Anda di halaman 1dari 9

Corrosion fatigue behaviour of a 15Cr6Ni precipitation-hardening

stainless steel in different tempers


C . - K . L I N a n d W. - J. T S A I
Department of Mechanical Engineering, National Central University, Chung-Li 32054, Taiwan
Received in final form 17 March 2000

A B S T R A C T Systematic fatigue experiments, including both high-cycle axial fatigue (SN curves)

and fatigue crack growth (FCG, da/dNDK curves), were performed on a precipitationhardening martensitic stainless steel in laboratory air and 3.5 wt% NaCl solution.
Specimens were prepared in three tempers, i.e. solution-annealed (SA), peak-aged
(H900) and overaged (H1150) conditions, to characterize the effects of ageing treatment
on the corrosion fatigue (CF) resistance. SN results indicated that fatigue resistance
in all three tempers was dramatically reduced by the aqueous sodium chloride environment. In addition, the smooth-surface specimens in H900 temper exhibited longer CF
lives than the H1150 ones, while those in SA condition stood in between. However,
for precracked specimens, the H1150 temper provided superior corrosive FCG resistance than the other two tempers. Comparison of the SN and FCG curves indicated
that early growth of crack-like defects and short cracks played the major role in
determining the CF life for smooth surface. The differences in the CF strengths for
the SN specimens of the given three tempers were primarily due to their inherent
differences in resistance to small crack growth, as they were in the air environment.
Keywords corrosion fatigue; precipitation-hardening martensitic stainless steel; ageing
treatment; crack-like defect; small crack growth.

INTRODUCTION

Precipitation-hardening martensitic stainless steels


(PHMSSs) have been widely used as structural components in a variety of applications in aircraft, chemical,
naval, nuclear and other industries due to their high
strength and toughness, good fabrication characteristics
and good corrosion resistance. Of the former, 17-4 PH
stainless steel is presently one of the most often used
alloys. It is usually obtained from the mill in the solutionannealed condition, with a microstructure consisting of
low-carbon equiaxed martensite and 510 vol% d-ferrite
stringers.1,2 By subsequent age-hardening treatment in
the temperature range of 482621 C, a wide range of
mechanical properties can be obtained.1,35 Peak values
of strength and hardness are obtained after ageing at
482 C (900 F), where precipitation of coherent copperrich precipitates occurs.1 Ageing at higher temperatures
(above 540 C) would result in the precipitation of
incoherent large copper-rich precipitates, lower strength
and hardness, and improvement in toughness and stress

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

corrosion cracking (SCC) resistance.69 A newer


15Cr6Ni PHMSS, Custom 450, which is a second
generation of the 17-4 PH alloy type, was produced to
avoid the formation of d-ferrite in the martensitic matrix
and improve the through-thickness properties, in particular, for heavy sections.1
Many premature service failures of PHMSS components have been reported to be caused by environmentassisted cracking.913 Investigations on corrosion resistance of 17-4 PH alloy, heat treated to different conditions, have been reported in the literature.2,69 The
emphasis in these investigations has been on the effects
of ageing treatment on the SCC resistance. In general,
for PHMSSs, the overaged condition is reported to have
higher SCC resistance than peak-aged condition and the
SCC resistance is increased with an increase in ageing
temperature.8 However, the effects of ageing treatment
on the corrosion fatigue (CF) behaviour of PHMSSs
have not been studied in much detail. More research
is necessary to characterize the effects of ageing treatment on the CF resistance in chloride environment

489

490

C . - K . L I N a n d W. - J . T S A I

for PHMSSs due to their extensive use in many components of primary load-carrying structures in such
environments.
The limited number of previous investigations4,11,1416
discussing CF behaviour of 17-4 PH steels were mainly
based on either high-cycle fatigue or fatigue crack growth
(FCG) experimental results. It was not possible to distinguish from these results the environmental effect on
the stages of initial and Stage II crack growth separately,
for a given alloy and ageing condition. This current
research was therefore planned to characterize the
environmental influence on the development of fatigue
damage for a 15Cr6Ni PHMSS, Custom 450, in an
aerated 3.5 wt% NaCl solution by systematic experiments, including both high-cycle axial fatigue (SN
curves) and FCG (da/dNDK curves). Three heat treatments, solution-annealed (SA), peak-aged (H900) and
overaged (H1150), were applied to assess whether the
overageing temper would improve the CF resistance as
effectively as it did for the SCC resistance. The results
obtained in salt water were then compared to those
obtained in laboratory air to characterize the effects of
the corrosive environment on the reduction of fatigue
resistance in the given alloys.
EXPERIMENTAL PROCEDURES

The 15Cr6Ni PHMSS used in the current study was


Custom 450 supplied by the vendor (Carpenter
Technology Corporation, Reading, PA, USA) in the
form of hot-rolled solution-annealed bars of 12.7 or
88.9 mm diameter. The small and large round bars were
used to fabricate high-cycle axial fatigue and FCG test
specimens, respectively. The chemical composition of
this alloy is given in Table 1. Three heat treatments were
applied in the present work, i.e. unaged SA, peak-aged
H900 and overaged H1150. Details of the heat treatment
procedures are given in Table 2. The mechanical properTable 1 Nominal chemical composition of Custom 450 stainless
steel (wt%)
Cr

Ni

14.79 6.47

Cu

Mo

Si

Mn

Fe

1.50

0.77

0.46

0.37

0.02

0.021

0.004 bal.

Table 2 Heat treatment conditions of Custom 450 stainless steels


tested
Type

Treatment

SA
H900
H1150

1038 C (1900 F)1 hair cool


SA482 C (900 F)4 hair cool
SA621 C (1150 F4 hair cool

ties for each temper are listed in Table 3. The monotonic


tensile properties were obtained by using tensile specimens having a cylindrical gauge section of 6 mm in
diameter and 25 mm in length. Both high-cycle axial
fatigue and FCG tests were carried out in laboratory air
and an aerated 3.5 wt% NaCl solution at room temperature on specimens machined parallel with the rolling
direction. High-cycle axial fatigue tests were conducted
as per ASTM E466 on axial smooth-surface specimens,
with a cylindrical gauge section of 6 mm in diameter
and 18 mm in length, to determine the stresslife (SN)
curves. FCG experiments were performed in accordance
with ASTM E647 on 6.35-mm-thick compact tension
(CT) specimens to determine the da/dNDK relationship. The salt water solution had a pH value of 7.2 both
before and after the tests. All CF tests were conducted
under freely corroding conditions, i.e. no external potential was applied to the specimens. Details of the specimen
geometry and experimental set-up for such CF tests
were described elsewhere.17
All fatigue tests were conducted on a closed-loop,
servohydraulic machine under a sinusoidal loading wave
form with a load ratio of R=0.1 and a frequency of f=
20 Hz. The high-cycle axial fatigue tests were run to
failure or to 2106 cycles where specimen was considered to be a runout. All CT specimens were first
fatigue precracked in laboratory air before testing in the
corrosive environment. The crack length and crack
closure level in the FCG tests was determined by the
compliance technique recommended by ASTM E647
using a clip gauge mounted on the front edge of the CT
specimen to monitor the crack-mouth-opening displacement during testing. Characterizations of the fracture
surface morphology were obtained by scanning electron
microscopy (SEM).
RESULTS AND DISCUSSION

Environmental effect on the fatigue behaviour


Figure 1 was a plot of the fatigue life as a function of
stress amplitude for axial smooth-surface specimens of
each temper tested in both laboratory air and 3.5 wt%
NaCl solution. It can be seen in Fig. 1 that the aqueous
sodium chloride environment indeed generated deleterious effects on the fatigue resistance of smooth surfaces
in each temper as evidenced by the remarkable reduction
in fatigue life. For each temper, the differences in the
fatigue lives between the two environments became
larger as the fatigue life increased. The ratios of the
fatigue strength at 106 cycles in 3.5 wt% NaCl solution
to the atmospheric fatigue strength were 0.89, 0.85 and
0.87 for tempers SA, H900 and H1150, respectively.
This implies that H900 temper was slightly more sensi-

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

C O R R O S I O N FAT I G U E B E H AV I O U R O F A P R E C I P I TAT I O N - H A R D E N I N G S TA I N L E S S S T E E L

Fig. 1 Comparison of SN curves under various environments for


Custom 450 stainless steel in different tempers: (a) SA; (b) H900;
and (c) H1150. (Arrows designate runout tests.)

tive to the environmental effects in terms of SN behaviour, as compared to the other two tempers. The
counterpart results of the FCG tests using precracked

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

491

CT specimens were plotted as (da/dN) versus (DK) in


Fig. 2. The fatigue crack growth rates (FCGRs) of long
cracks in salt water were not significantly different from
those in air for each temper except at DK>20 MPa m1/2
for H900 temper. Apparently, the corrosive environment
did not generate as severe damage on the FCG resistance
of precracked CT specimen as it did in the case of
smooth-surface specimen. For example, salt water, in
comparison to air, could reduce the fatigue life of the
smooth surface in H900 temper by an order of magnitude
[Fig. 1(b)], while it only increased the FCGR of the
precracked specimen in the same temper by a maximum
factor of two and a half [Fig. 2(b)]. It was initially
assumed that the less deleterious effects of the aggressive
environment on the FCGR of long cracks might be
attributed in part to a greater crack closure effect induced
by corrosion products. However, no significant difference in the crack closure level between the CT specimens
tested in air and salt water was detected for the given
three tempers, except for SA temper at low DK regions
where enhanced crack closure effects have been found
in salt water. Apparently, for the given cyclic loading
conditions ( f =20 Hz and R=0.1), the given salt water
did not generate the same damageing effects on the long
crack growth behaviour as it did on the SN behaviour.
In particular, the aggressive environment did not exert
any detrimental influence on the long crack growth in
H1150 temper as shown in Fig. 2(c).
The above comparisons suggest that the corrosive
environment exerted more detrimental influence on the
initiated fatigue crack of microstructural dimensions
(Stage I cracking) than on the phase of Stage II crack
growth for the given alloys. It has been similarly recognized by Miller and Akid18 that the fatigue lifetime of
plain specimens for several steels in both an inert air
and an aggressive environment was controlled by the
initial growth of crack-like defects or short cracks of microstructural dimensions. Several processes may
occur during the CF lifetime, i.e. surface film breakdown, stress-assisted pitting or preferential dissolution,
environmentally assisted Stage I followed by Stage II
crack growth, and crack coalescence. However, the predominant processes determining the CF lifetime are
those that are strongly dependent on the combined
action of cyclic stresses and the aggressive environment
which together often lead to strain localization and
enhanced growth of defects/short cracks; these include
the pitting, preferential dissolution, Stage I and Stage Ito-Stage II crack growth processes.18
In the case of steels which show a fatigue limit in air,
it has been postulated that the initiated fatigue crack of
microstructural dimensions can not overcome the relevant dominant microstructural barrier to continue
growth.18 It was generally assumed that pits and/or

492

C . - K . L I N a n d W. - J . T S A I

localized corrosion of emerging slip steps or extrusion,


somehow induced by the CF process, prematurely
initiate fatigue cracking or reduce the applied stress
required to initiate fatigue cracks.1820 Research21 has
shown that surface film breakdown and pit development
is the first phase of crack-like defect growth. Apparently,
in the current study, the generation of crack-like defects
followed by their enhanced growth in salt water is likely
to occur at an earlier stage compared to that under the
atmospheric conditions thereby resulting in a decrease
in the fatigue lifetime for axial smooth-surface specimens. This can be supported by the SEM fractography
observations.
Fatigue fracture surface morphology near the crack
initiation sites in the axial smooth-surface specimens of
SA and H1150 tempers tested in salt water at long-life
regime is exemplified in Fig. 3. It can be seen in Fig. 3
that the CF fracture of this smooth-surface specimen
originated from a corrosion-induced surface defect which
was absent in the specimens tested in air. Not only can
the corrosion-induced surface defect serve as a stress
concentration site to become a Stage I crack, but its
enhanced growth under the synergism between the
aggressive environment and the cyclic stresses can also
reduce the transition period from Stage I to Stage II
crack growth. However, when the crack length was large
enough to overcome the dominant microstructural barrier, the environmental effects became less important.18
This is supported by the experimental evidences presented in Fig. 1 where differences in lifetime between
the air and salt water were reduced with an increase in
the applied stress levels and a decrease in the lifetime
and test duration, and in Fig. 2 which shows little
differences in the FCGRs of long cracks under the
atmospheric and aqueous sodium chloride environments.
It is therefore proposed that the fatigue lifetimes of the
SN specimens tested in 3.5 wt% NaCl solution, as
compared to the atmospheric environment, were significantly reduced by the faster initial growth of crack-like
defects and shorter transition period from Stage I to
Stage II cracking, in particular at the lower applied
stress levels.
Note the visible corrosion-induced surface defects
serving as the fatigue fracture origins for the cases of SA
and H1150 specimens could not be clearly identified for
H900 ones, as shown in Fig. 4. It may be noted in Fig. 4
that the H900 specimen failed from a subsurface
inclusion very close to the free surface. The subsurface
inclusions are also favourable sites for premature fatigue
cracking under corrosive environments as they are
microstructural stress concentrations which would lead
to strain localization and enhance the preferential dissolution or film rupture at the surfaces near them.18
This might explain why the fatigue lives in H900 speciFig. 2 Comparison of FCGR curves under various environments
for Custom 450 stainless steel in different tempers: (a) SA;
(b) H900; and (c) H1150.

C O R R O S I O N FAT I G U E B E H AV I O U R O F A P R E C I P I TAT I O N - H A R D E N I N G S TA I N L E S S S T E E L

(a)

(a)

(b)

(b)

Fig. 3 SEM fractograph of a Custom 450-SA smooth-surface axial


fatigue specimen tested in 3.5 wt% NaCl solution: (a) stable crack
growth region; and (b) fracture origin (i: crack initiation site).

mens were still significantly reduced by the corrosive


environment even though no visible-sized surface defects
were identified at the fatigue fracture origins. Two
previous studies2,22 consistently reported that the pitting
resistance for a comparable PHMSS, 17-4 PH, in an
acidic chloride medium took the following sequence:
peak-aged>overaged>solution-annealed. The pitting
resistance of the 17-4 PH stainless steel was found to be
improved when copper was present as a fine, coherent

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

493

Fig. 4 SEM fractograph of a Custom 450-H900 smooth-surface


axial fatigue specimen tested in 3.5 wt% NaCl solution: (a) stable
crack growth region; and (b) fracture origin (i: crack initiation site).

precipitate form (peak-aged temper) rather than as a


coarse incoherent precipitate form (overaged temper) or
in solid solution (unaged SA temper).22 In peak-aged
condition, repassivation is likely to occur at the pit sites
as a result of formation of a protective copper oxide film
and the pitting resistance is improved compared to SA
and overaged tempers.2 As the reformed austenite in
overaged microstructures is rich in austenite stabilizing
elements, e.g. copper and nickel in solid solution form,

494

C . - K . L I N a n d W. - J . T S A I

the autopassivation effect is thus diminished.22 This


might explain why the corrosion-induced surface defects
were visible at the fracture initiation sites in SA and
H1150 SN specimens (Fig. 3), but similar-sized surface
defects could not be detected for H900 ones (Fig. 4).
Comparison of CF resistance in different tempers
Comparisons of the SN curves in different tempers
under the air and salt water environments were shown
in Fig. 5. Figure 5 indicated in both air and salt water
that peak-aged H900 temper exhibited the highest
fatigue strength of smooth surface among the given
three tempers, while the overaged H1150 temper showed
the lowest one. Apparently, the overaged temper did not
provide superior CF resistance of smooth surface than
the peak-aged temper, as it generally did in the SCC
case for PHMSSs.8 As shown in Fig. 5, the fatigue
strength of axial smooth-surface specimen tested in both

Fig. 5 Comparison of SN curves for differently heat-treated


Custom 450 stainless steels tested in: (a) air; and (b) salt water.
(Arrows designate runout tests.)

air and 3.5 wt% NaCl solution appeared to be increased


with the monotonic tensile strength and hardness of the
given alloys. That is, the tensile strength, hardness and
fatigue strength in smooth surface for the given three
tempers takes the following order: H900>SA>H1150.
Comparison of the FCGRs of long cracks for various
tempers in saline solution was presented in Fig. 6. As
shown in Fig. 6, the given PHMSS exhibited the highest
and lowest FCGRs in H900 and H1150 tempers, respectively, while the SA condition stood in between. The
distinctly superior resistance to long crack growth possessed by H1150 over H900 temper in salt water might
be partially attributed to the greater toughness generated
in H1150 temper. An earlier study by Rack and Kalish4
similarly indicated that the fracture toughness of a
comparable PHMSS, 17-4 PH, exhibited a distinct minimum at peak-aged H900 temper leading to higher
FCGRs in both dry argon and water-saturated argon, as
compared to an overaged H1100 condition. The variation of FCGRs of long cracks in different tempers may
also be related to the nature of the slip mode. In peakaged microstructures which contain predominantly
shearable, coherent precipitates, the dominant mode of
deformation is generally one of planar slip as the dislocations can cut through the particles.23 On the other
hand, the non-shearability of the incoherent precipitates
in overaged microstructures is believed to result in
dislocations either looping or bypassing particles and a
diffuse/wavy slip pattern.23 This dislocation looping of
non-shearable precipitates can effectively remove strain
localization resulting in lower driving force for crack
extension. In addition, for precipitation-hardening alloys,

Fig. 6 Comparison of FCGR curves for differently heat-treated


Custom 450 stainless steels tested in salt water.

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

C O R R O S I O N FAT I G U E B E H AV I O U R O F A P R E C I P I TAT I O N - H A R D E N I N G S TA I N L E S S S T E E L

the uniformly distributed incoherent precipitates in overaged microstructures would be expected to act as potent
hydrogen trapping sites innocuously trapping hydrogen
in the matrix, lowering the concentration of hydrogen
at the crack tip, and thus reducing deleterious hydrogen
embrittlement (HE) effects on the FCG behaviour.24,25
On the other hand, a coarse and planar slip pattern
would enhance the susceptibility to HE as considerable
hydrogen transported by mobile dislocations could reach
the crack tip areas due to a larger extent of dislocation
transport of hydrogen.24,25 These differences in resistance to HE can be supported by Fig. 2 indicating that
the FCGRs of long cracks in both air and salt water for
the H1150 temper were almost the same, while larger
FCGRs at certain DK ranges in saline solution were
observed for H900 and SA tempers in comparison with
the atmospheric environment.
As shown in Fig. 5, smooth-surface H900 specimens
had the longest CF lives in comparison to SA and H1150
ones. However, the FCG data presented in Fig. 6 indicated that H900 temper produced the fastest Stage II
FCGRs in salt water for the precracked CT specimens,
as compared to the other two heat-treated conditions.
Comparisons made in Figs 5 and 6, again, indicate that
the SN behaviour in 3.5 wt% NaCl solution was predominated by the early growth of crack-like defects or
short cracks to a Stage II crack for the given PHMSSs.
Apparently, H900 temper provided a better resistance to
the combined deleterious effects from the applied cyclic
stresses and the aggressive environment on the initial
growth of crack-like defects, as compared to SA and
H1150 tempers. However, the superiority of the H900
to H1150 and SA tempers in corrosive SN behaviour
may be essentially attributed to its greater inherent
resistance to small crack growth. By comparing the
fatigue strengths at 106 cycles in air and salt water for
each temper (see Fig. 1), it can be found that the amounts
of reduction in fatigue strength at 106 cycles caused by
the corrosive environment were 11%, 15% and 13% for
SA, H900 and H1150 tempers, respectively. Apparently,
the extent of the detrimental effects from the corrosive
environment on the SN behaviour was slightly increased
in the sequence of tempers SA, H1150 and H900.
Although the SN specimens in H900 temper were
slightly more sensitive to the environmental effects than
the SA ones, they still last much longer to fail than did
SA ones in salt water. That is, the corrosive environment
did not significantly change the relative superiority of
the given three tempers in resistance to small crack
growth and accordingly the trend of SN behaviour.
This is supported by Fig. 5(a) and (b) which shows that
the trends of the relative differences of the given three
tempers in the SN behaviour were not significantly
different between the air and 3.5 wt% NaCl solution

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

495

environments except that the SN curves were all shifted


toward lower stress levels in salt water. Therefore, the
differences in the CF strengths of the given three
tempers were primarily due to their inherent differences
in resistance to small crack growth, as they were in the
air environment.
Murakami and Endo26 have observed that the dependence of the threshold stress intensity range for small
crack growth, DKth,SC , on the materials hardness and
the existing flaw size (area) for a wide variety of metals
could be expressed by the following unified equation:
DKth,SC =3.3103 (Hv +120)(area)1/3

(1)

where DKth,SC is in MPa m1/2 , Hv is the materials


Vickers hardness, and area is in mm and determined
by projecting defects or cracks onto the plane normal to
the maximum tensile stress. The fatigue limit, Se , for a
specimen with existing defects was thus derived as:26
Se =1.43(Hv +120)/(area)1/6

(2)

where Se is the nominal stress and has units of MPa.


Note that both Eqs (1) and (2) were derived for a stress
ratio of R=1. It may be noted that the materials
having higher Vickers hardness would show higher
DKth,SC and accordingly higher fatigue strength.
As the flaw sizes at the fatigue fracture origins for the
specimens tested in air could not be clearly identified
from the SEM fractographs, only those tested in salt
water were used to check whether the trend of the SN
behaviour observed in the current study could be predicted by the model of Murakami and Endo.26 The SEM
fractographs for the failed SN specimens having the
longest CF lives in each temper were selected to estimate
the flaw sizes (including surface defects and subsurface
defects around inclusions) at the fracture origins using
the method outlined by Murakami and Endo.26 In this
regard, the values of area for the observed defects in
the corrosive specimens were determined as 22, 18 and
25 mm for SA, H900 and H1150 tempers, respectively.
Substituting these flaw sizes together with the hardness
values given in Table 3 into Eq. (2) yields the estimated
fatigue limits at R=1 in salt water as 342, 397 and
311 MPa for SA, H900 and H1150 tempers, respectively.
By taking a mean stress correction factor for the fatigue
limit as [(1R)/2]a, where a=0.226+Hv 104 , the
predicted CF limits at R=0.1 for SA, H900 and H1150
tempers would be 280, 326 and 255 MPa, respectively.27
Similarly, based on these flaw sizes, the DKth,SC values
in salt water were determined by Eq. (1) as 3.70, 3.88
and 3.59 MPa m1/2 for SA, H900 and H1150 tempers,
respectively. According to these predictions, the thresholds for small crack growth and fatigue strengths in the
corrosive environment were expected to take the follow-

496

C . - K . L I N a n d W. - J . T S A I

Table 3 Mechanical properties of Custom 450 stainless steel in different tempers

Temper

Ultimate tensile
strength (MPa)

Yield strength
(MPa)

Elastic modulus
(GPa)

Elongation
(in 25 mm) (%)

Vickers hardness
(Hv )

V-notch impact
toughness (J)

SA
H900
H1150

1065
1405
1015

1041
1354
678

190
202
141

12.2
15.3
23.2

280
329
252

133
54
132

ing trend: H900>SA>H1150. The experimental SN


behaviour shown in Fig. 5 indeed exhibited this trend in
both air and 3.5 wt% NaCl aqueous solution.
As shown in Fig. 5, the observed fatigue strengths in
salt water at 106 cycles with a stress ratio of R=0.1 were
~315, 363 and 260 MPa for SA, H900 and H1150
tempers, respectively. The observed CF strengths at 106
cycles for R=0.1 were apparently greater than the
predictions by Murakami and Endos model.26,27 The
underestimate of the CF strengths at 106 cycles was due
to the fact that the predictions were made based on the
larger flaw sizes observed in the failed specimens with
CF lives shorter than 106 cycles. Nevertheless, the CF
strengths of the SN specimens for the given three
tempers did exhibit a trend as predicted by the small
crack growth resistance model of Murakami and Endo.26
In addition, if the observed fatigue strengths at 106
cycles in the air environment were adopted in a reverse
process using Murakami and Endos model to estimate
the largest flaw sizes which would not grow to failure,
the expected critical flaw sizes were found to be 7.2, 5.1
and 13 mm for SA, H900 and H1150 tempers, respectively. As the observed flaw sizes causing the fatigue
failures of SN specimens in the aqueous sodium chloride
environment were significantly greater than these estimated critical flaw sizes by a factor of ~two or three,
the corrosive environment indeed resulted in premature,
initiated growth of defects and substantially reduced the
fatigue lives in each temper. Therefore, the reduction of
fatigue strength in each temper by the corrosive environment and the trend of the differences of the given three
tempers in CF strengths were predictable by the model
of Murakami and Endo.26 This implies that the differences in the corrosive SN behaviour for the given three
tempers mainly resulted from their inherent differences
in resistance to small crack growth; i.e. the thresholds
for small crack growth and fatigue strengths in the
corrosive environment apparently took the following
trend: H900>SA>H1150.
From the above discussion, it is obvious that for the
given stainless steel, peak-aged treatment producing the
highest tensile strength and hardness also generated
the highest fatigue strength in the atmospheric and
aqueous sodium chloride environments due to the
enhanced resistance to small crack growth. It can be

seen in Figs 3 and 4 that in 3.5 wt% NaCl solution the


fatigue cracks essentially grew by a transgranular mode
regardless of heat-treated condition. This is different
from the cracking pattern in SCC where intergranular
crack growth path is generally seen in PHMSSs.69 It
has been reported69 that the SCC resistance in PHMSSs
can be improved by overageing at a higher temperature
as compared to the peak-aged temper. However, ageing
at a higher temperature from H900 to H1150 temper
to increase the SCC resistance might not guarantee an
equivalent improvement in the CF resistance for the
given PHMSSs simply because the cracking pattern
shifts from intergranular to transgranular mode for CF
loading. The favourable CF strength in smooth surface
of peak-aged material seems to arise from its greater
resistance to initial fatigue crack growth despite its lower
resistance to Stage II crack growth in 3.5 wt% NaCl
solution, as compared to unaged and overaged
conditions.
CONCLUSIONS

1 Comparison of the SN curves and FCGR curves of


long cracks in both air and 3.5 wt% NaCl solution
for a 15Cr6Ni PHMSS in different tempers indicated
that the corrosive environment exerted more influence
on the SN behaviour than on the Stage II crack
growth.
2 Heat-treating a PHMSS to a peak-aged temper would
generate the highest CF strength of smooth surface
in salt water while to an overaged temper would result
in the lowest FCGR of Stage II crack growth in the
same aggressive environment.
3 For the given alloys tested in salt water, the CF lives
of smooth-surface specimens were predominated by
the early growth of defects or short cracks rather than
by the Stage II crack growth. This is inferred from
the fact that peak-aged temper exhibited higher FCGR
of Stage II crack growth yet produced longer fatigue
lifetime of smooth surface as compared with the
overaged and solution-annealed tempers.
4 The superiority of the peak-aged temper to solutionannealed and overaged tempers in corrosive SN
behaviour may be essentially attributed to its greater
inherent resistance to small crack growth, as similar

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

C O R R O S I O N FAT I G U E B E H AV I O U R O F A P R E C I P I TAT I O N - H A R D E N I N G S TA I N L E S S S T E E L

trends of the relative differences in SN behaviour


among the given three tempers were observed in the
air and aqueous sodium chloride environments.

14

Acknowledgements

15

This work was supported by the Taiwan Power Company


and managed by the National Science Council of the
Republic of China (Taiwan) under Contract nos NSC87-TPC-E-008-004 and NSC-88-TPC-E-008-002.

16

REFERENCES

17

1 W. F. Smith (1993) Structure and Properties of Engineering Alloys,


2nd edition, McGraw-Hill, New York.
2 U. K. Mudali, A. K. Bhaduri and J. B. Gnanamoorthy (1990)
Corrosion behavior of 17-4 PH stainless steel. Mater. Sci. Tech.
6, 475481.
3 F. B. Pickering (1976) Physical metallurgy of stainless steel
developments. Int. Metals Rev. 21, 227268.
4 H. J. Rack and D. Kalish (1974) The strength, fracture toughness, and low cycle fatigue behavior of 17-4 PH stainless steel.
Metall. Trans. 5, 15951605.
5 R. D. K. Mishra, G. Y. Prasad, T. V. Balasurbramanian and
P. R. Rao (1987) On variation of impact toughness in 17-4
precipitation hardened stainless steel. Scripta Metall. 21,
10671070.
6 K. S. Raja and K. P. Rao (1993) On the hardness criterion for
stress corrosion cracking resistance of 17-4 PH stainless steel.
J. Mater. Sci. Lett. 12, 963966.
7 D. L. Dull and L. Raymond (1973) A test procedure to evaluate
the relative susceptibility of materials to stress corrosion cracking. Corrosion 29, 205212.
8 J. E. Truman (1981) Stress-corrosion cracking of martensitic
and ferritic stainless steels. Int. Metals Rev. 26, 301349.
9 C. S. Carter, D. G. Farwick, A. M. Ross and J. M. Uchida
(1971) Stress corrosion properties of high strength precipitation
hardening stainless steels. Corrosion 27, 190197.
10 F. J. Heymann, V. P. Swaminathan and J. W. Cunningham
(1981) Steam turbine blades: considerations in design and a
survey of blade failure. Report no. CS-1967, Electrical Power
Research Institute, Palo Alto, CA, USA.
11 B. C. Syrett, R. Viswanathan, S. S. Wing and J. E. Wittig
(1982) Effect of microstructure on pitting and corrosion fatigue
of 17-4 PH turbine blade steel in chloride environments.
Corrosion 38, 273282.
12 V. P. Swaminathan and J. W. Cunningham (1983) Correlations
between LP blade failures and steam turbine characteristics. In:
Corrosion Fatigue of Steam Turbine Blade Materials (Edited by
R. I. Jaffee), Pergamon Press, New York, pp. 3.13.16.
13 R. M. Thompson, G. B. Kohut, D. R. Canfield and W. R. Bass

2000 Blackwell Science Ltd. Fatigue Fract Engng Mater Struct 23, 489497

18

19

20
21

22
23

24

25

26

27

497

(1991) Sulfide stress cracking failures of 12Cr and 17-4 PH


stainless-steel wellhead equipment. Corrosion 47, 216220.
M. U. Islam, G. Campbell and R. Hsu (1990) Fatigue and
tensile properties of EB welded 17-4 PH steel. Welding J.
68, 4550.
L. E. Willertz, T. M. Rust and V. P. Swaminathan (1983) High
cycle corrosion fatigue of some steam turbine blade alloys. In:
Corrosion Fatigue of Steam Turbine Blade Materials (Edited by
R. I. Jaffee), Pergamon. Press, New York, pp. 3.753.106.
T. M. Rust and V. P. Swaminathan (1983) Corrosion fatigue
testing of steam turbine blading alloys. In: Corrosion Fatigue of
Steam Turbine Blade Materials (Edited by R. I. Jaffee), Pergamon.
Press, New York, pp. 3.1073.129.
C.-K. Lin and S.-T. Yang (1998) Corrosion fatigue behavior of
7050 aluminum alloys in different tempers. Engng Fracture
Mech. 59, 779795.
K. J. Miller and R. Akid (1996) The application of microstructural fracture mechanics to various metal surface states. Proc. R.
Soc. Lond. A 452, 14111432.
C. Laird and D. J. Duquette (1972) Mechanisms of fatigue
crack nucleation. In: Corrosion Fatigue: Chemistry, Mechanics and
Microstructure (Edited by O. Devereux, A. J. McEvily and R. W.
Staehle), National Association of Corrosion Engineers,
Houston, pp. 88117.
D. J. McAdam (1926) Stressstrain relationships and the corrosion-fatigue of metals. Proc. ASTM 26, 224254.
R. Akid, Y. Z. Wang and U. S. Fernando (1993) The influence
of loading mode and environment on short fatigue crack growth
in a high strength steel. In: CorrosionDeformation Interactions
(Edited by T. Magnin and J. M. Gras), Les Editions de Physique,
Les Ulis, France, pp. 659670.
K. S. Raja and K. P. Pao (1995) Pitting behavior of type
17-4 PH stainless steel weldments. Corrosion 51, 586592.
E. A. Starke Jr and G. Lutjering (1979) Cyclic plastic deformation and microstructure. In: Fatigue and Microstructure
(Edited by M. Meshii), American Society for Metals, Metal
Parks, OH, pp. 205243.
P. Munn and B. Andersson (1990) Hydrogen embrittlement of
PH 138Mo steel in simulated real-life tests and slow strain
rate tests. Corrosion 46, 286295.
G. M. Pressouyre and I. M. Bernstein (1981) An example of
the effect of hydrogen trapping on hydrogen embrittlement.
Metall. Trans. A 12, 835844.
Y. Murakami and E. Endo (1986) Effects of hardness and crack
geometries on DKth of small cracks emanating from small
defects. In: The Behavior of Short Fatigue Cracks (Edited by K.
J. Miller and E. R. de los Rios), Mechanical Engineering
Publications. London, pp. 275293.
Y. Murakami, Y. Uemura, Y. Natsume and S. Miyakawa (1990)
Effect of mean stress on the fatigue strength of high-strength
steels containing small defects in nonmetallic inclusions. Trans.
Jpn. Soc. Mech. Engng 56, 1074.

Anda mungkin juga menyukai