Вы находитесь на странице: 1из 11

Materials Science and Engineering A 428 (2006) 1–11

Effect of boron on fatigue crack growth behavior in


superalloy IN 718 at RT and 650 ◦C
L. Xiao a , D.L. Chen b , M.C. Chaturvedi a,∗
aDepartment of Mechanical and Manufacturing Engineering, University of Manitoba, Winnipeg, Man., Canada R3T 2N2
b Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street, Toronto, Ont., Canada M5B 2K3
Received 27 April 2005; accepted 30 August 2005

Abstract
The effect of boron on the fatigue crack growth rate (FCGR) in Inconel 718 (IN 718) was studied at room temperature (RT) and 650 ◦ C. The
results showed that the addition of B improved the fatigue crack propagation resistance of IN 718. The higher the B concentration, higher was
the fatigue threshold at 650 ◦ C. While the FCGRs increased as the test temperature increased from RT to 650 ◦ C in the Paris regime, a rapid drop
in the FCGRs in the near-threshold regime and a higher fatigue threshold at 650◦ were observed due to the oxide-induced crack closure. The
fracture surfaces were observed to exhibit transgranular cracking with fatigue striations in specimens tested at RT, and a mixture of transgranular
and intergranular cracking at 650 ◦ C. The fracture mode changed from intergranular cracking to transgranular cracking and plastic deformation
marks increased with increasing B concentration at 650 ◦ C. The micromechanism of improvement in the fatigue crack growth resistance due to B
addition was further studied via observations of crack growth path, fractography and TEM examination of the plastic zone ahead of the crack tip.
The plastic deformation mode within the crack tip plastic zone was planar slip, along with twinning, on {1 1 1} planes. A crystallographic cracking
model was, thus, proposed on the basis of restricted slip or twinning. The improvement in the fatigue crack growth resistance in IN 718 due to B
addition was mainly attributed to the increase in the grain boundary cohesion via minimizing the deteriorative effect of oxygen and the increase in
the resistance to the dislocation movement at the crack tip.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Inconel 718 superalloy; Fatigue crack growth rate; Fatigue threshold; Slip; Twin

1. Introduction addition, they examined the microstructure at the tip of fatigue


cracks using transmission electron microscopy and showed that
Due to its superior mechanical properties and acceptable IN 718 exhibits very inhomogeneous deformation within the
weldability, Inconel 718 (IN 718) is being extensively used plastic zone ahead of the crack tip. Microtwinning played an
as a structural material in the aerospace industry for a vari- important role in maintaining the plastic deformation within the
ety of applications, such as, gas turbine engine disks, blades plastic zone corresponding to the near-threshold FCGR regime
and shafts, where the resistance to fatigue crack growth is an [4]. Yuen and Roy [5] reported that the near-threshold FCGR
important design consideration [1,2]. Therefore, a considerable decreased and fatigue threshold value, Kth , increased with
attention has been paid to the fatigue crack growth behavior of increasing grain size from 22 to 91 ␮m due to an increase in
IN 718 superalloys [1–9]. These studies have demonstrated that the roughness-induced closure [5]. However, Osinkolu et al. [7]
a variation in test temperature, composition and microstructure suggested that the FCGR in IN 718 is insensitive to the grain
can produce an order of magnitude of differences in the fatigue size in air at 650 ◦ C. James and Mills [1,8,9] studied the FCGR
crack growth rates (FCGRs) in this alloy. Clavel and Pineau [3,4] in weldments of IN 718 and reported that FCGRs in welded
reported that an increase in test temperature from 25 to 650 ◦ C or specimens were higher than those observed in wrought speci-
a decrease in the test frequency at 550 ◦ C produced a significant mens due to the presence of residual stresses in the as-welded
increase in the FCGR, especially in the near-threshold regime. In condition.
Although the FCGR in IN 718 has been extensively stud-
ied, relatively little is known about the effect of boron on the
∗ Corresponding author. Tel.: +1 204 474 6675; fax: +1 204 261 6735. fatigue crack growth behavior. The presence of boron in super-
E-mail address: mchat@cc.umanitoba.ca (M.C. Chaturvedi). alloys and weld metal can have a significant effect on their

0921-5093/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2005.08.206
2 L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11

mechanical properties, depending upon whether this element is at a rate of 50 ◦ C/h, and held at 621 ◦ C for a total aging time of
present as a free atomic species or incorporated within particular 8 h before air cooling to RT.
microstructural features, such as grain boundaries, inclusions or
precipitates [10–13]. The nature of distribution of B in the alloy, 2.2. Fatigue crack growth testing and microstructural
to certain extent, will depend upon its concentration. Therefore, observations
one of the purposes of this investigation was to study the effect
of boron on the fatigue crack propagation behavior of IN 718 The FCGR tests were performed with a 30 kN Instron servo-
superalloy at RT and 650 ◦ C. hydraulic fatigue testing system, which was interfaced with a
In addition, it has been known for sometime that small addi- computer for the test control and data acquisition. The specimens
tions of boron and zirconium are extremely beneficial to the were heated to 650 ◦ C using a resistance furnace having an accu-
creep rupture properties of nickel-base superalloys [10–12]. The racy of ±3 ◦ C. A sinusoidal loading waveform at a frequency
commonly held view is that these elements help prevent prema- of 30 Hz and a load ratio (R = Pmin /Pmax ) of 0.05 was applied.
ture cavity and microcrack formation at the grain boundaries, All tests were performed in air. Crack lengths were measured
and thus improve the rupture lifetime and ductility. The opti- using a direct current potential drop (DCPD) technique. Load
mum concentrations of B and Zr depend somewhat on the base shedding technique using an automated K-decreasing proce-
composition of the alloys, but in most alloys, that contain B dure in accordance with ASTM E647-99 was used to determine
and Zr, it is typically of the order of 60 ppm B and 500 ppm Zr the near-threshold fatigue crack growth rates. The K-gradient,
[10–12]. The reason why B minimizes grain boundary cracking C = (1/K)(dK/da), was selected to be −0.1.
has, however, remained a subject of debate. Fatigue crack propagation paths were examined with an opti-
The mechanisms of fracture and particularly the disloca- cal microscope and fracture surfaces were observed using a
tion configurations that arise during low and high cycle fatigue JSM-5900LV scanning electron microscope. Some of the fatigue
of nickel-base superalloys have been also extensively studied crack growth tests were interrupted for a detailed microstructural
[3,13–18]. However, very limited work has been done concern- examination of the crack tip region using transmission electron
ing the micromechanisms of deformation at the crack tip and microscope (TEM). The material in front of the crack tip was
of fracture during fatigue crack propagation [3,19,20]. A fur- removed by a numerically controlled wire EDM, and the TEM
ther purpose of this investigation, therefore, was to identify the foils were then prepared using a combination of precision dim-
micromechanism of fatigue crack propagation in conjunction pling, mechanical grinding and electro-polishing techniques to
with the effect of B on FCGR in IN 718 at 650 ◦ C. ensure that the electron transparent region in the foil was within
the plastic zone directly ahead of the crack tip. The foils were
2. Experimental examined using a JEOL-2000FX electron microscope operating
at 200 kV.
2.1. Materials and heat treatment
3. Results
Four groups of IN 718 superalloys with a similar base chem-
ical composition of (wt%) 18.45% Cr, 53.43% Ni, 18.83% Fe, 3.1. Microstructure of the heat-treated IN 718
2.91% Mo, 4.86% Nb, but containing 12, 29, 60 and 100 ppm
B, respectively, were produced via a vacuum induction melt- A typical SEM micrograph of the heat-treated IN 718 is
ing (VIM) + vacuum arc remelting (VAR) for this research. A shown in Fig. 1. It is seen that grain boundaries of the solu-
split 140 kg heat was made. The first half of heat was poured tion treated material were extensively decorated with particles
into a 108 mm electrode with a B level of 60 ppm. Additional of various morphologies, ranging from globular to almost con-
B was added into the remainder of the heat to increase the B tinuous films. EDS X-ray microanalysis revealed that these
level to 100 ppm and it was also poured into a 108 mm elec-
trode. Both electrodes were VAR’d into 140 mm ingots. Ingots
were homogenized at 1163 ◦ C for 16 h and then press-forged
to 76 mm × 76 mm billet at 1093 ◦ C. The billets were further
rolled to 12.7 mm × 79 mm plates at 1066 ◦ C. A similar pro-
cessing procedure was used to produce alloys with 12 and
29 ppm B. Standard compact tension (CT) specimens with a
thickness of 12.7 mm were machined from the plate in the T–L
orientation (T, transverse direction or the direction of least defor-
mation and L, longitudinal direction or the direction of prin-
cipal deformation) using electro-discharge machining (EDM)
techniques.
The specimens were heat-treated by using a commercial heat
treatment (CHT) procedure, consisting of a solution treatment at
954 ◦ C for 1 h and then air cooled to RT. The specimens were pre-
cipitation hardened by aging at 718 ◦ C for 8 h, cooled to 621 ◦ C Fig. 1. SEM image of as heat-treated IN 718.
L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11 3

Fig. 2. TEM dark field image of heat-treated IN 718 in [1 0 0] orientation taken


with [1 1 0] reflection, showing the presence of ␥ and ␥ precipitates.

particles were ␦ (Ni3 Nb) phase. Some boron containing precipi-


tate particles were observed at grain boundaries, as indicated
by an arrow in Fig. 1, and an evidence of segregation of B
on grain boundaries of IN 718 superalloy was observed by Fig. 3. Influence of boron concentration on the fatigue crack growth behavior
of IN 718 at RT.
one of the authors by SIMS analysis in an earlier studies on
the weldability of IN 718 [21,22]. TEM examination showed
that the aged IN 718 is precipitation strengthened primarily tested at RT, as indicated by a continuous decrease in K values
by ␥ precipitates, which has a Ni3 Nb based DO22 ordered with decreasing FCGR (Fig. 3).
body-centered tetragonal structure. A typical dark field TEM Fig. 4 shows the effect of B concentration on the FCGRs of the
micrograph taken with (1 1 0) ␥ reflection is presented in Fig. 2. specimens fatigued at 650 ◦ C. Again, the alloy with the lowest
It shows the presence of disc-shaped ␥ particles with a size concentration of B (12 ppm) had the worst fatigue crack propa-
of about 30 nm in diameter and 5 nm in thickness, which have gation resistance. The FCGRs decreased as the B concentration
the following cube–cube relationship with the parent fcc ␥ increased from 12 to 29 ppm. The alloy with 60 ppm B had a
matrix: [0 0 1]␥ ||0 0 1␥ and {1 0 0}␥ ||{1 0 0}␥ , as has been lower FCGR than the alloy with 29 ppm B at the near-threshold
also observed by others [23–27]. In addition, some spherical- regime, whereas a reversed response was observed at the higher
shaped ␥ precipitates of approximately 5 nm in diameter were
also observed in the dark field images obtained with (1 1 0)
reflection.

3.2. Effect of boron on the fatigue crack growth rate at RT


and 650 ◦ C

The FCGRs, da/dN, obtained at RT, as a function of the stress


intensity factor range, K, are shown in Fig. 3 for the four IN
718 alloys with different concentrations of B. It is seen that the
addition of boron affects the FCGR of IN 718, particularly at
the lower K levels. The alloy with the lowest B concentra-
tion (12 ppm) exhibited the highest near-threshold crack growth
rates, and it decreased as the B concentration increased from 12
to 29 ppm. While a higher crack growth rate was observed in the
alloy with 60 ppm B than in the alloy with 29 ppm B at the higher
K value and near-threshold regions. The difference, however,
was reduced to near zero or values decreased at intermediate
K values. The alloy with 100 ppm B had a significantly lower
FCGR compared to the other three alloys with lower concentra-
tions of B. These results suggest that the addition of B within the
concentration levels considered in this study generally improved
the fatigue crack growth resistance of IN 718 at RT. No obvious Fig. 4. Influence of boron concentration on the fatigue crack growth behavior
fatigue crack growth threshold was observed in the four alloys of IN 718 at 650 ◦ C.
4 L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11

Table 1
Paris parameters of IN 718 with different concentrations of B at RT and 650 ◦ C
Alloys B concentration Paris law
(ppm)
RT 650 ◦ C

12 12 da/dN = 5 × 10−16 da/dN = 1 × 10−11


K6.18 K3.27
29 29 da/dN = 2 × 10−16 da/dN = 9 × 10−12
K6.27 K3.17
60 60 da/dN = 1 × 10−15 da/dN = 1 × 10−12
K5.62 K3.99
100 100 da/dN = 2 × 10−19 da/dN = 1 × 10−13
K8.20 K4.48

K values, thus causing a cross-over of FCGR curves between


the alloys with 29 and 60 ppm B. The best fatigue crack growth
resistance was again observed in the alloy with 100 ppm B. It
is evident that the B addition significantly enhanced the fatigue
crack growth resistance of IN 718 tested at 650 ◦ C as well. In
addition, a rapid drop in the FCGRs in the near-threshold regime
was observed in all the four alloys tested, as shown in Fig. 4,
indicating the presence of a fatigue crack growth threshold.
The relationship between the da/dN and K in the interme-
diate crack propagation regime for the specimen fatigued at RT
and 650 ◦ C may be expressed by the Paris law, as illustrated
in Table 1. A list of fatigue crack propagation threshold values
obtained with different B concentrations and test temperatures
is given in Table 2. Since the FCGR decreased continuously
with decreasing K value at RT (Fig. 3), the fatigue crack prop-
agation threshold was defined as the K value corresponding
to a growth rate of ∼10−10 m/cycle. The fatigue threshold at
650 ◦ C was well defined as the K value corresponding to the
sharp decrease in the FCGR (Fig. 4). As seen from Table 2, the
higher the B concentration the higher was the fatigue threshold
at 650 ◦ C. The fatigue threshold was observed to be higher at
650 ◦ C than at RT.

3.3. Effect of test temperature on the fatigue crack growth


rate

To compare the fatigue crack growth behavior of the four


alloys at different test temperatures, a direct comparison of their
FCGRs is presented in Fig. 5(a and b). It is seen that for each of Fig. 5. Influence of test temperature on the fatigue crack growth behavior of IN
the four alloys there is a considerable difference in the FCGRs 718 alloys with different B concentrations: (a) 12 B and 60 ppm B and (b) 29 B
between RT and 650 ◦ C. The FCGRs increased noticeably as and 100 ppm B.
the test temperature increased. However, in the near-threshold
regime the FCGRs at 650 ◦ C reduced drastically, giving rise to
the fatigue threshold values higher than those observed at RT.
Table 2 Such a trend was observed in all the alloys with different B
Fatigue threshold values of IN718 alloys with different concentrations of B at concentrations. Further examination showed that the test tem-
RT and 650 ◦ C perature mainly influenced the FCGR in the lower crack growth
Alloys B concentration (ppm) Fatigue threshold regime. The extent of improvement in the fatigue crack growth
resistance due to B addition seems to be larger at 650 ◦ C than that
RT 650 ◦ C
observed√at RT. The fatigue threshold value increased
√ from about
12 12 7.6 8.5 7.6 MPa m at RT to approximately 8.5 MPa √m at 650 ◦ C for
29 29 8.6 9.4 the alloy with 12 ppm B, from 8.6 to 9.4 MPa m for the alloy
60 60 7.8 10.6 √
100 100 – 11.6
with 29 ppm B, and from 7.8 to 10.6 MPa m for the alloy with
60 ppm B. This could probably be due to the influence of B in
L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11 5

Fig. 6. Fatigue crack growth paths in IN 718 with different B concentrations fatigued at RT: (a) 12 ppm B, (b) 29 ppm B, (c) 60 ppm B and (d) 100 ppm B.

minimizing the deteriorative effect of oxygen on the FCGR to or the preferred crystallographic cracking planes was found to
some extent, which will be discussed later. be about 120◦ . It was also observed that slip was heavily con-
centrated on the {1 1 1} slip planes in IN 718 during low cycle
3.4. Metallography and fractography fatigue and fatigue crack growth at RT and 650 ◦ C [3,14–17].
Fig. 8 shows the fractographs of specimens with different
The fatigue crack growth paths in alloys with different con- concentrations of B fatigued at RT. A typical feature is seen
centrations of B fatigued at RT are shown in Fig. 6. The alloys to be the presence of transgranular fracture with a great deal
with 12 and 60 ppm B displayed a relatively smooth and straight of fatigue striations. Some secondary cracks along the grain
growth path (Fig. 6(a and c)), whereas the alloys with 29 and boundaries can be observed in Fig. 8(a–c). This mode of frac-
100 ppm B exhibited a tortuous and rough path. Plastic defor- ture suggests that the crack grew via the usual plastic blunting
mation marks could be observed in the region around the fatigue mechanism. In the alloy with low B concentration, the fracture
cracks, as indicated by arrows in Fig. 6(b and d). The improved mode was transgranular with an appearance of crystallographic
fatigue crack growth resistance in the alloys with 29 and 100 ppm fracture (Fig. 8(a)). There were no significant differences in
B in comparison with the alloys with 12 and 60 ppm B can be the fracture modes among the alloys with 12, 29 and 60 ppm
partly attributed to the enhancement of the fatigue crack closure B, as shown in Fig. 8(a–c). However, as the B concentration
caused by the fracture surface roughness in the near-threshold increased from 60 and 100 ppm, the nature of fracture changed
regime. significantly, and was characterized by the presence of a high
Regions of multiple crystallographic cracking in the alloy density of plastic tear ridges and fatigue striations, as shown in
with 100 ppm B were observed by optical microscope on the Fig. 8(d).
surface of specimen in the vicinity of the fatigue crack, as shown Fractographic observations of the specimens tested at 650 ◦ C
in Fig. 7. Further examination showed that they were induced by showed that the cracks propagated basically in an intergranular
an extensive plastic deformation via planar slip and/or by twin- mode when B concentration was 12 ppm (Fig. 9(a)), and by a
ning. Some secondary cracks were observed within the plastic mixture of intergranular and transgranular mode in the alloy with
deformation bands, as indicated by arrows in Fig. 7. The angle 29 ppm B, as shown in Fig. 9(b). In the alloy with 60 ppm B, the
between the two groups of parallel plastic deformation traces fracture surface had a mixed transgranular/intergranular nature

Fig. 7. An optical micrograph showing the crystallographic cracking during fatigue crack growth in IN 718 with 100 ppm B fatigued at RT.
6 L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11

Fig. 8. Typical fracture surfaces in IN 718 fatigued at RT: (a) 12 ppm B, (b) 29 ppm B, (c) 60 ppm B and (d) 100 ppm B.

with a large number of fatigue striations (Fig. 9(c)). On a closer were observed on the {1 1 1} primary slip planes (Fig. 10(a)).
examination, some secondary cracks along the grain boundaries This indicates that slip was activated on the {1 1 1} slip planes
were also observed in the alloys with 29 and 60 ppm B, as indi- within the plastic zone ahead of the crack tip. In addition, defor-
cated by arrows in Fig. 9(b and c). The fracture mode exhibited mation twins were also observed to be activated in the crack tip
a transition from intergranular to transgranular fracture as the B regions, as shown in Fig. 10(b). Further examination revealed
concentration increased from 60 to 100 ppm (Fig. 9(d)), where that some plastic deformation traces were present within the
secondary cracks along the grain boundaries were not observed. twins. Twinning was observed to be more abundant at 650 ◦ C
The characteristic fatigue striations observed in the alloy with than at RT. The interaction between the twins was also observed,
higher B concentrations suggest that the fatigue crack propaga- as indicated by an arrow in Fig. 11(a), and Fig. 11(b) shows that
tion at 650 ◦ C also occurred via the plastic blunting mechanism twin bands were produced in the alloy with 100 ppm B fatigued
at 650 ◦ C (Fig. 9(c and d)). The change in the fracture mode, at 650 ◦ C. These results imply that twinning also plays a role in
from the intergranular to transgranular cracking as the B con- fatigue deformation of IN 718.
centration increased, confirms that the major role of boron added
to IN 718 is to increase the grain boundary strength of the
alloy. 4. Discussion

3.5. Microstructural examination of the fatigue crack tip 4.1. Fatigue crack growth mechanism

In order to gain insight into the fatigue crack growth mech- TEM observations showed that the ␥ precipitates were
anism operating in these alloys, some fatigue tests were inter- sheared by dislocations during deformation of IN 718 superal-
rupted at K values corresponding to the intermediate crack loy [17,18,23,24]. As the accumulated plastic strain increased,
growth regime, and then TEM foils from the plastic zone at the dislocations were confined to discrete planar bands, which
the tip of the fatigue crack were prepared for TEM analysis. led to the formation of intense deformation bands. The shearing
As shown earlier, at the end of the double aging treatment, mechanism of ␥ particles has been discussed by several authors
the microstructure consisted of homogeneously distributed ␥ [15,17,25–27]. Sundararaman et al. [25] reported that when the
phase. The TEM micrographs in the crack tip regions in the size of ␥ particles in IN 718 superalloy exceeded a critical
Paris regime are presented in Fig. 10(a and b). Some dislocations value (∼10 nm), they were sheared by deformation twins, and
L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11 7

Fig. 9. Typical fracture surfaces in IN 718 fatigued at 650 ◦ C: (a) 12 ppm B, (b) 29 ppm B, (c) 60 ppm B and (d) 100 ppm B.

the smaller precipitates were sheared by the movement of paired along the crack growth path in IN 718, as indicated by arrows in
dislocations. Clavel and Pineau [3], however, showed that the Fig. 7.
precipitates in IN 718 were sheared by the propagation of twins. A slip-dependent and/or twin-dependent fatigue cracking
Oblak et al. [26,27] observed that the shearing of ␥ particles model based on the restricted availability of planes is proposed,
appeared to take place by the coupled motion of a/21 1 0 pairs as described schematically in Fig. 12. Crystallographic cracking,
in IN 718 deformed to 3% strain at RT. Worthem et al. [15], on the which has been previously reported in nickel-base superalloys
other hand, reported that the shearing of ␥ precipitates occurred [28] and aluminium–lithium alloys [29], is a crack propagation
by slip only during cyclic deformation and no significant shear- mode along the preferred crystallographic planes. The formation
ing of ␥ precipitates was observed after the monotonic strains, of planar slip bands or twinning bands is a necessary condition
and no deformation twins were identified. Xiao et al. [23] fur- for the crystallographic cracking [28,30]. The precipitates were
ther confirmed the occurrence of shearing of ␥ particles by sheared in the course of cyclic straining, resulting in the for-
slip in IN 718 during cyclic deformation. Therefore, based upon mation of planar bands or precipitate-free channels. The cyclic
the observations in this investigation and the findings reported plastic deformation in front of the crack tip proceeded via the
by other authors [3,15,23–27], it seems likely that the plastic formation and propagation of planar bands [23–27]. Thus, the
deformation occurred at the crack tip during fatigue crack prop- fatigue cracks grew primarily along the planar slip or twinning
agation was by slip and twinning on {1 1 1} planes. Deformation bands on {1 1 1} planes intersecting with the crack tip, as shown
was inhomogeneous and channelization of deformation seems schematically in Fig. 12. Twinning planes in fcc alloys are invari-
to have occurred in planar deformation bands after ␥ precip- ably the close-packed {1 1 1} planes, which are also the slip
itates were sheared. The trailing dislocation in the same slip planes, while the twinning directions are 1 1 2. An increase in
plane would repeatedly shear the ␥ precipitates during cyclic crack length is produced by shear de-cohesion at the crack tip,
deformation and continuously reduce their size to such an extent and the extent of such de-cohesion is dependent on the extent
that they offered very little or no resistance to the movement of of intense plastic deformation ahead of the crack tip and the
dislocations. This phenomenon would cause cyclic softening applied stress–strain field. Fig. 12(b) is a three-dimensional rep-
and would apparently lead to the formation of precipitate-free resentation of the four {1 1 1} slip or twinning planes in a fcc
deformation bands [18]. The development of soft precipitate- crystal, and Fig. 12(c) shows projection of three {1 1 1} planes,
free regions along slip planes might lead to early crack initiation. i.e., (1 1 1), (1 1 1) and (1 1 1) planes, when the foil normal is
Indeed, a large numbers of secondary cracks were generated parallel to [1 1 1]. The angle between the traces of the above
8 L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11

Fig. 11. Transmission electron micrographs showing twin substructures formed


Fig. 10. Transmission electron micrographs showing deformation substructures in IN 718 in the crack tip plastic zone corresponding to the Paris regime: (a)
of IN 718 in the crack tip region corresponding to the Paris regime: (a) slip an interaction between two twins in the alloy with 12 ppm B at RT and (b)
dislocation lines formed on the (1 1 1) slip plane in the alloy with 60 ppm B at deformation microtwin bands formed in the alloy with 100 ppm B at 650 ◦ C.
650 ◦ C and (b) a twin formed in the alloy with 60 ppm B at 650 ◦ C.

Fig. 12. A schematic diagram illustrating the fatigue crack growth by shearing planes of slip or twinning: (a) crack growth, (b) three-dimensional structure of {1 1 1}
slip or twinning planes in the [112̄]–[1̄10]–[1 1 1] space and (c) projections of {1 1 1} planes on (1 1 1) plane.
L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11 9

three {1 1 1} planes with the (1 1 1) plane is either 60 or 120◦ , stress above the critical stress for the preferential slip would
and is in good agreement with the experimental observations be required for dislocations to escape the pinning by B atoms
illustrated in Fig. 7. within the cross-slip plane. As a result, B addition can influence
the slip distribution and the relevant mode of cracking in these
4.2. Effect of boron on plastic deformation at the tip of alloys. In the alloy with higher B concentration these defor-
fatigue crack mation bands disperse before a crack could be initiated. It has
been shown that, at a lower B concentration, however, plas-
As shown in the previous section, the addition of B leads to tic deformation was concentrated in the local regions due to
a lower value of FCGR and a higher fatigue threshold value in cross-slip being easily activated, and it led to the crack initiation
IN 718 at RT and 650 ◦ C. The mechanism responsible for the within the bands [16]. Microscopically, strong slip concentra-
improvement in the fatigue crack growth resistance of IN 718 tion within the persistent slip bands would lead to fatigue crack
due to the addition of B is discussed next. nucleation and favor crack propagation, thereby would decrease
Slip reversibility, crack branching and resistance to active the fatigue crack growth resistance of IN 718 with lower B con-
environments can be critical factors influencing the fatigue crack centration. The planar slip characteristics in the higher B alloy
propagation resistance of metals and alloys. It seems likely that has been attributed to crack branching and mixed mode I and
the mechanism of fatigue crack propagation of IN 718 in the II crack propagation [31]. The zig-zag crack paths thus pro-
Paris regime is associated with the crystallographic fracture in duced, as shown in Fig. 7, would promote roughness-induced
the favored crystallographic directions, induced by the separa- crack closure which in turn would reduce the effective stress
tion of atomic bonds directly ahead of the crack tip (Fig. 7). intensity factor range Keff —the driving force for fatigue crack
Accommodation of the plastic strain in front of the crack tip propagation. By contrast, the alloy with lower B concentration
appears to occur via fracturing or shearing of ␥ -Ni3 Nb precipi- showed relatively straight fracture paths, resulting in smaller
tates, which may occur as a result of high local stresses, or some crack closure. A combination of changes in the slip character in
limited amount of dislocation motion or micro-twinning activity the plastic zone ahead of the crack tip and the closure contribu-
(Fig. 11). These plastic deformation features were observed to tion behind the crack tip could account for the increased fatigue
be localized, and very fine (Figs. 10 and 11). crack growth resistance of the material with increasing B con-
One mechanism for the observed improvement in the fatigue centration. The flatter surface in the alloy with 60 ppm B shown
crack propagation resistance due to B addition could be related in Fig. 6(c) corresponds to a lower fatigue threshold (Table 2)
to the increased grain boundary cohesion arising from B seg- and/or faster near-threshold crack growth rate (Fig. 3) due to the
regation, which has been demonstrated to occur in IN 718 by lower crack closure level. However, the reason why the crack
SIMS analysis [21,22]. This would lead to a reduced tendency growth path in the alloy with 60 ppm B, shown in Fig. 6(c),
of cracking along grain boundaries in the higher B alloy. Stress is a little flatter than that observed in the alloy with 29 ppm B
and strain concentrations at the grain boundary regions, induced shown in Fig. 6(b), is not understood and further studies are
by a lack of sufficient number of slip systems, can be released by required.
either intergranular cracking or transgranular cracking, depend-
ing on the cohesion strength of the boundary relative to that 4.3. Effect of boron on fracture mode of IN 718
of the bulk material. Metallographic and SEM examinations
of the specimens revealed a much higher incidence of grain The difference in the fatigue crack growth rates of IN 718
boundary cracking in the lower B alloys at RT and 650 ◦ C, between RT and 650 ◦ C may be explained in terms of slip
as shown in Figs. 8 and 9. This indicates that the boundaries reversibility and oxide penetration when deformation occurs
with a higher B concentration are more resistant to intergran- by the shearing of precipitates. Temperature influences fatigue
ular cracking. In general, intergranular cracking occurs at a crack growth behavior via its effect on both the localized strain at
faster rate than transgranular cracking. Therefore, the absence the crack tip and the kinetics of environmental interactions. It has
of intergranular cracking in alloys with higher B concentration been long recognized that air, or more specifically oxygen, can
would partly account for their superior fatigue crack growth significantly degrade the fatigue properties of superalloys [6],
properties. thus leading to a significant increase in the fatigue crack prop-
Another possible mechanism for the improvement could be agation rate at 650 ◦ C, as shown in Fig. 5. The most common
via direct elastic interaction between dislocations and B atoms explanation for this effect is that oxygen diffuses into the grain
in solution, and by boride particles. As an interstitial element, boundaries and makes them more susceptible to intergranular
B atoms may form clusters which, along with boride particles, cracking [6]. The change in the fracture mode from transgran-
would act as obstacles to the dislocation movement by imped- ular cracking at RT to intergranular cracking at 650 ◦ C in the
ing the dislocation slip, and thus enhance the FCG resistance. alloy with lower B concentration, as shown in Figs. 8(a) and 9(a),
It was observed that planar deformation bands were generated corroborates that the role of oxygen is to reduce the grain bound-
in the fatigued IN 718 with the higher B concentration [16,17]. ary energy in such a way that the crack growth path becomes
This was suggested to be due to the formation of preferential intergranular. On the other hand, it is the environmental inter-
paths for the dislocation movement, resulting in a lower crit- actions at 650 ◦ C that give rise to a rapid reduction or a halt in
ical stress for the slip of subsequent dislocations. This would the fatigue crack growth when it is smaller than approximately
make cross-slip relatively more difficult, since an additional 10−8 m/cycle, leading to a higher fatigue threshold as compared
10 L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11

to the RT tests (Fig. 5). In this situation, an oxide layer would transgranular cracking, indicating that the beneficial role of
be expected to form on the freshly cracked surface in the wake B was to an increase in the grain boundary cohesion.
of the crack tip, causing the oxide-induced crack closure, which 5. TEM examination of the material in the crack tip regions
is especially significant in the near-threshold regime [32]. showed that the mode of plastic deformation was mainly slip,
The fractographic examination indicated that an increase in B but twinning also played a role.
concentration reduced the tendency to intergranular separation 6. During fatigue of IN 718 plastic deformation occurred via
in IN 718 fatigued at 650 ◦ C, as shown in Fig. 9. The intergran- planar slip or twinning in the vicinity of the crack in the
ular cracking was basically eliminated in alloys with higher B Paris regime. A crystallographic cracking mechanism was
concentrations, and entirely transgranular fracture was observed proposed on the basis of planar slip or twinning plane.
in the alloy with 100 ppm B (Fig. 9(d)). This implies that B 7. Fatigue crack growth occurred primarily along the planar slip
drastically suppressed the occurrence of intergranular cracking, bands or twinning bands on {1 1 1} planes intersecting with
which is most likely controlled by the grain boundary cohesion. the crack tip.
The major effect of B addition appeared, therefore, to increase 8. Addition of B increased the resistance to dislocation move-
the grain boundary cohesion and to minimize the deleterious ment due to the interaction between B atoms and dislocations,
effect of oxygen in the air environment. Furthermore, the segre- thus giving rise to an increased FCG resistance. At 650 ◦ C
gation of B at grain boundaries may tie up vacancies and reduce boron suppressed the occurrence of intergranular fracture
grain-boundary diffusion reactions, and thus prevent a premature which was basically controlled by the grain boundary cohe-
formation of cavities and microcracks at the grain boundaries. sion. The major effect of B addition appeared, therefore, to
That is, the beneficial role of B is to counteract the accelerating increase the grain boundary cohesion and to minimize the
effect of oxygen on the nucleation and growth of cavities, or deleterious effect of oxygen in the air environment at ele-
act as a getter of detrimental impurity elements, thus mitigate vated temperatures.
their negative effects [11]. Accordingly, at 650 ◦ C the fracture
mode changed from intergranular cracking in the alloys with Acknowledgement
lower B concentration to transgranular cracking in the alloys
with higher B concentration, giving rise to an increased fatigue The authors would like to thank the Natural Sciences and
crack growth resistance with increasing B concentration, as seen Engineering Research Council of Canada for the financial
in Fig. 4. support.

5. Summary and conclusions References

1. Addition of boron improved the fatigue crack growth resis- [1] L.A. James, ASME J. Eng. Mater. Technol. 103 (1981) 234.
tance of IN 718 alloy at room temperature (RT) and at 650 ◦ C, [2] T. Connolley, P.A.S. Reed, M.J. Starink, Mater. Sci. Eng. A 340 (2003)
particularly in the near-threshold regime. The alloy with 139.
[3] M. Clavel, A. Pineau, Metall. Trans. 9A (1978) 471.
lower B concentration exhibited faster fatigue crack growth
[4] M. Clavel, A. Pineau, Mater. Sci. Eng. 55 (1982) 157.
rates (FCGRs) than the alloy with higher B concentration. [5] J.L. Yuen, P. Roy, Scripta Metall. 19 (1985) 17.
The higher the B concentration was, the higher was the [6] J.L. Yuen, P. Roy, W.D. Nix, Metall. Trans. 15A (1984) 1769.
fatigue threshold at 650 ◦ C. The intermediate stable crack [7] G.A. Osinkolu, G. Onofrio, M. Marchionni, Mater. Sci. Eng. A 356
propagation behavior could be expressed via the Paris equa- (2003) 425.
[8] L.A. James, W.J. Mills, ASME J. Eng. Mater. Technol. 107 (1985) 34.
tion.
[9] W.J. Mills, L.A. James, ASME J. Eng. Mater. Technol. 107 (1985) 41.
2. The FCGRs of IN 718 increased noticeably as the test temper- [10] S. Floreen, J.M. Davidson, Metall. Trans. 14A (1983) 895.
ature increased from RT to 650 ◦ C due to the slip reversibility [11] T.J. Garosshen, T.D. Tillman, G.P. McCarthy, Metall. Trans. 18A (1987)
and environmental interaction. The extent of improvement in 69.
fatigue crack propagation resistance due to B addition was [12] W.D. Cao, R.L. Kennedy, in: E.A. Loria (Ed.), Superalloys 718, 625,
found to be larger at 650 ◦ C than at RT. 706 and Various Derivatives, The Minerals, Metals & Materials Society,
1997, p. 511.
3. In spite of the presence of a well-defined fatigue crack growth [13] H.F. Merrick, Metall. Trans. 5 (1974) 891.
threshold at 650 ◦ C, no obvious fatigue threshold region was [14] T.H. Sanders Jr., R.E. Frishmuth, G.T. Embley, Metall. Trans. 12A
observed for the alloys fatigued at RT. The rapid drop in the (1981) 1003.
FCGRs in the near-threshold regime and the higher fatigue [15] D.W. Worthem, I.M. Robertson, F.A. Leckie, D.F. Socie, C.J. Altstetter,
threshold value at 650◦ , compared to those obtained at RT, Metall. Trans. 21A (1990) 3215.
[16] L. Xiao, D.L. Chen, M.C. Chaturvedi, Metall. Trans. 35A (2004) 3477.
was due to the oxide-induced crack closure. [17] L. Xiao, D.L. Chen, M.C. Chaturvedi, Metall. Trans. 36A, in press.
4. The typical fracture surface obtained at RT was characterized [18] S. Kalluri, K.B.S. Rao, G.R. Halford, M.A. McGaw, in: E.A. Loria
by the transgranular fracture together with fatigue striations (Ed.), Superalloys 718, 625, 706 and Various Derivatives, The Minerals,
for the alloys with different concentrations of B. However, the Metals & Materials Society, 1994, p. 593.
[19] C. Mercer, W.O. Soboyejo, in: E.A. Loria (Ed.), Superalloys 718, 625,
fracture mode changed from the predominantly transgranular
706 and Various Derivatives, The Minerals, Metals & Materials Society,
cracking at RT to intergranular cracking at 650 ◦ C in the alloy 1997, p. 577.
with lower B concentration. With increasing B concentration, [20] C. Mercer, A.B.O. Soboyejo, W.O. Soboyejo, Mater. Sci. Eng. A 270
the fracture mode changed from intergranular cracking to (1999) 308.
L. Xiao et al. / Materials Science and Engineering A 428 (2006) 1–11 11

[21] X. Huang, M.C. Chaturvedi, N.L. Richards, J. Jackman, Acta Mater. 45 [28] R.B. Scarlin, Metall. Trans. 7A (1976) 1535.
(1997) 3095. [29] D.L. Chen, M.C. Chaturvedi, Metall. Mater. Trans. 31A (2000)
[22] W. Chen, M.C. Chaturvedi, Acta Mater. 45 (1997) 2735. 1531.
[23] L. Xiao, D.L. Chen, M.C. Chaturvedi, Scripta Mater. 52 (2004) 603. [30] S.P. Lynch, Fatigue Mechanisms A Symp. Sponsored by ASTM Com-
[24] D. Fournier, A. Pineau, Metall. Trans. 8A (1977) 1095. mittee E-9 on Fatigue National Bureau of Standards National Science
[25] M. Sundararaman, P. Mukhopadhyay, S. Banerjee, Acta Metall. 36 Foundation Kansas City, Mo., 22–24, 1978, p. 174.
(1988) 847. [31] C.J. Beevers, R.L. Carlson, in: R.A. Smith (Ed.), Fatigue Crack Growth
[26] D.F. Paulonis, J.M. Oblak, D.S. Duvall, Trans. ASM 62 (1969) 611. 30 Years of Progress, Pergamon Press, 1984, p. 89.
[27] J.M. Oblak, D.F. Paulonis, D.S. Duvall, Metall. Trans. 5A (1974) 143. [32] D.L. Chen, B. Weiss, R. Stickler, Eng. Fract. Mech. 53 (1996) 493.

Вам также может понравиться