You are on page 1of 8

International Journal of Fatigue xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

International Journal of Fatigue


journal homepage: www.elsevier.com/locate/ijfatigue

Fatigue crack growth in the heat affected zone of a hydraulic turbine


runner weld
Alexandre Trudel a,b, Michel Sabourin a, Martin Lévesque b, Myriam Brochu b,⇑
a
Global Technology Center in Sustainable Hydro, ALSTOM Renewable Power Canada Inc., Sorel-Tracy, Canada
b
Department of Mechanical Engineering, École Polytechnique de Montréal, Montreal, Canada

a r t i c l e i n f o a b s t r a c t

Article history: The fatigue crack growth behavior of a CA6NM weld heat affected zone (HAZ) was investigated. Fatigue
Received 7 November 2013 crack growth tests in river water environment were conducted on as-welded and post-weld heat treated
Received in revised form 7 March 2014 specimens at load ratios R = 0.1 and R = 0.7. For a fully open crack, i.e. at R = 0.7, the HAZ fatigue behavior
Accepted 11 March 2014
was similar to that of the base metal. When crack closure occurred, i.e. at R = 0.1, the HAZ showed a lower
Available online xxxx
near threshold crack growth resistance. The post-weld heat treatment was beneficial at R = 0.1 by reliev-
ing tensile residual stresses, while its effect was negligible at R = 0.7. The crack trajectory was influenced
Keywords:
by the weld’s yield strength mismatch which promoted deviation towards the soft material (base metal).
Stainless steels
Heat affected zone
The main conclusion of this study is that the HAZ does not represent a weak link when its high load ratio
Fatigue crack growth fatigue behavior is considered. This confirms the validity of currently used fatigue assessment methods of
Crack closure hydraulic turbine runners.
Residual stresses Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction growth threshold and Paris’ relation constants. These parameters


are well known for materials typically used to manufacture turbine
Hydraulic turbine runners are one of the most critical parts in runners such as martensitic stainless steel CA6NM [2,3].
hydroelectric power plants. Their main role is to convert the ki- However, experience has shown that fatigue cracks in turbine
netic energy of flowing water into mechanical energy, which is runners can grow through and along weld heat affected zones
then transmitted to the generator that converts it to useful electri- (HAZ), which can have a heterogeneous microstructure and a gra-
cal energy. Turbine runners are typically engineered for service dient of mechanical and fatigue properties. The HAZ of many dif-
lifetimes of 70 years. During service, these components are sub- ferent materials have been shown to exhibit a different fatigue
jected to important cyclic loads generated from transient operation crack growth behavior than their base metal counterpart. Weld-in-
and complex hydraulic phenomena, such as high frequency pres- duced residual stresses are frequently considered to explain such
sure fluctuations caused by rotor–stator interactions [1]. These differences. Elevated fatigue crack growth rates are often measured
cyclic loads are taken into account in fatigue analyses relying on [4–6] when tensile residual stresses develop in the HAZ vicinity,
fracture mechanics concepts. A damage tolerance approach is nec- while compressive residual stresses promote crack closure, which
essary to account for discontinuities present in turbine runners. is beneficial [7,8]. Furthermore, microstructural effects are often
These discontinuities are found in the form of inevitable casting invoked to explain different fatigue crack growth characteristics
and welding defects from which fatigue cracks can grow. In addi- in the HAZ, as this zone often has characteristic microstructural
tion, the runner’s blades are often welded to their support with features. Grain coarsening and/or refinement, as well as formation
partial penetration T-joints, resulting in an un-welded portion that of secondary phases such as reformed austenite in the case of
is treated as a crack-like defect. In these regards, the main purposes CA6NM [9] often occur in the HAZ. These microstructural features
of the fatigue analyses are to establish the maximum allowable de- can affect the fatigue crack growth behavior of the HAZ by influ-
fect size, within detection limits, as well as to size the un-welded encing the extent of, inter alia, roughness-induced crack closure
ligament of the partial penetration joints. These analyses rely on [10], crack path deflection [11,12], secondary fracture mechanisms
experimentally determined parameters such as the fatigue crack occurrences such as intergranular fracture [13] and crack tip plas-
tic zone processes such as transformation-induced plasticity
[14,15]. Moreover, previous studies have shown that the degree
⇑ Corresponding author. of yield strength mismatch between filler metal, HAZ and base
E-mail address: myriam.brochu@polymtl.ca (M. Brochu).

http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
0142-1123/Ó 2014 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
2 A. Trudel et al. / International Journal of Fatigue xxx (2014) xxx–xxx

Nomenclature

B compact tension specimen thickness W compact tension specimen width


C intercept of Paris’ relation
CE intercept of Elber’s relation Acronyms
da/dN fatigue crack growth rate in mm/cycle AW as-welded
Kcl stress intensity factor at closure BM base metal
Kmin minimum stress intensity factor FM filler metal
Kmax maximum stress intensity factor HAZ heat affected zone
DK stress intensity factor range (Kmax  Kmin) HT heat treated
DKeff effective stress intensity factor range (Kmax  Kcl) PWHT post-weld heat treatment
DKth threshold stress intensity factor range TRIP transformation-induced plasticity
m slope of Paris’ relation
mE slope of Elber’s relation
R Load ratio (Kmin/Kmax)

metal can influence the crack trajectory [16–18]. This can indi- 2.2. Microhardness measurements
rectly affect the fatigue crack growth behavior if the crack deviates
from HAZ sub-regions of different fatigue crack growth resistances Vickers microhardness profiles were performed across the HAZ,
[19]. Additionally, welded hydraulic turbine runners always under- starting in the filler metal and ending in the base metal in accor-
go a post-weld heat treatment at typically 600 °C, which is a man- dance with ASTM E384 [27]. The measurements were taken per-
datory manufacturing step as it softens hard and brittle as-welded pendicularly to the welding direction on as-welded and heat
constituents proven to deteriorate the weld’s fracture toughness treated specimens. The microhardness was measured at every
[20]. This tempering treatment also relieves the residual stresses 100 lm on a length of 12 mm. For each measurement, a force of
known to form in the vicinity of CA6NM welds [21–23] and pro- 100 gf was applied for 15 s. These measurements were useful to
motes the formation of reformed austenite [24]. estimate the size of the HAZ, which is confined between the fusion
Investigations concerned with the fatigue crack growth behav- line and base metal.
ior in the HAZ of CA6NM welds are sparse [25] and come short
of providing a detailed explanation of influencing factors such as
previously mentioned. This study aims to fulfill this shortcoming 2.3. Fatigue crack growth rate testing
by thoroughly characterizing the fatigue crack growth behavior
of the HAZ that develops in CA6NM and 410NiMo welds. The gen- As-welded (AW) and heat treated (HT) compact tension (CT)
eral objective is to identify the main influencing factors in order to specimens were machined from the welded plates to a width of
provide a satisfactory explanation of the characteristic HAZ fatigue W = 50.8 mm and a thickness of B = 12.7 mm in accordance with
crack growth behavior. Ultimately, this study provides insight ASTM E647 [28]. The starter notch was positioned in the heat af-
regarding the fact that the HAZ represents a weak link in terms fected zone (HAZ), parallel to the welding direction and close to
of fatigue crack growth within hydraulic turbine runners, or not. the fusion line, as shown in Fig. 1b. Fatigue crack growth tests were
To reach these objectives, fatigue crack growth tests in aqueous realized as per standard ASTM E647 under load control with a
environment were realized in order to establish the fatigue thresh- 100 kN servo-hydraulic machine and an automated custom pro-
old and Paris’ relation constants at load ratios of R = 0.1 and R = 0.7. gram. Load ratios of R = 0.1 and R = 0.7 and a constant frequency
As-welded and post-weld heat treated specimens were tested in of 20 Hz were used. During the tests, the crack was immersed in
order to study the effect of post-weld heat treatment on the HAZ room temperature synthesized water simulating waters from
fatigue crack growth resistance. The results are analyzed in terms Outardes River, which drives three major hydroelectric power
of residual stresses relief and crack closure mechanisms. A discus- plants in the province of Quebec, Canada. The mineral composition
sion on the weld strength mismatch is also provided to explain the of the water used can be found in [9]. A K-decreasing procedure as
observed gradual crack deviation from the HAZ to the base metal. per standard ASTM E647 was applied to reach the threshold stress
intensity factor range (DKth), which was established for growth
rates close to 108 mm/cycle. This was followed by a K-increasing
2. Experimental procedures procedure up to a stress intensity factor range (DK) that corre-
sponded to a crack length of 40 mm (0.8 W). Crack length was
2.1. Materials and specimen preparation monitored using the compliance method as per standard ASTM
E647 with a crack mouth clip gauge. Optical measurements on
The materials used in this study are martensitic stainless steel both sides of the crack were periodically performed to calibrate
alloy CA6NM and filler metal of matching chemical composition and validate the compliance method. The linear portion of the fati-
410NiMo. A fully automated flux-cored arc welding (FCAW) pro- gue crack growth curves obtained were fitted with a Paris type
cess was used to deposit a 40 mm thick layer of filler metal (FM) relation, da/dN = CDKm, where da/dN is the growth rate in mm/
on the surface of a 50 mm thick rectangular base metal (BM) cycle and C and m are material specific constants to be determined
CA6NM plate (Fig. 1a). The chemical composition of the CA6NM al- from the experimental data. Crack closure was assessed with crack
loy used can be found in Table 1. The mechanical properties of the opening displacement and load data. A 2 % compliance offset crite-
CA6NM alloy used [2] along with typical values for 410NiMo [26] rion was used to determine the stress intensity factor at closure
are shown in Table 2. Th4e welded plate was cut lengthwise in (Kcl), as suggested in standard ASTM E647. This allowed for
two equal parts (Fig. 1a), one of which was subjected to a 600 °C the determination of the effective stress intensity factor range
post-weld heat treatment (PWHT) for two hours, then air cooled defined as DKeff = Kmax  Kcl, where Kmax is the maximum applied
down to room temperature. stress intensity factor.

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
A. Trudel et al. / International Journal of Fatigue xxx (2014) xxx–xxx 3

Welding direction
Cut plane
FM 410NiMo

FM
40 mm
410NiMo
fusion line
HA Z
50 mm BM
CA6NM

BM CA6NM

(a) (b)
Fig. 1. (a) Fatigue specimen sampling layout from welded plates and (b) fatigue specimen showing the notch location in the HAZ.

Table 1
Chemical composition of BM CA6NM (wt%) [2]. (a) 400
FM
Material C Mn Si S P Cr Ni Mo
380
CA6NM 0.02 0.66 0.59 0.008 0.031 13.04 4.07 0.53 HAZ
360 BM

Microhardness (HV)
Table 2 340
Mechanical properties.

Material Yield Tensile Young’s Elongation 320


strength (MPa) strength (MPa) modulus
CA6NM [2] 763 837 206 GPa 27.0%
300
410NiMo [26] 849 925 – – Soft region

280
As−Welded
3. Results
260
−4 −2 0 2 4 6 8
3.1. Microhardness
Distance from fusion line (mm)
Fig. 2a and b respectively show the Vickers microhardness pro-
files obtained for the as-welded and heat treated conditions. The (b) 400
PWHT significantly reduced the hardness of the filler metal from
a mean microhardness of 365 HV to 320 HV. However, the microh- 380
ardness of the base metal was not affected by PWHT and main-
tained a mean microhardness of 290 HV. Fig. 2 shows that for
Microhardness (HV)

360
both as-welded and heat treated conditions, the HAZ was a transi- FM
tion region in which the microhardness decreased, starting from HAZ
340
the fusion line, over a distance of approximately 4.5 mm.
BM
It is worth mentioning that, just before reaching the microhard-
ness lowest plateau characterizing the base metal, a slightly softer 320
region was identified and considered within the HAZ. This soften-
ing can be attributed to the decomposition of martensite in stable 300
austenite during welding [21].
280
Soft region
3.2. Fatigue testing results Heat Treated
260
3.2.1. Fatigue behavior of the as-welded HAZ −4 −2 0 2 4 6 8
Fig. 3a shows the fatigue crack growth rates da/dN against DK Distance from fusion line (mm)
for the as-welded HAZ tested at R = 0.1 and 0.7. Results show sim-
Fig. 2. Vickers microhardness profiles measured across the HAZ, from the FM to the
ilar fatigue crack growth rates for both load ratios up to a DK value
BM in the (a) as-welded condition and (b) heat treated condition.
of 5.0 MPa m1/2. However, for DK values above 5.0 MPa m1/2, the
growth rates measured at R = 0.1 are lower than those measured
at R = 0.7 and the growth rates differences increase with DK. The The occurrence of closure in the specimen tested at R = 0.1 corre-
closure stress intensity factor was also determined and is reported lates with the lower crack growth rates measured, when compar-
on a graph of Kcl/Kmax against DK in Fig. 3b. Comparing the closure ing with the crack growth rates obtained at R = 0.7 without closure.
evolution at R = 0.7 and R = 0.1 reveals that while the crack is fully In ductile materials, such as stainless steel, a certain amount of
open (Kcl/Kmax = 0) at R = 0.7 for all DK, the specimen tested at plasticity-induced crack closure is expected, especially at low load
R = 0.1 is subjected to closure for DK values above 5 MPa m1/2. ratios like R = 0.1 and at low DK, in the threshold regime [29].

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
4 A. Trudel et al. / International Journal of Fatigue xxx (2014) xxx–xxx

tensile residual stresses were present in the as-welded specimens


(a)
−4
10
AW HAZ, R = 0.1 and were responsible for the absence of crack closure, and hence
AW HAZ, R = 0.7 higher growth rates at R = 0.1. It was also shown in [9] that crack
growth caused the relaxation of the crack tip tensile residual stres-
ses (30% reduction for a 10 mm crack extension), which would ex-
plain the fact that closure becomes significant as the crack grows
−5
10 and DK increases at R = 0.1. Moreover, a significant amount of re-
tained austenite was previously reported in the HAZ [9] and the
strain-induced transformation of austenite into martensite at the
da/dN (mm/cycle)

tip of a growing fatigue crack in CA6NM was demonstrated in


[24]. This transformation-induced plasticity (TRIP) is known to
accentuate crack closure with increasing DK, where greater cyclic
−6
10 plastic zone sizes result in a greater amount of austenite trans-
forming into martensite. The TRIP effect could have contributed
to the increasing closure level with increasing DK. Additional work
is however needed to clearly characterize the influence of the TRIP
effect on the fatigue crack growth resistance of CA6NM welds HAZ.
−7
10
3.2.2. Fatigue behavior of the heat treated HAZ
Fig. 4a shows the fatigue crack growth rates against DK for the
heat treated HAZ tested at R = 0.1 and R = 0.7. Fig. 4b shows the clo-
sure level evolution in the form of Kcl/Kmax against DK. A load ratio
effect can be clearly observed as the growth rates are significantly
−8
10 lower at R = 0.1 than at R = 0.7 for all DK values. This behavior can
1 10 50 be correlated with the significant closure levels measured at R = 0.1
ΔK (MPa√m) and the corresponding absence of closure at R = 0.7. The Kcl/Kmax
ratio at R = 0.1 decreases gradually with increasing values of DK,
(b) 1 AW HAZ, R = 0.1
which is an expected behavior.
AW HAZ, R = 0.7
3.2.3. Fatigue behavior of the HAZ with respect to the base metal
0.8 Fig. 5 shows the growth rates against DK for both as-welded
and heat treated HAZ, and both load ratios. The fatigue crack
growth curves of the base metal, which were obtained in [2,3]
0.6 for the same base metal and experimental parameters, are pre-
K /Kmax

sented for R = 0.1 and R = 0.7 as lines.


cl

0.4 Behav ior at R ¼ 0:7


It can be seen from the black filled symbols that the growth
rates measured in the as-welded and heat treated HAZ at R = 0.7
0.2
are similar and comparable to the R = 0.7 base metal characteristic
curve. In the near threshold regime, the HAZ is characterized by
0 slightly higher crack growth rates than the base metal. In a previ-
ous study, higher growth rates were measured at constant DK of
0 5 10 15 20 8 MPa m1/2 and 20 MPa m1/2 when comparing the HAZ to the base
metal for a fully open crack [9]. As was explained and modeled, the
ΔK (MPa√m) finer martensitic microstructure of the HAZ led to a less tortuous
Fig. 3. Fatigue testing results obtained in the as-welded HAZ. (a) Growth rates and
crack path than in the base metal, which reduced toughening by lo-
(b) closure level (Kcl/Kmax) against DK for R = 0.1 and R = 0.7. cal mixed modes of crack advance. Furthermore, the growth rate
difference between the HAZ and base metal vanishes at higher
DK. This could be explained by the fact that significant macro-
Roughness and oxide-induced crack closure can also be significant scopic crack deviation was observed throughout the tests, where
for steels [30,31]. On the other hand, at high load ratios and/or high the crack deviated away from the nominal crack growth plane to-
DK, closure usually becomes less significant as crack tip opening wards the base metal side. This behavior and its effect on the mea-
displacements increase. The results obtained with the specimen sured growth rates are further discussed in Section 4.2.
tested at R = 0.1 in the near threshold regime contradicts this typ-
Behav ior at R ¼ 0:1
ical behavior since the crack was closure free at DK values below
5 MPa m1/2 and closure developed at higher DK values. The ab- At R = 0.1, the heat treated HAZ (empty inverted triangles)
sence of closure during fatigue crack growth testing of steel welds shows a fatigue behavior typical of the base metal behavior at
at low R values was previously observed and explained by the pres- R = 0.1. This is correlated with the significant crack closure mea-
ence of tensile residual stresses [4–6]. This explanation can be ap- sured in this specimen (Fig. 4b). On the other hand, the crack
plied to our specimen since the crack initially propagated closure growth rates measured in the as-welded HAZ (empty squares) at
free at near threshold levels. Furthermore, though residual stresses low DK values are comparable to the base metal behavior tested
were not measured in this study, it was shown in [9] that crack tip at R = 0.7. This behavior correlates with the absence of closure pre-
tensile residual stresses of 250 MPa were present in specimens viously reported for this specimen for DK values below 5 MPa m1/2
produced from the same welded plate. It is therefore deducted that (Fig. 3b).

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
A. Trudel et al. / International Journal of Fatigue xxx (2014) xxx–xxx 5

(a) 10−4 10
−4

HT HAZ, R = 0.1 AW HAZ, R = 0.1


HT HAZ, R = 0.7 HT HAZ, R = 0.1
AW HAZ, R = 0.7
HT HAZ, R = 0.7
BM, R = 0.1
−5 BM, R = 0.7
10−5 10

da/dN (mm/cycle)
da/dN (mm/cycle)

−6
10−6 10

−7
10−7 10

−8
10−8 10
1 10 50 1 10 50

ΔK (MPa√m) ΔK (MPa √m)


Fig. 5. Growth rates against DK at R = 0.1 and 0.7 along with the base metal R = 0.1
(b) 1 HT HAZ, R = 0.1 and R = 0.7 curves.
HT HAZ, R = 0.7

0.8 Considering that significant residual stresses are present in the


as-welded state, this behavior is in agreement with the residual
stresses and crack closure evolution as explained in Section 3.2.1.
0.6
Kcl /Kmax

When tensile residual stresses act to fully open the crack, the spec-
imen adopts a crack growth behavior comparable to the base metal
tested at R = 0.7. On the other hand, when closure is fully devel-
0.4
oped following the residual stress relief with crack advance, the
crack growth behavior is closer to the R = 0.1 base metal curve.
0.2 Hence, the apparently lower fatigue crack growth curve slope m
of this case is in fact a transition from an effective high load ratio
behavior caused by tensile residual stresses to the nominal low
0 load ratio behavior.
The results discussed are summarized in Table 3 where DKth
0 5 10 15 20 values and Paris’ relation constants are reported for every test. In
ΔK (MPa√m) the case of the as-welded HAZ tested at R = 0.1, two distinct rela-
tions are proposed to account for the closure transition previously
Fig. 4. Fatigue testing results obtained in the heat treated HAZ. (a) Growth rates discussed; one for DK values below 5 MPa m1/2 (before closure
and (b) closure level (Kcl/Kmax) against DK for R = 0.1 and R = 0.7. transition) and a second for DK values above 14 MPa m1/2 (after
closure transition). It can be seen that, before the closure transi-
In the near threshold regime of the heat treated HAZ, higher tion, the slope m correlates well with the slopes obtained from
growth rates and a lower threshold are found, when compared the tests conducted at R = 0.7, while after the closure transition,
with the base metal R = 0.1 behavior. This may be explained by the slope is increased and is comparable to the slope obtained at
the microstructural effects reported in [9]. At low load ratios, like R = 0.1 in the heat treated HAZ.
R = 0.1, a less tortuous crack path can lead to a reduced contribu-
tion of roughness-induced crack closure. However, this behavior
is in contradiction with the fact that the heat treated HAZ shows 4. Discussion
a behavior comparable to the R = 0.1 characteristic curve of the
base metal for higher DK values. The gradual crack deviation to- 4.1. Effect of PWHT on HAZ fatigue behavior
wards the base metal, which was also observed in the specimens
tested at R = 0.1, might offer an explanation and is further dis- In Fig. 6, the growth rates obtained from the fatigue tests were
cussed in Section 4.2. plotted against the effective stress intensity factor range (DKeff) to
As for the as-welded specimen, whereas the measured growth isolate the effect of crack closure. It can be seen that the results ob-
rates are similar to the R = 0.7 characteristic base metal curve at tained for all specimens fall onto a single curve for which the El-
low DK values, the growth rates gradually merge towards the ber’s relation constants (da=dN ¼ C E DK m eff ) are shown based on all
E

R = 0.1 base metal curve for DK values higher than 5.0 MPa m1/2, data. This result implies that, in a closure free situation, the load

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
6 A. Trudel et al. / International Journal of Fatigue xxx (2014) xxx–xxx

Table 3
Fatigue threshold and Paris’ relation constants for AW and HT HAZ tested at R = 0.1 and R = 0.7. Coefficients of determination (R2) are also shown to show goodness of fit.

R = 0.1 R = 0.7
AW HT BM [2] AW HT BM [3]
Before transition After transition
DKth (MPa m1/2) 1.99 3.51 4.50 1.87 1.95 2.07
C 7.80  109 1.67  109 1.08  109 9.86  1010 1.08  108 1.04  108 5.51  109
m 2.87 3.14 3.33 3.34 2.78 2.74 3.04
R2 0.9904 0.9936 0.9930 – 0.9991 0.9971 –

−4 deviation of the order of 80 lm/mm towards the base metal was


10
AW HAZ, R = 0.1 observed in the as-welded and heat treated HAZ tested at both load
HT HAZ, R = 0.1 ratios. This crack deviation might have affected the measured
AW HAZ, R = 0.7 growth rates since it was proven experimentally in [9] that the
HT HAZ, R = 0.7 HAZ crack growth resistance varies from fusion line towards base
metal. Fig. 7 shows the crack growth curves calculated from the
−5
10 experimentally determined parameters of Table 3, along with the
highest growth rates measured in the HAZ at constant DK values
of 8 MPa m1/2 and 20 MPa m1/2 taken from [9]. This shows that
da/dN (mm/cycle)

the growth rates measured in this study are lower than the highest
growth rates measured in our previous study. This can be attrib-
−6
uted to the fact that the crack deviated from HAZ sub-regions of
10 slightly lower resistance to fatigue crack growth towards the base
metal, which was identified in [9] as having the highest resistance
to crack growth because of its coarser microstructure.
This gradual deflection of the crack towards base metal can be
explained by considering the yield strength mismatch between
−7 the filler metal and base metal as shown in Table 2 by their differ-
10 da/dN = 1.02⋅10−8(ΔKeff)2.80 ent respective yield strengths. In yield strength mismatched welds,
R2 = 0.9916 an asymmetric plastic zone forms which causes crack deviation to-
wards the side of lower yield strength [16–18]. In order to confirm
that the crack did deviate from the higher yield strength HAZ (hard
region) near filler metal towards the lower yield strength base me-
−8
10 tal (soft region), post-fatigue testing Vickers microhardness maps
1 10 50 were realized on a surface encompassing the fatigue crack of the
ΔKeff (MPa√m) as-welded and heat treated fatigue specimens tested at R = 0.1
(Fig. 8). For the as-welded HAZ (Fig. 8a), the crack, initially in a re-
Fig. 6. Fatigue crack growth rates against DKeff for as-welded and heat treated HAZ gion of approximately 375 HV gradually deviated towards a softer
tested at R = 0.1 and 0.7. region having a microhardness of approximately 320 HV. Similarly,
for the heat treated HAZ (Fig. 8b), the crack gradually deviated

ratio and the PWHT do not significantly affect the crack growth
resistance of CA6NM weld HAZ.
−4
At R = 0.1, the closure measurements showed that the crack was 10
fully open in the as-welded HAZ at low DK (DK = DKeff), which has Highest da/dN
measured in HAZ [9]
been attributed to crack tip tensile residual stresses. This suggests HT HAZ, R = 0.1
that the PWHT is beneficial to the crack growth resistance only in AW HAZ, R = 0.1
HT HAZ, R = 0.7
particular conditions for which tensile residual stresses prevent
da/dN (mm/cycle)

AW HAZ, R = 0.7
crack closure. With reduced residual stresses after PWHT, the crack
is allowed to close when growing at low R values such as R = 0.1.
Similar behavior has been observed in many steel welds, including
−5
in the fusion zone of a CA6NM weld [4,25]. 10
As for the results obtained at R = 0.7, Fig. 6 shows that the fati-
gue crack growth driving force, DKeff, was equal to the applied DK
for both as-welded and heat treated specimens. For high load ratios
such as R = 0.7, fatigue cracks are typically fully open even in the
absence of tensile residual stresses [32]. Therefore, the effect of
PWHT is negligible at R = 0.7.
−6
10
8 10 20
4.2. Effect of macroscopic crack path deviation on crack growth
characterization ΔK (MPa√m)
Fig. 7. Fatigue crack growth rates against DK for as-welded and heat treated HAZ
Despite the fact that considerable efforts were deployed to en- tested at R = 0.7 and R = 0.1 along with the highest growth rates measured in the
sure that the fatigue crack would grow in the HAZ, gradual crack HAZ taken from [9].

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
A. Trudel et al. / International Journal of Fatigue xxx (2014) xxx–xxx 7

(a)

Vickers Microhardness
FM

HAZ

BM

HAZ
boundaries

Crack path

10 mm

(b)

FM

HAZ

BM

Fig. 8. Vickers microhardness maps around the crack of (a) as-welded and (b) heat treated specimens tested at R = 0.1 showing the crack receding from hard material and
gradually deviating towards soft material.

from a high microhardness region of approximately 330 HV to- post-weld heat treated specimens were tested at R = 0.1 and 0.7
wards a lower microhardness region of approximately 290 HV. in order to study the effect of load ratio and PWHT on the HAZ
These findings show that it is the crack tip asymmetric plastic- resistance to crack growth. From the results, the following conclu-
ity that drives the macroscopic crack trajectory in the case of crack sions can be drawn:
tip yield mismatch. On the other hand, it is often extrinsic factors,
such as crack closure, that drive the resistance to crack growth.  The HAZ was found to have a fatigue behavior similar to the
These two phenomena are independent, where the former occurs base metal when the crack was fully open. This was the case
in the plastic zone ahead of the crack tip, while the latter relates for the as-welded and heat treated HAZ tested at R = 0.7 as
to the contact of fracture surfaces behind the crack tip. Conse- well as the as-welded HAZ tested at R = 0.1 which was sub-
quently, a fatigue crack can indeed deviate towards a zone of high- jected to tensile residual stresses at crack tip. On the other
er resistance to crack growth in the vicinity of mismatched welds. hand, the heat treated HAZ tested at R = 0.1 was subjected
This is the case of the present study, where the crack deviated to- to closure and showed a near threshold crack growth resis-
wards the most fatigue resistant constituent (base metal). tance inferior to the base metal R = 0.1 fatigue behavior. At
higher DK, its fatigue behavior was comparable to the base
metal.
5. Conclusion  When corrected for closure by plotting da/dN against DKeff,
the fatigue crack growth rates of the as-welded and heat
The objective of this study was to characterize the fatigue crack treated HAZ tested at R = 0.1 and 0.7 fell onto a single curve.
growth behavior of a CA6NM and 410NiMo weld HAZ, typical of This indicates that load ratio and PWHT affected the mea-
what can be found in hydraulic turbine runners. As-welded and sured crack growth rates only by modifying the extent of

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006
8 A. Trudel et al. / International Journal of Fatigue xxx (2014) xxx–xxx

crack closure. The PWHT was beneficial to the fatigue crack [6] Vaidya WV, Horstmann M, Ventzke V, Petrovski B, Kocak M, Kocik R, et al.
Structure–property investigations on a laser beam welded dissimilar joint of
growth resistance of the HAZ tested at R = 0.1 by relieving
aluminium AA6056 and titanium Ti6Al4V for aeronautical applications. Part II:
detrimental tensile residual stresses that prevented crack Resistance to fatigue crack propagation and fracture. Materialwissenschaft und
closure. However, since the growth rates were similar Werkstofftechnik 2009;40:769–79.
before and after PWHT, the PWHT had no effect on the [7] Itatani M, Fukakura J, Asano M, Kikuchi M, Chujo N. Fatigue crack growth
behavior of weld heat-affected zone of type 304 stainless steel in high
HAZ fatigue behavior for a load ratios of R = 0.7, where clo- temperature water. Nucl Eng Des 1994;153:27–34.
sure was not significant. [8] Pouget G, Reynolds AP. Residual stress and microstructure effects on fatigue
 For all tested specimens, the crack growth plane, initially crack growth in AA2050 friction stir welds. Int J Fatigue 2008;30:463–72.
[9] Trudel A, Lévesque M, Brochu M. Microstructural effects on the fatigue crack
positioned in the HAZ near the fusion line, deviated towards growth resistance of a stainless steel CA6NM weld. Eng Fract Mech 2014;15:
the base metal. This behavior could be caused by the yield 60–72.
strength mismatch between the filler metal, HAZ and base [10] Xiong Y, Hu XX. The effect of microstructures on fatigue crack growth in Q345
steel welded joint. Fatigue Fract Eng M 2012;35:500–12.
metal. The crack deviation most probably affected the mea- [11] Sanghoon K, Donghwan K, Tae-Won K, Jongkwan L, Changhee L. Fatigue crack
surements of the HAZ crack growth behavior, as microstruc- growth behavior of the simulated HAZ of 800 MPa grade high-performance
ture (and crack growth resistance) of the tested material steel. Mater Sci Eng A–Struct 2011;528:2331–8.
[12] Xuedong W, Qingyu S, Xin W, Zenglei Z. The influences of precrack
evolved with crack length. In this regard, we believe that orientations in welded joint of Ti–6Al–4V on fatigue crack growth. Mater Sci
constant DK fatigue tests with the crack growing perpendic- Eng A–Struct 2010;527:1008–15.
ular to the weld, such as in [9], constitute a more rigorous [13] Kang D-H, Lee J-K, Kim T-W. Corrosion fatigue crack propagation in a heat
affected zone of high-performance steel in an underwater sea environment.
method to compare the fatigue behavior of the filler metal,
Eng Fail Anal 2011;18:557–63.
HAZ and base metal of a weld. However, welding direction [14] Engström H, Westin EM. Effect of post-weld straining on temper-rolled
relative to the specimen geometry should be taken into con- austenitic stainless steel welds. Weld World 2008;52:87–99.
sideration for results analysis as it can influence residual [15] Cheng X, Petrov R, Zhao L, Janssen M. Fatigue crack growth in TRIP steel under
positive R-ratios. Eng Fract Mech 2008;75:739–49.
stresses and fatigue crack growth properties. [16] Pippan R, Flechsig K, Riemelmoser FO. Fatigue crack propagation behavior in
the vicinity of an interface between materials with different yield stresses.
Engineers who are responsible for the runner’s design base their Mater Sci Eng A–Struct 2000;283:225–33.
[17] Zhang H, Zhang Y, Li L, Ma X. Influence of weld mis-matching on fatigue crack
calculations on a hypothetical scenario where the crack is continu- growth behaviors of electron beam welded joints. Mater Sci Eng A–Struct
ously subjected to tensile residual stresses, which often results in a 2002;334:141–6.
fully open crack. Our results show that these assumptions can be [18] Bhat S. Tip parameter approximation and fatigue growth of crack towards
inclined weld interface between strength mismatched steels. Int J Damage
applied to a crack growing in an as-welded HAZ since tensile Mech 2011;20:752–82.
residual stresses are significant. In addition, this approach is [19] Nascimento MP, Voorwald HJC. Considerations about the welding repair
conservative since residual stresses are normally relieved during effects on the structural integrity of an airframe critical to the flight-safety, In:
Procedia engineering. Prague: Czech Republic; 2010. p. 1895–903.
post-weld heat treatment and with subsequent crack growth, [20] Bilmes P, Llorente C, Ipiña J. Toughness and microstructure of 13Cr4NiMo
which allows the crack to close at low load ratios. The results of high-strength steel welds. J Mater Eng Perform 2000;9:609–15.
this study bring additional confidence in currently used fatigue life [21] Thibault D, Bocher P, Thomas M. Residual stress and microstructure in welds
of 13%Cr–4%Ni martensitic stainless steel. J Mater Process Technol 2009;209:
calculation methods by confirming that the HAZ that develops in
2195–202.
turbine runner welds is not a weak link in terms of fatigue crack [22] Thibault D, Bocher P, Thomas M, Gharghouri M, Côté M. Residual stress
growth, as long as the high load ratio crack growth behavior (fully characterization in low transformation temperature 13%Cr–4%Ni stainless
open crack) is considered. steel weld by neutron diffraction and the contour method. Mater Sci Eng A–
Struct 2010;527:6205–10.
[23] Moisan É, Sabourin M, Bernard M, Bui-Quoc T. Residual stress measurements
Acknowledgements in hydraulic turbine welded joints. Presented at the IAHR 23rd symposium on
hydraulic machinery and systems. Japan: Yokohama; 2006.
[24] Thibault D, Bocher P, Thomas M, Lanteigne J, Hovington P, Robichaud P.
This work was supported by Alstom Hydro, Hydro-Québec and Reformed austenite transformation during fatigue crack propagation of
the Natural Sciences and Engineering Research Council of Canada. 13%Cr–4%Ni stainless steel. Mater Sci Eng A-Struct 2011;528:6519–26.
The help of technologists Alexandre Lapointe and Carlo Baillargeon [25] Pukasiewicz AGM, Henke SL, Casas WJP. Effect of post-weld heat treatment on
fatigue crack propagation in welded joints in CA6NM martensite stainless
from Hydro-Québec, as well as of Bénédict Besner from École Poly- steel. Weld Int 2006;20:6.
technique de Montréal is gratefully acknowledged. [26] (2013). 410NiMo AC-DC Datasheet. <http://www.hobartbrothers.com/uploads/
pdf/datasheets/410NiMo_ACDC.pdf.>
[27] ASTM. ASTM E384–11e1 Standard Test Method for Knoop and Vickers
References Hardness of Materials. ed. ASTM International; 2011.
[28] ASTM. ASTM E647–11e1 Standard test method for measurement of fatigue
[1] Sabourin M, Gagné J-L, St-Hilaire GSA. De La Bruère-Terreault J. Mechanical crack growth rates. ed. ASTM International; 2011.
loads and fatigue analysis of a francis runner. Presented at the HydroVision. [29] Trudel A, Brochu M, Lévesque M. Residual stress effects on the propagation of
Canada: Montreal; 2004. fatigue cracks in the weld of a CA6NM stainless steel. Presented at the 13th
[2] Lanteigne J, Sabourin M, Bui-Quoc T, Julien D. The characteristics of the steels international conference on fracture. China: Beijing; 2013.
used in hydraulic turbine runners. Presented at the IAHR 24th symposium on [30] Suresh S, Zamiski G, Ritchie D. Oxide-induced crack closure: an explanation for
hydraulic machinery and systems. Brazil: Foz Do Iguassu; 2008. near-threshold corrosion fatigue crack growth behavior. Metall Mater Trans A
[3] Sabourin M, Thibault D, Bouffard DA, Levesque M. New parameters influencing 1981;12:1435–43.
hydraulic runner lifetime. Presented at the 25th IAHR symposium on hydraulic [31] Gray Iii GT, Williams JC, Thompson AW. Roughness-induced crack closure: an
machinery and systems. Romania: Timisoara; 2010. explanation for microstructurally sensitive fatigue crack growth. Metall Trans
[4] Ohta A, Suzuki N, Maeda Y. Unique fatigue threshold and growth properties of A 1983;14(A):421–33.
welded joints in a tensile residual stress field. Int J Fatigue 1997;19:303–10. [32] McEvily AJ, Ritchie RO. Crack closure and the fatigue-crack propagation
[5] Jang C, Cho P-Y, Kim M, Oh S-J, Yang J-S. Effects of microstructure and residual threshold as a function of load ratio. Fatigue Fract Eng M 1998;21:847–55.
stress on fatigue crack growth of stainless steel narrow gap welds. Mater
Design 2010;31:1862–70.

Please cite this article in press as: Trudel A et al. Fatigue crack growth in the heat affected zone of a hydraulic turbine runner weld. Int J Fatigue (2014),
http://dx.doi.org/10.1016/j.ijfatigue.2014.03.006

You might also like