain which is near to [011] zone under the same imaging condition.

High-Cycle Fatigue Behavior of Type 316L Stainless Steel

417

which were related to the dynamic recovery of multiple dislocations caused. Acknowledgments The authors would like to acknowledge the technical support provided by Mr. Jiunn-Shyoung Huang and KenFeng Chien. REFERENCES
1) D. R. Forsyth, B. L. Silverblatt, T. R. Mager, W. A. Bamford, J. A. Tortorice, J. T. Crane and I. L. W. Wilson: Aging management evaluation for reactor internals, (WCAP-14573, 1997), chap. 2, p. 30. 2) D. J. Wulpi: Understanding How Components Fail, (ASM, 1985), pp. 117162. 3) A. J. Allen, D. J. Buttle, C. F. Coleman, F. A. Smith and R. L. Smith: Microstructural examination of fatigue accumulation in critical LWR components, EPRI NP-5590, Jan. 1988. chap. 2, pp. 114. 4) V. N. Shah and P. E. MacDonald: Residual life assessment of major light water reactor components-overview, (NUREG/CR-4731, EGG242469, Vol. 1, June 1987), pp. 1941. 5) V. N. Shah and P. E. MacDonald: Aging and life extension of major light water reactor components, (Elsevier Science Publisher, 1993), Chap. 2, pp. 1921 and chap. 16, pp. 565567. 6) K. H. Luk: Boiling-water reactor internals aging degradation study, (ORNL NUREG/CR-5754, 1993), pp. 2224. 7) K. H. Luk: Pressurized-water reactor internals aging degradation

study, (ORNL NUREG/CR-6048, 1993), pp. 3132. 8) K. Iida: Nucl. Eng. Des. 138 (1992) 297312. 9) S. Jeelani, R. Natarajan and G. R. Reddy: Int. J. Fatigue 3 (1986) 159 164. 10) K. C. Liu and M. L. Grossbeck: Use of subsize fatigue specimens for reactor irradiation testing, (ASTM STP 888, 1986), pp. 276289. 11) J. J. Yeh, J. Y. Huang and R. C. Kuo: High-cycle fatigue behavior of type 316L Stainless Steel, International symposium on experimental mechanics (ISEM), Taipei Grand Hotel, Taiwan, Dec. 2830, 2002. 12) Standard Practice for Conducting Force Controlled Constant Amplitude Axial Fatigue Tests of Metallic Materials, (ASTM E 466-96, 1997), pp. 466470. 13) M. Itow, Y. Abe, A. Sudo and T. Kaneko: Seventh international symposium on environmental degradation of materials in nuclear power systems-water reactors, ed. by G. Airey et al., (Breckenridge, Colorado, August 710, 1995), pp. 541550. 14) C. Fukuoka, H. Yoshizawa, Y. G. Nakagawa and M. E. Lapides: Metall. Trans. A 24A (1993) 22092216. 15) C. Laird: Metall. Trans. A 8 (1977) 851860. 16) C. Fukuoka and Y. G. Nakagawa: Scr. Mater. 34 (1996) 14971502. 17) K. Katagiri, A. Omura, K. Koyanagi, J. Awatani, T. Shiraishi and H. Kaneshiro: Metall. Trans. A 8 (1977) 17691773. 18) D. Kuhlmann-Wilsdorf and C. Laird: J. Mater. Sci. Eng. 27 (1977) 137156. 19) J. Awatani, K. Katagiri and H. Nakai: Metall. Trans. A 9 (1978) 111 116. s : Fatigue of Metallic Materials, (Elservier 20) M. Klesnil and P. Luka Scientic Publishing Company, 1980) p. 30. 21) C. Y. Chen and W. M. Stobbs: Ultramicroscopy 58 (1995) 289305. 22) Robert E. Reed-Hill: Physical Metallurgy Principles, (Van Nostrand Company, New York 1973), pp. 267325.

Potrebbero piacerti anche