Вы находитесь на странице: 1из 9

Low Cycle Fatigue Behavior of Ren6 80 at Elevated Temperature

STEPHEN D. ANTOLOVICH, S. LIU, AND R. BAUR Low cycle fatigue of Ren6 80 was studied at 871 and 982 ~ It was found that when the data were represented on the basis of plastic strain, the life increased with decreasing frequency and imposition of a 90 s hold at maximum strain. Transmission electron microscopy studies showed that the 7' coarsened and an interfacial array of edge dislocations developed. The density of dislocations in the matrix was very low. Light optical microscopy revealed that cracks generally inititiated at oxide spikes in surface connected grain boundaries. A crack initiation criterion based on the maximum stress and oxide depth at the time of crack initiation was found to represent the data very well. Based on that representation, an expression for the initiation fatigue life was developed. That expression includes temperature, frequency and cyclic stress strain parameters as variables.

I. INTRODUCTION L O W cycle fatigue (LCF) at elevated temperature is an important consideration in the design of turbine components such as disks and turbine blades. Various methodologies for predicting the life of such components have been proposed 1~ and one of these methodologies ~has been the subject of an international conference? In addition, an exhaustive review of available data and life prediction schemes has recently been made. 6 It is striking that various approaches seem to work well for some alloy systems but not for others and that a universally applicable fatigue law has not been discovered. This is probably due to the fact that in different classes of materials, different forms of damage can occur. For example, grain boundary sliding and cavitation might occur relatively easily in those systems where there is no pinning of the boundaries by precipitate or carbide particles. On the other hand, some systems may be very prone to environmental attack. Still others will fail due to the accumulation of deformation debris. Not only will the failure mode depend on the system being considered, but also on the temperature, strain range and strain rate. In addition, interactions between the modes of damage accumulation can occur. Metallurgical factors have been discussed in review papers. 7,8 In this study, the Ni base superalloy Ren6 80 was investigated primarily at 871 ~ (1600 ~ and at 982 ~ (1800 ~ Ren6 80 is currently used as a turbine blade material in this temperature regime.

II. EXPERIMENTAL PROCEDURE A. Materials and Heat Treatment Test specimens of Ren6 80 were cast to shape. The composition (wt pct) is 3.0A1, 5.0Ti, 14.0Cr, 3.9W, 4.0Mo, 9.8Co, 0.17C, Ni bal. The heat treatment has been described elsewhere9 and results in a mixture of small spherical 7' particles and large V' cubes. After heat treatment, the specimens were machined by low stress grinding. B. Mechanical Testing The LCF behavior was investigated using longitudinal specimens at 871 and 982 ~ Testing was done using an MTS in the strain control mode. The cycle character was generally zero-tension-zero (R, = %,,/%,x ~ 0.05) and the test frequency was adjusted to maintain loading and unloading strain rates of either 0.5 pct/min or 50 pct/min. Some testing was done with a 90 s hold time at the maximum strain. Temperature was maintained at either 871 or 982 ~ using an RF generator and the maximum variation along the gage length was 3 ~C. C. Metallography

STEPHEN D. ANTOLOVICH is Professor of Materials Science, University of Cincinnati, Cincinnati OH 45221. S. LIU, formerly Research Associate at the University of Cincinnati, is presently Research Metallurgist, Special Metals, New Hartford, NY 13413. R. BAUR is Senior Engineer, General Electric Co., Aircraft Engine Group, Evendale OH 45215. Manuscript submitted February 15, 1980. METALLURGICAL TRANSACTIONS A

1. Optical. In order to examine microstructural features, as well as the relationship between the crack path and the microstructure, tested specimens were sectioned on a plane containing the stress axis and crack initiation site. 2. Scanning Electron Microscopy (SEM). The fracture surfaces were characterized by SEM. SEM was also used to investigate microstructural features at high magnifications. 3. Transmission Electron Microscopy (TEM). Thin wafers about 0.5 mm thick were cut perpendicular to the specimen axis about 2 to 3 mm from the fracture surface. The thickness of the wafer was reduced to 0.13 mm by mechanical polishing. TEM foils were then
VOLUME 12A, MARCH 1981--473

ISSN 0360-2133/81/0311-0473500.75/0 9 1981 AMERICAN SOCIETY FOR METALS AND THE METALLURGICAL SOCIETY OF AIME

10
871%

[]

o9

871~

o~
LU
Z

ld
101
UJ O Z
rr

<
rr Z m

th
KEY
STRAIN

<

cc I-,-

p03

rr

-2
STRAIN

O9 10 O m
I-O3

KEY TIME~D HOLD

RATE
%/MIN

(SEC)
0 90

O Fo9 /
13-

RATE HOLD TIME.


SEC

%/MIN 0.5
~o

0 [ 90
0
C]

II
9

<
J

5 50
J , L ,l,~L

9 ~}
i

9
, i ii,Li]
, t ,

l i l{

&. 10 ~

,,,,, 5

10

ld

J ll,,il

I llIIII

1@

1(~

10

10 a 10" 10 CYCLES TO INITIATION N i

CYCLES TO FAILURE NF Fig. 3--Coffin-Manson behavior of Rene 80 at 871 ~ Cycles to failure.

Fig. 1--Coffin-Manson behavior of Ren6 80 at 871 ~ initiation.

Cycles to
I I p I 9 i me

- ' ' ,

, - . , i

982b
9" \

BoA kko
(5 Z
10 ~
E~

982~;

uJ
z

_z

z
I.-.u? 0 I...._A
1~32
STRAIN

rr

CO
KEY HOLD TIME (SEC) 0 []
l

[]

STRAIN

KEY ~ HOLD rIME


(SEC) 0 9O

n []
X

RATE %/MIN O5

o3 <

90 9
, I I I , i~li i

g.

RATE 7~ /MIN 5

0
i , , ,,

9
,,I . . . . . . . . I

so i

10 2
1 0* lO

50
i

~ , I 1,11103

0a
CYCLES

10'

ld

1r
Cycles to

1@

TO INITIATION,

N,

Fig. 2--Coffin-Manson behavior of Ren~ 80 at 982 ~ initiation.

Cycles to

CYCLES TO FAILURE, Nf Fig. 4--Coffin-Manson behavior of Ren~ 80 at 982 ~ failure.

made by electropolishing using a Fischione twin-jet unit. They were subsequently examined in a JEM 200A transmission electron microscope operated at 200 Kv. An in situ stereographic projection of all foils was made to facilitate the excitation of desired reflections.

III. RESULTS AND DISCUSSION A. Mechanical Properties

D. Oxidation Studies Since environmental interactions have been previously demonstrated to be of prime importance for this alloy, 9 the oxidation kinetics were studied both by optical microscopy and by measuring the weight gained by a polished sample as it oxidized in air. The optical microscopy entailed sectioning a specimen that had been oxidized between 1 and 500 h and measuring the longest oxide spikes that were observed along grain boundaries. The other procedure was carried out using a microbalance connected to a strip chart recorder to provide a continuous record of weight gain vs time for temperatures in the range of interest.
4 7 4 - - V O L U M E 12A, M A R C H 1981

1. Tensile Properties. Tensile test results have been reported previously. 9 In general the strength decreases and ductility increases with increasing test temperature. Strain rate effects are small at 760 ~ but become pronounced at higher temperatures. Thus to investigate the effects of thermally activated deformation (i.e., creep) most testing was done above 760 ~ 2. Low Cycle Fatigue (LCF). The LCF test results are listed in Tables I and II and Coffin-Manson plots are shown in Figs. 1 to 4. Low rate and hold time tests exhibit a marked improvement in LCF compared to high rate continuous cycling. The reasons for this behavior are discussed in a subsequent section. Cumulative glide plots are shown in Figs. 5 and 6. In general, the material shows initial rapid softening followed by a regime in which the stress falls off less rapidly. This softening can result from fatiguing of a
METALLURGICAL TRANSACTIONS A

cold worked structure, 1~from shearing of precipitates I1 or from structural instabilities) 2 The reasons for the observed softening behavior will be discussed later in light of microstructural observations. Finally near the end of the test when cracks are present between the extensometer probes the stress again drops off very

rapidly. Crack initiation was defined by linearly extrapolating the midlife and end of life regimes of the cumulative glide plots and taking the point of intersection as N,. Cyclic stress strain curves are shown in Figs. 7 and 8. Figs. 7 and 8 show a reduction of stress amplitude with

Table I. LCF Oata for Ren6 80 at 871 ~


Plastic Strain Strain Range Specimen Temperature Rate Aep Number (~ (~ pct/min) pet 16LV12 16LV3 16LV15 16LV9 16V11 16LVI4 16LV13 16LV7 16LV21 16LV22 16LV23 16LV17 16LVi9 16LVI8 16LVI6 16LV24 871 871 871 871 871 871 871 871 871 871 871 871 871 871 871 871 50 45 0.50 50 0.50 50 50 29 80 85 91 50 0.50 0.50 50 0.50 loading 50 unloading 50 50 0.50 0.50 50 50 0.013 0.020 0.048 0.048 0.051 0.051 0.051 0.062 0.12 0.135 0.14 0.23 0.25 0.25 0.25 0.255 Total Strain Range Act pet 0.487 0.50 0.474 0.781 0.473 0.736 0.800 0.67 0.85 0.91 0.97 1.25 0.815 0.82 1.06 0.82 Stress Maximum Young's Range Stress Modulus Cycles to Cycles to Ao %~x EX10 -3 Initiation Failure (MPa) (MPa) (MPa) Nj Nf 771 755 650 1095 661 1138 1092 954 1172 1241 1241 1472 827 781 1360 965 586 731 340 677 344 658 732 732 627 662 669 775 416 378 710 428 158 163 139 140 154 141 144 158 154 161 148 147 147 158 160 172 9780 -1945 1244 2682 1417 936 -710 470 330 230 455 421 202 485 11760 2980 -1342 --1226 710 980 -496 --732 312 541

Comments Failed Failed outside extensometer Removed at N, Failed Failed, temperature loss causing failure, removed at 2742 Removed at N t Failed Failed Failed Removed at 591 Failed Removed at N i Removed at N s Stopped after 75 pet drop in load from mid-life Failed Failed

16HLV2 16HLV1 16HLV6 16HLV5 16HLV3 16HLV4

871 871 871 871 871 871

0.05 0.05 0.25 0.25 0.25 0.25

0.465 0.490 0.73 0.70 0.86 0.885

776 792 607 709 1015 1085

239 246 258 317 309 377

168 163 136 134 151 158

7440 7008 580 662 631 665

-12867 -929 1584 --

Tensile hold. Removed at N~ Tensile hold. Failed Tensile hold. Removed at N t Tensile hold. Failed Tensile hold. Stopped after 75 pet drop in load from mid-life Tensile hold. Removed at Nt

Table II. LCF Data lot Ren6 at 982 *C


Plastic Strain Strain Range Specimen Temperature Rate A% Number (~ (t pct/min) Pct 18LV1 18LV2 18LV3 18LV4 18LV8 18LV7 18LV5 18LV6 18HLV1 18HLV9 18HLV4 18HLV10 18HLV5 18HLV6 18HLV11 18HLV8 982 982 982 982 982 982 982 982 982 982 982 982 982 982 982 982 0.42 0.42 53.1 54.0 54.0 54.0 0.46 0.46 0.46 0.47 54.0 49.4 0.46 0.46 54.5 54.0 0.046 0.040 0.040 0.040 0.136 0.140 0.184 0.192 0.084 0.090 0.104 0.110 0.196 0.212 0.222 0.230 Total Strain Range Act Pet 0.282 0.300 0.468 0.476 0.624 0.624 0.532 0.552 0.332 0.342 0.476 0.436 0.532 0.532 0.628 0.624 Stress Maximum Young's Range Stress Modulus Cycles to Cycles to Ao Oraax EXI0 -3 Initiation Failure (MPa) (MPa) (MPa) N~ Nf 355 401 622 639 731 706 485 476 381 399 611 585 495 496 654 705 186 214 341 409 374 374 252 245 168 177 182 189 226 234 241 258 138 142 144 142 146 142 143 129 137 139 139 138 145 139 143 138 9720 4772 2119 1525 752 753 523 473 1437 1499 835 984 526 331 405 378 -5445 2511 --931 877 -1620 -1457 -573 --483

Comments Removed at Ni Failed Failed Removed at 1810 OX Removed at 848 OX Failed Failed Removed at 503 Failed OX Removed at N, Failed OX Removed at 984 Failed Removed at N~ Removed a t N t Failed

NOTE: OX indicates failure outside extensometer probes.

METALLURGICAL TRANSACTIONS A

VOLUME 12A, MARCH 1981--475

increasing temperature and decreasing strain rate. This observation will also be discussed later in terms of structural changes. 3. Effect of HoM Time. Just as the life increased with decreasing frequency for the continuous cycling tests, the effect of introducing a 90 s hold was to further increase the life and this increase was all the more pronounced as the plastic strain decreased. This is a further manifestation of the fact that a negative creep/ fatigue interaction is not operative for this alloy. The tensile hold tests showed longer lives than was observed for compressive holds. The tensile holds develop lower maximum stress than compressive holds and if maximum stress is an important factor in determining the life, such behavior would be expected.

It has been suggested in a previous paper 9 that a combination of response stress and environmental damage interact to define crack initiation. This hypothesis is investigated further in subsequent sections. B. Metallography

500

CO W qs FO3

250

1. Optical and SEM. The surface region of an LCF specimen is shown in Fig. 9 which is typical of what was observed for all test conditions: oxide spikes in the grain boundaries and a uniform wall of oxidation elsewhere. Such behavior was observed in a previous study 9 and the oxide spikes were usually the source of crack formation. However, for high rate hold time specimens creep voiding was observed in addition to surface cracking and this can be seen in Fig 10. Even though this new failure mode began to appear under these conditions, the general trends in life were not changed. It would be reasonable to assume that if the test conditions were made more extreme in terms of hold times and strain rates that the creep voiding could become a dominant failure mode but this was not the case in the studies reported here. Due to oxidation of the fracture surface features, detailed SEM fractography was not carried out. In

871~ c~
uJ

103

o ~4~ 9
t I t

~
400
CYCLE

::3
F--

,< 09 O3
!JJ

Y%--S~
KEY
STRAIN RATE HOLD TIME

800

I-CO

(SEC)
%IMIN

O O

90 9
III J J i gill

Fig. 5---Cumulative glide behavior of Rene 80 at 871 ~ for continous cycling at 0.50 pct/min.

102

5 50
, i i , iJiJi i , i ll,,i

D
J

15 3
I

1~ 2

161

10(

;J

STRAIN AMPLITUDE % Fig. 7--Cyclic stress/strain behavior of Rene 80 at 871 ~

9 9

STRESS RANGE MEAN STRESS

10a

982~

CO 03 W OZ I-CO

250

982 ~ C

.i t3

I-_J 13_

03

or" I-O3

102

STRAIN %/MIN 5 5O

KEY RATE HOLD O 0

TIME

ISEC} 90 I

I 4OO CYCLE

, i,i,I

i lllll

gill

8O0

1 0 -3

10-2 STRAIN AMPLITUDE

10-1
%

100

Fig. 6---Cumulative glide behavior of Ren6 80 at 982 ~ for continuous cycling at 0.50 pct/min. 476--VOLUME 12A, MARCH 1981

Fig. 8--Cyclic stress/strain behavior of Ren6 80 at 982 ~ METALLURGICAL TRANSACTIONS A

general, cracks initiated intergranularly with some evidence of transgranular propagation on the fracture surface. The final overload region had a characteristic dendritic appearance typical of these cast alloys. 2. TEM. A micrograph of the starting structure is shown in Fig. i 1. The structural changes as a function of test conditions were studied in detail, the results of which are summarized in Figs. 12 and 13. For all test conditions there was a tendency for the large 7' to coarsen and for the small 7' to disappear. For Ren6 77, in which the 7' is structurally similar to that in Ren6 80, it has been shown that coarsening of the 7' morphology takes place early in the fatigue process and requires plastic deformation to occur. ~3Such behavior is likely to be the case for Ren6 80. A comparison of Figs. 12 and 13 shows that the morphological changes in V' are most pronounced for high temperatures, loading rates and imposed hold times. At the temperatures and strain rates employed in this study, interfacial networks of near-edge dislocations are formed on the 7', similar to those observed elsewhere. 9,~4 The tendency for this to occur is again most developed for high temperatures, low loading rates and imposed hold times, Fig. 13(d). These observations are all

consistent with the observed cumulative glide and cyclic stress strain behavior. The continual drop off in stress with increasing cycles seen in Figs. 5 and 6 can be explained in terms of progressive coarsening. Similarly coarsening is more pronounced at high temperatures and low rates resulting in the cyclic stress/strain behavior seen in Figs. 7 and 8. It is difficult to imagine that the structural changes and development of the interracial dislocations seen in Figs. 12 and 13 can be damaging. In fact, as has been previously pointed out, 9 the coarsened particles are associated with increased ducility and the interracial dislocations reduce the matrix/precipitate strain energy. In view of the above observations, failure does not occur as a result of reaching a critical dislocation density and arrangement. Instead, the most damaging

Fig. 9---Oxide spike initiating surface crack. Specimen 16LV18, plastic strain 0.25 pct, life 732 cycles.

Fig. 10~Grain boundary voiding for high ramp rate hold time specimen. Specimen 16HLV1, plastic strain 0.05 pct, life 12867 cycles. METALLURGICAL TRANSACTIONS A

Fig. 11--Initial structure of Ren6 80 showing: (a) grain structure by light optical microscopy, (b) precipitate structure by dark field TEM, g = (100), (c) precipitate/dislocation structure by bright field TEM, g = (200). VOLUME 12A, MARCH 1981--477

event appears to be the d e v e l o p m e n t of oxide spikes along grain boundaries a n d interdendritic regions. Implications of this microstructural observation are used in the next section to suggest a possible model for high temperature L C F crack initiation. C. M o d e l D e v e l o p m e n t As a first order approximation, it will be a s s u m e d that as an oxide spike develops, at some point, u n d e r the action of the cyclic stress it loses c o h e r e n c y with the

base metal a n d b e c o m e s a well-defined crack.* Intui*Another possibility is penetration of oxygen or nitrogen along the boundaries which causes an embrittled region that eventually cracks. Such a mechanism is also qualitatively consistent with the following discussion. tively, the point of cracking depends on the size of the oxide spike a n d the m a g n i t u d e of the stress, which in turn depends on the underlying microstructure of the metal. The f o r m a t i o n of an oxide spike is a thermally activated process a n d in m a n y systems follows well-

(a)

(b)

(~ rd) Fig. 12--Precipitate/dislocation microstructure of Ren~ 80 continuously cycled at 871 ~ (a) Specimen 16LVI2 (low strain, high rate) g (200), (b) Specimen 16LVI6 (high strain, high rate) g ~ (200), (c) Specimen 16LV11 (low strain, low rate) g = (200), (d) Specimen 16LV18 (high strain, low rate) g = (111). See Table I for test conditions. 478--VOLUME 12A, MARCH 1981 METALLURGICALTRANSACTIONS A

k n o w n kinetics which m a y be expressed mathematically as;

observations can be represented b y the following equation:


t, = ,~ ( D t ) '~

z,

~ ~

[1]

[2]

where l i = spike depth at initiation, D = diffusion constant, t, = time to crack initiation, a n d ct = geometric constant. W e have verified that Eq. [1] is applicable to grain b o u n d a r y oxidation of Ren6 80 b y the optical microscopy experiments cited previously. T h e results of those

where m = 0.49 at 982 ~ a n d 0.46 at 760 ~ The data were also used to determine an a p p a r e n t activation energy for grain b o u n d a r y oxidation of 9.6 k c a l / m o l . Further verification of parabolic oxidation kinetics was o b t a i n e d f r o m the m i c r o b a l a n c e experiments. T h e resuits of these experiments s h o w e d that the weight gain

(a)

(b)

~O fd) Fig. 13--Precipitate/dislocation microstructure of Ren~ 80 subject to a 90 s hold at maximum strain during LCF at 982 ~ (a) Specimen 18HLV4 (low strain, high rate) g = (200), (b) Specimen 18HLV8 (high strain, high rate) g = (200), (c) Specimen 18HLV1 (low strain, low rate) g (200), (d) Specimen 18HLV5 (high strain, low rate) g = (200). METALLURGICAL TRANSACTIONS A VOLUME 12A, MARCH 1981--479

per unit area could be described by the following equation: AW _A _ = Ct]/2 [31

of more customary variables, Eq. [1] may be substituted into Eq. [5] and appropriate expressions may also be substituted for the maximum stress and time to initiation: [61

where A W = weight gain, A = area, C constant, and t = time. The constant depends on temperature and was measured as 7.6 X 10 -4 (gm/cm2).s 1/2 at 982 ~ and 4.9 X 10 -4 (gm/cm2)-s in at 928 ~ It has been shown elsewhere 15that cyclic deformation significantly increases the oxidation rate in Ren6 77, a microstructurally similar alloy. The effect of cyclic deformation on the form of the oxidation equation is not known for these systems. Lacking this information it seems reasonable to assume that the form of the oxidation process can be represented by Eq. [1]. Using these assumptions, relative oxide spike depths were calculated for purposes of developing a correlation. T o do this the spike depth was taken as unity for the test of shortest duration. The depth for any other test was simply found by taking the ratio of the test times as implied by Eq. [1]:
l, = ( t / t o ) 1/2
[4]

t, = ( l / .

+ th)N,

[7]

where d = mean stress, K -- fatigue strength coefficient, A% = plastic strain range, n' = cyclic strain hardening exponent, v = continuous cycle frequency, t h = hold time, and N, = cycles to initiation. Finally, from Figs. 14 and 15 the exponent p in Eq. [6] may be taken as 1/4 with the result:

+ K

Da2( 1

th) N ,

[8]

Temperature effects are accounted for explicitly in the diffusion constant: D = Doexp - Q/RT [9]

where t = test time and to = test time for reference specimen. Next, a crack initiation criterion of the form shown below was adopted as an operating hypothesis:

o ? . x . / e = Co

[51

where o m"x = maximum stress at initiation, Co = material constant, and p = exponent. Values of o, ~" were measured and plotted as a function of the computed l, values for both 871 and 982 ~ The results are shown in Figs. 14 and 15. The differences in constant reflect the fact that for each temperature the depth was taken as unity for the shortest test time. The results, to a high degree of confidence, reflect the applicability of Eq. [5] as an initiation criterion. It has been shown elsewhere 16J7 that Eq. [5] is also applicable to other materials and test conditions. To develop an expression for the fatigue life in terms

where the symbols have their usual meamng. Equation [8] represents an explicit evaluation of the correlations developed in Figs. 14 and 15. The increases in life that were observed for low frequencies and hold times can be explained in terms of structural coarsening. The structure is unstable relative to plastic deformation and the stress drops off significantly. This necessitates a large value of l, in Eq. [5] and a correspondingly large number of cycles. Of course a prediction, in the true sense of that word, must include the ability to predict the stress response as a function of test times and temperatures. This means that the structural changes must also be predicted as a function of these variables. Obviously such studies must be undertaken if the ideas advanced here are to be put in more fundamental terms. In all tests carried out for this study, the mean stress 8 was non-zero. However, in many instances stable systems are studied in fully reversed fatigue (# = 0). In this case, and taking a --~ 1 for convenience, Eq. [8] reduces to an expression of the form:

N, = C2 871 ~
lO 3~ ~[~

e x p ~o -Ac

,10,

CONTINUOUS CYCLING, Omax = 758 I,-23 cf:.95)

IS SR::: T RATE R A N I [ HO(L:EGTIME (SECI TIME [ 1%/MIN ~

%Lo

It is noteworthy that for a given temperature and frequency Eq. [lO] has the same form as Coffin's frequency modified fatigue life with an explicit expression for the frequency and temperature dependence. Furthermore, the exponent fl in the Coffin-Manson equation is given by:
1

Oma==75791~ -29 If: 94)


10~ 10 0 RELATIVE . . . . . 101 OXIDE . DEPTH iI 10 2 i i I i .... 1(

fl-

8n'

[11]

AT INITIATION

Fig. 1 4 - - M a x i m u m stress at initiation as a function of oxide depth at initiation for Renr 80 tested at 871 ~ T h e equations of each line along with the coefficients of determination are given. 4 8 0 - - V O L U M E 12A, M A R C H 1981

It is interesting to compare the qualitative predictions of Eq. [10] with what is frequently observed for stable materials tested in fully reversed fatigue at high temperatures. For example, a decrease in life is predicted for low frequencies, long hold times and high temperMETALLURGICAL TRANSACTIONS A

103
D~
CON3 LNUOUS r ~ CYC lING

982 ~
STRAIN ] RATE 9B5) %/MIN O5 c t m a x = 3 8 5 9 I, -21 ( | =

KEY
HOLD 0 TIME 90 ,. (SEC)

ALL DATA
~:~ Omax:38291, -25 ~t: 93)

t tli,i

I IIII[

I I I l

10 ~

101

102

103

RELATIVE OXIDE DEPTH AT INITIATION Fig. 15--Maximum stress at initiation as a function of oxide depth at initiation for Ren6 80 tested at 982 ~ The equations of each line along with the coefficients of determination are given.

specimens rapidly loaded to maximum strain was any internal void formation seen. Even in these cases there were also surface initiation sites. 4) The metallographic observations were used to develop an intuitive correlation between maximum stress and depth of oxidation at the time of crack initiation. All data were well-represented by this correlation. 5) The correlation demonstrated above was used to develop an expression for the crack initiation life. The expression has the form of the Coffin-Manson law for a given set of experimental conditions. In addition to cyclic stress/strain parameters frequency, hold time, and temperature are explicitly incorporated.

ACKNOWLEDGMENT atures and such behavior is, in fact, often observed, especially for the conditions enumerated above. Finally, values of n' in the neighborhood of 0.15 to 0.20 are frequently observed and values of the CoffinManson exponent in the range 0.6 to 0.8 are predicted which again agrees with observations. At this point we would consider the model to be a promising initial approach to understanding high temperature LCF behavior in certain Ni base superalloys. Providing a more fundamental basis will require detailed studies of dynamic oxidation and coarsening phenomena. Based on the above discussion, it is clear one cannot conclude that accumulated deformation damage is the sole cause of LCF cracking merely because the data can be represented by a Coffin-Manson plot. Such a representation is also consistent with cracking by environmental damage. Identification of specific damage mechanisms would involve investigations of the temperature and frequency dependence of the fatigue life as well as detailed metallographic studies. CONCLUSIONS 1) The fatigue life of Ren6 80 at both 871 and 982 ~ increases with decreased frequency and imposition of a 90 s hold time. Such behavior is at variance with the concept of a negative creep/fatigue interaction. 2) The ?' precipitate in Ren6 80 coarsened and developed an interracial array of edge dislocations with very few dislocations in the matrix. Such changes do not constitute damage in any fundamental sense. 3) Cracks initiated at oxide spikes in surface connected grain boundaries. Only in the case of hold time The authors wish to express their appreciation to the Air Force Office of Scientific Research for financial support of this project (AFOSR Grant No. 76-2952), and to Dr. A. Rosenstein, Program Manager, for his very useful assistance.

REFERENCES
1. S. S. Manson: ASTM STP 520, 1973, p. 744. 2. L. F. Coffin: Proc. Second lnternational Conf. on Mech. Behavior of Materials, p. 866, ASM, Metals Park, OH, 1976. 3. S. Majumdar and P. S. Maiya: Proc. A S M E - M P C Symposium on Creep-Fatigue Interaction, MPC-3, p. 323, ASME, N.Y., 1976. 4. W. Ostergren: J. Test. Eval., 1976, vol. 4,p. 327. 5. A GARD Conference Proceedings, AGARD-CP-243, 1978. 6. L. F. Coffin, S. S. Manson, A. E. Carden, and L. K. Severud: ORNL Report # 5073, Jan. 1977. 7. M. Gell and G. R. Leverant: ASTM STP 520, 1973, p. 37. 8. Stephen D. Antolovich: La Fatigue des materiaux et des structures, C. Bathias and J. P. Bailon, eds., p. 465, Maloine S. A., Paris, 1980. 9. Stephen D. Antolovich, P. Domas, and J. L. Strudel: Met. Trans A, 1979, vol. 10A, p. 1859. 10. C. E. Feltner and C. Laird: Acta. Metall., 1967, vol. 15, p. 1621. 11. R. E. Stoltz and A. G. Pineau: J. Mater. Sci. Eng., 1978, vol. 31, p. 275. 12. D. Fournier and A. Pineau: Met. TransA, 1977, vol. 8A, p. 1095. 13. E. Rosa: M. S. Thesis, University of Cincinnati, 1978. 14. A. Lasalmonie and J. L. Strudel: Philos. Mag., 1975, vol. 32, p. 937. 15. Stephen D. Antolovich, Eduardo Rosa, and Andr6 Pineau: Mater. Sci. Eng., in press. 16. Stephen D. Antolovich and N. Jayaraman: Paper presented at the 27th Army Materials Research ConferEnce, Lake George, NY, July, 1980. Paper in press in the conference proceedings. 17. Stephen D. Antolovich, R. Baur, and S. Liu: Superalloys 1980, J. K. Tien, S. T. Wlodek, H. Morrow III, M. Gell, and G. E. Maurer, eds., p. 605, ASM, 1980.

METALLURGICAL TRANSACTIONS A

VOLUME 12A, MARCH 1981--481

Вам также может понравиться