You are on page 1of 14

Low-Cycle Fatigue Behavior of INCONEL 718 Superalloy

with Different Concentrations of Boron at Room Temperature


L. XIAO, D.L. CHEN, and M.C. CHATURVEDI

Symmetrical push-pull low-cycle fatigue (LCF) tests were performed on INCONEL 718 superalloy
containing 12, 29, 60, and 100 ppm boron (B) at room temperature (RT). The results showed that
all four of these alloys experienced a relatively short period of initial cyclic hardening, followed
by a regime of softening to fracture at higher cyclic strain amplitudes (t /2  0.8 pct). As the
cyclic strain amplitude decreased to t /2  0.6 pct, a continuous cyclic softening occurred with-
out the initial cyclic hardening, and a nearly stable cyclic stress amplitude was observed at t /2
 0.4 pct. At the same total cyclic strain amplitude, the cyclic saturation stress amplitude among
the four alloys was highest in the alloy with 60 ppm B and lowest in the alloy with 29 ppm B.
The fatigue lifetime of the alloy at RT was found to be enhanced by an increase in B concentra-
tion from 12 to 29 ppm. However, the improvement in fatigue lifetime was moderate when the B
concentration exceeded 29 ppm B. A linear relationship between the fatigue life and cyclic total
strain amplitude was observed, while a “two-slope” relationship between the fatigue life and
cyclic plastic strain amplitude was observed with an inflection point at about p /2  0.40 pct.
The fractographic analyses suggested that fatigue cracks initiated from specimen surfaces, and trans-
granular fracture, with well-developed fatigue striations, was the predominant fracture mode. The
number of secondary cracks was higher in the alloys with 12 and 100 ppm B than in the alloys
with 29 and 60 ppm B. Transmission electron microscopy (TEM) examination revealed that typi-
cal deformation microstructures consisted of a regularly spaced array of planar deformation bands
on {111} slip planes in all four alloys. Plastic deformation was observed to be concentrated in
localized regions in the fatigued alloy with 12 ppm B. In all of the alloys,   precipitate particles
were observed to be sheared, and continued cyclic deformation reduced their size. The observed
cyclic deformation softening was associated with the reduction in the size of   precipitate parti-
cles. The effect of B concentration on the cyclic deformation mechanism and fatigue lifetime of
IN 718 was discussed.

I. INTRODUCTION Cao and Kennedy[4] studied the effect of P and B on


the creep deformation of INCONEL** 718 (IN 718).* They
IN the late 1950s, it was realized that small additions of
boron (B) and zirconium (Zr) were extremely beneficial to **INCONEL is a trademark of INCO Alloys International, Huntington,
the creep-rupture properties of nickel-base superalloys. Since WV.
then, the addition of B in superalloys is generally considered
to be beneficial, especially for the grain boundary sensitive concluded that both P and B increased the resistance to
mechanical properties, such as creep resistance. Work by creep deformation, but the effect of a combined addition of
Pennington[1] showed that B (up to 0.01 pct) and Zr (up to B and P on creep strength of IN 718 was much greater than
0.1 pct) improved high-temperature strength, ductility, and the sum of the individual effects of P and B. Garosshen
notch sensitivity of several alloys, and Darmara[2] observed et al.[5] studied the effect of B, C, and Zr on the mechani-
that as little as 15 ppm B was sufficient to double the rup- cal properties of nickel-base superalloy. They reported that
ture life and ductility of WASPALOY* Much of the earlier both C and B had a strong influence on the formation of
grain boundary precipitates, and B resulted in the forma-
*WASPALOY is a trademark of Special Metals Corporation, Huntington, tion of an intergranular M3B2 boride. Both B and Zr were
WV. observed to be critical for improvement in the mechanical
properties of the alloys, although B levels above the solu-
work on the effect of minor elements in superalloys has been bility limit resulted in no further improvement or reduc-
summarized in a review by Holt and Wallace.[3] tion in strength. Floreen and Davidson[6] studied the effect
of B and Zr on the creep and fatigue crack growth behav-
ior of Ni-base superalloy. They reported that additions of
L. XIAO, Research Associate, and M.C. CHATURVEDI, Distinguished B and Zr markedly improved the smooth specimen creep
Professor and Canada Research Chair, are with the Department of Mechan- properties and the threshold stress intensity values for creep
ical and Manufacturing Engineering, University of Manitoba, Winnipeg, crack growth. However, they had no effect on the crack
MB, Canada, R3T 2N2. Contact e-mail: mchat@cc.umanitoba.ca D.L. growth rates during creep or fatigue deformation. No
CHEN, Associate Professor, is with the Department of Mechanical,
Aerospace and Industrial Engineering, Ryerson University, Toronto, ON,
changes in microstructure, fracture appearance, or grain
Canada, M5B 2K3. boundary sliding behavior due to B and Zr additions were
Manuscript submitted December 20, 2004. observed.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, OCTOBER 2005—2671


The improvement in mechanical properties of superalloys of nickel-base superalloys.[15] It was, therefore, considered
due to the addition of B has been attributed to a variety of necessary to further study the effect of B on the plastic defor-
mechanisms.[1–7] They include (1) decreased grain bound- mation mechanism, especially at room temperature (RT),
ary diffusivity, (2) increased grain boundary interfacial at which the grain boundaries may not play a significant role
strength, (3) lowering of grain boundary surface energy, in plastic deformation of IN 718 and especially during LCF,
(4) removal of tramp elements by precipitating them as sta- which was the primary goal of this work.
ble compounds, (5) formation of denuded zones on either
side of the grain boundaries in which the precipitation of
 was absent or was of a lower volume fraction than that II. MATERIALS AND EXPERIMENTAL
present in the grains, (6) slower agglomeration of grain PROCEDURES
boundary M23C6 carbide concurrently with depletion of 
A. Materials and Heat Treatment
from the adjacent matrix, and (7) modification in fine  or
M23C6 carbide morphologies. Four wrought IN 718 based alloys, which had the same
A large amount of work has been done to understand the base chemical composition of (wt pct) 18.45Cr, 53.43Ni,
effect of B on the grain boundary behavior of IN 718.[3,5,7,8–14] 18.83Fe, 2.91Mo, 4.86Nb, but contained 12, 29, 60, and
It is generally agreed that B segregates to grain boundaries 100 ppm B, respectively, were produced via vacuum induc-
due to its small atomic size and its low solubilities in  and tion melting vacuum arc remelting for this research
. Systematic research work has been conducted by by Special Metal Corporation (Huntington, WV) (12 and
Chaturvedi and co-workers[9–14] to characterize the grain 29 ppm B) and by ALLVAC (Monroe, NC) (60 and
boundary segregation and precipitation in alloy 718. By sec- 100 ppm B). A split 140-kg heat was made. The first half
ondary-ion mass spectrometry (SIMS), Chaturvedi and co- of the heat was poured into a 108-mm electrode with a B
workers obtained direct evidence of atomic segregation of level of 60 ppm. Additional B was added into the remain-
B at high-angle grain boundaries, and also observed B to ing heat to increase the B level to 100 ppm and poured into
form borides.[9–12] Most of Chaturvedi and co-workers’ inves- a 108-mm electrode. Both electrodes were vacuum arc
tigations were concerned with the role of B addition and its remelted into 140-mm ingots. Ingots were homogenized at
effect on the grain boundary behavior during welding and 1163 °C for 16 hours and then press forged to 76 76 mm
not on the mechanical properties of superalloys; however, billet at 1093 °C. The billets were further rolled to 12.7
due to the strength requirements of aircraft engine and power 79 mm plates at 1066 °C. A similar processing procedure
generation turbine disk alloys at high temperatures, where was used to produce alloys with 12 and 29 ppm B. The
the grain boundaries play an important role in the defor- actual chemical composition of these alloys is given in
mation behavior, many significant investigations have been Table I. Cylindrical specimens, 6.25 mm in diameter and 10 mm
conducted by other researchers. However, despite the obvi- in gage length, were machined for the LCF tests. The speci-
ous importance of minor element additions, the actual mech- mens were given a commercial heat treatment. This treatment
anism by which these elements improve mechanical consisted of a solution treatment at 1227 K for 1 hour and
properties remains unclear. To date, only a few studies have then air cooling to RT, followed by aging at 990 K for 8 hours,
reported the influence of B on either the microscopic defor- cooling to 895 K at a rate of 50 K/h, and holding at 895 K
mation mechanism or the low-cycle fatigue (LCF) properties for a total aging time of 8 hours before air cooling to RT.

Table I. Composition of the Selected IN 718 (Weight Percent Unless Designated ppm)

Alloy Mn Si Cr Ni Co Fe Mo
12 0.01 0.02 18.45 53.43 0.01 18.83 2.91
29 0.07 0.08 17.97 53.92 0.71 17.61 2.96
60 0.01 0.01 17.85 53.50 0.01 18.46 2.88
100 0.01 0.01 17.87 53.55 0.01 18.46 2.88
P W V Nb Ti Al B (ppm) C (ppm)
0.002 0.001 0.03 4.86 0.97 0.46 12 120
0.007 0.07 0.02 5.12 0.94 0.47 29 225
0.003 0.01 0.02 5.33 0.99 0.54 60 40
0.003 0.01 0.02 5.33 0.99 0.54 100 40
Alloy Cu Ta Hf Ag (ppm) Pb (ppm) Bi (ppm) Ca (ppm) Mg (ppm)
12 0.001 0.004 0.004 5 3 0.3 50 36
29 0.02 0.005 0.004 5 3 0.3 50 10
60 0.02 0.01 — 2 1 0.1 — 100
100 0.02 0.01 — 1 1 0.1 — 57
S (ppm) Sn (ppm) Cd (ppm) Sb (ppm) O (ppm) N (ppm) Zr
10 20 50 20 3 2 0.001
3 20 50 20 1 65 0.001
4 5 — — 5 22 0.01
3 5 — — 8 22 0.01

2672—VOLUME 36A, OCTOBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


B. Fatigue Tests and Microstructural Examination
The LCF tests were carried out on a computerized Instron
(Canton, MA) 8502 servohydraulic testing system, which was
interfaced with FastTrack II software. Uniaxial pull-push
LCF tests were performed under total strain control mode
in air, employing a symmetrical triangular strain-wave cycle.
A nominal total axial strain rate of 3 10 3 s 1 was used
for all the tests. Axial strain was measured and controlled by
an extensometer of axial gage length of 10 mm. The stress-
strain values were processed to determine the cyclic stress
responses and the plastic strain amplitudes as a function of
number of cycles.
Fractographic analysis was carried out using a JSM-
5200LV scanning electron microscope (SEM). Thin foils
for transmission electron microscopy (TEM) were prepared
from 3-mm discs cut from regions adjacent to the fracture
surface and perpendicular to the loading axis, and twin-jet
electropolished to perforation in a 10 pct solution of per-
chloric acid in methanol, at 243 K temperature and at 12 V.
Fatigue deformation substructures were observed using a
JEOL*-2000FX transmission electron microscope operated
*JEOL is a trademark of Japan Electron Optics Ltd., Tokyo.

at 180 kV.

III. RESULTS
A. Microstructure of the Heat-Treated Material
The microstructure of the heat-treated alloys, containing
different amounts of B, was characterized. It was found that
the grain boundaries in all four of the alloys were exten-
sively decorated with particles of various morphologies, Fig. 1—SEM image of the heat-treated IN 718 with 29 ppm B: (a) SEM
ranging from globular to needle, and a typical SEM micro- micrograph and (b) EDS.
graph of the heat-treated material containing 29 ppm B is
shown in Figure 1(a). Energy-dispersive spectroscopy (EDS) with (010) reflection and presented in Figure 2(d), shows
revealed that these particles were (Ni3Nb) phase, as indi- the disc-shaped   particles with an average diameter of
cated by the arrows in Figure 1(a). Some B containing pre- 30 nm and an average thickness of 5 nm. In addition, some
cipitate particles were observed at grain boundaries, as coherent spherical  precipitates of approximately 5 nm in
suggested by the presence of B peaks in the corresponding diameter were also observed in the dark-field images with
EDS spectrum shown in Figure 1(b). Strong evidence of (010) reflection.
segregation of B at grain boundaries in B containing IN 718
superalloy was observed by SIMS analysis in the earlier
B. Cyclic Deformation Behavior
studies on weldability of IN 718 superalloy.[9–12] The grain
size of all four alloys was about 6 m; however, its distri- The LCF test results on materials with different concen-
bution was observed to be inhomogeneous in some local trations of B are given in Table II. The cyclic stress response
regions. curves as a function of number of cycles for specimens with
IN 718 is a precipitation-hardened superalloy in which different concentrations of B are given in Figures 3(a) and
the strengthening is mainly due to the presence of 13 to (b). It is seen that the material exhibited a relatively short
14 vol pct fraction of coherent   precipitates with an period of cyclic hardening in the early stages, extending to
ordered DO22 type bct crystal structure. It also contains 3 a few cycles, at the higher cyclic strain amplitudes (t /2 
to 4 vol pct of  with L12 type ordered crystal structure. 0.8 pct) followed by a continuous cyclic softening to frac-
A representative thin-film TEM microstructure is shown ture. However, the cyclic hardening was absent and contin-
in Figure 2(a), which is in [001] orientation, as suggested uous cyclic softening was observed as the cyclic strain
by its selected area diffraction pattern (SADP) shown in amplitude decreased to t /2  0.6 pct. A nearly stable peak
Figure 2(b) and its interpretation in Figure 2(c). The sec- stress amplitude was observed at t /2  0.4 pct in all the
ondary spots in the SADP were identified to be due to   alloys. This indicated that the cyclic saturation stage was
and  precipitates. The (100), (010), and (110) reflections reached (Figures 3(a) and (b)). Toward the end of the tests,
in the SADP are both due to the   and  precipitates, the stress amplitude decreased rapidly, indicating the for-
whereas, (1/2 10) and (1 1/2 0) type reflections are only due mation of macrocracks and their subsequent fast growth. A
to   phase.[16,17] The dark-field TEM micrograph, taken closer examination of the plots revealed that the initial cyclic

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, OCTOBER 2005—2673


Fig. 2—TEM micrographs of the heat-treated IN 718 with 29 ppm B showing the presence of   and   precipitates: (a) bright-field image, (b) SAD
pattern with [001] zone axis, (c) corresponding key, and (d) dark-field image with 010 reflection showing   and  precipitates.

Table II. Results of Low-Cycle Fatigue Tests at RT for IN 718 Superalloy with Different Concentrations of B

Specimen
Alloy Parameter 1 2 3 4 5 6
12 (12 ppm B) t /2, pct 0.4 0.6 0.8 1.0 1.2 1.4
p /2, pct 0.0115 0.104 0.276 0.447 0.711 0.886
a, MPa 765.5 964.3 1019 1087.7 1125.9 1174.4
Nf, cycle 13,151 4287 2665 1558 464 247
29 (29 ppm B) t /2, pct 0.4 0.6 0.8 1.0 1.2 1.4
p /2, pct 0.0045 0.115 0.279 0.449 0.6195 0.812
a, MPa 755.6 938.9 995.1 1046 1134.6 1142.2
Nf, cycle 29,068 6988 4273 2186 860 650
60 (60 ppm B) t /2, pct 0.398 0.596 0.794 1.0 1.197 1.4
p /2, pct 0.0035 0.0815 0.2245 0.406 0.584 0.75
a, MPa 761 1003 1064 1122 1158 1190
Nf, cycle 32,195 9360 5732 2682 1187 773
100 (100 ppm B) t /2, pct 0.398 0.596 0.8 1.0 1.197 1.397
p /2, pct 0.0035 0.085 0.253 0.424 0.608 0.776
a, MPa 742 971 1042 1088 1124 1151
Nf, cycle 48,517 10,949 4406 1930 1236 775

2674—VOLUME 36A, OCTOBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 3—Cyclic stress response curves as a function of number of cycles:
(a) alloy with 12 ppm B fatigued at different cyclic total strain amplitudes;
and (b) a comparison of the cyclic stress response of alloys with different
concentrations of B fatigued at selected total strain amplitudes.

stress response value increased as the cyclic strain ampli-


tude increased, and the difference in cyclic stress amplitudes
at different cyclic strain amplitudes was reduced as the cyclic
deformation progressed (Figure 3(a)). Figure 3(b) shows the
effect of B on the cyclic stress response of the material
fatigued at t /2  1.4, 0.6, and 0.4 pct, respectively. It is
observed that at the same cyclic strain amplitude, such as
t /2  1.4 pct, the initial cyclic hardening weakened as B
concentration increased (Figure 3(b)). The cyclic stress ampli-
tude value of the alloy with 60 ppm B was the highest among Fig. 4—Effect of B concentration on the cyclic saturation stress amplitude
the four alloys fatigued at the same cyclic strain amplitude. and the cyclic plastic strain amplitude at the midlife (Nf /2) in IN 718 fatigued
The stress amplitude value of alloys with 12 and 29 ppm B at various cyclic total strain amplitudes: (a) cyclic saturation stress ampli-
was lower than that of the other two alloys. tude vs B concentration; (b) cyclic plastic strain amplitude vs B concen-
tration; and (c) cyclic hysteresis loops at the midlife in the alloys with
Figures 4(a) and (b) show the effect of B on the cyclic different concentrations of B at t /2  1.4 pct.
saturation stress amplitude defined as the cyclic stress ampli-
tude corresponding to the midlife, Nf /2, and plastic strain
amplitude at the midlife of the specimens fatigued at vari-
ous total cyclic strain amplitudes. The plastic strain ampli- stress). It was observed that among the four alloys, for a
tude is equal to the difference between the total cyclic strain given constant total strain amplitude, the cyclic stress ampli-
amplitude and the elastic strain amplitude, i.e., p  tude at the midlife was highest in the 60 ppm B alloy and
t-/E (E is the Young’s modulus, and  is the elastic lowest in the alloy with 29 ppm B (Figure 4(a)). The cyclic

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, OCTOBER 2005—2675


plastic strain amplitude corresponding to the midlife first 12 to 29 ppm B, but only a moderate improvement could
decreased as the B concentration increased from 12 to be observed above 29 ppm B. Furthermore, it is seen in Fig-
60 ppm, and then increased when it increased from 60 to ure 5(a) that the experimental fatigue characteristics of these
100 ppm, as shown in Figure 4(b). Therefore, within the four alloys could be basically rationalized in terms of the
range of B concentrations considered in the present inves- Coffin–Manson type equation relating the total strain ampli-
tigation, the variation in cyclic plastic strain amplitude with tude, t /2, to the fatigue lifetime, Nf. In comparison, the
B concentration displayed a minimum value in the alloy familiar Coffin–Manson law cannot be used to describe the
containing 60 ppm B. Furthermore, the cyclic stress-strain relationship between the cyclic plastic strain amplitude and
hysteresis loop of IN 718 with different concentrations of lifetime over the whole of the strain range examined, as
B at the midlifetime at t /2  1.4 pct is shown in Fig- shown in Figure 5(b). The p /2-Nf curves for all four alloys
ure 4(c). As the B concentration increased from 12 to exhibited two slopes and an inflection point occurred at
29 ppm, the slope of the elastic stress-strain curve in the about p /2  0.40 pct. Using the least-squares fitting
stage of elastic deformation or the Young’s modulus was method, the coefficient and exponent in the Coffin–Manson
observed to decrease slightly in IN 718. This led to a slight equation were determined and are listed in Table III. Eight
decrease in both the cyclic saturation stress and the cyclic Coffin–Manson type relationships were obtained for the four
plastic strain amplitude at a given total strain amplitude. alloys.
When the B concentration increased from 29 to 100 ppm, A comparison of the fatigue lifetime curves of IN 718 as
no obvious change in the Young’s modulus and related hys- a function of cyclic plastic strain amplitude at RT and 650 °C
teresis loops was observed. is shown in Figure 6. The results show that B has a stronger
effect on the fatigue lifetime at 650 °C than at RT. Also,
C. Effect of B on LCF Lifetime of IN 718 the two slopes in the Coffin–Manson plot disappeared at
650 °C.
The LCF lifetime (Nf) curves as a function of cyclic total
strain amplitude (t /2) and the plastic strain amplitude
(p /2) were determined for all four alloys, and are shown D. Fractographic Examination
in Figures 5(a) and (b). The results show that B affects the The fracture surfaces were examined by SEM to gain
fatigue lifetime of IN 718. The LCF life of the alloys insight into the effect of B on the fatigue failure mechanism.
increased with increasing concentration of B, and a rapid The results showed that fatigue crack basically initiated from
improvement occurred when B concentration increased from the specimen surface. Figure 7 shows typical fractographs

Fig. 5—Influence of B concentration on the LCF lifetime of IN 718: (a) t /2 vs Nf and (b) p /2 vs Nf.

Table III. Coffin–Manson Parameters for IN 718 Superalloy with Different B Concentrations

p /2  0.40 pct p /2  0.40 pct


Coffin–Manson Correlation Coffin–Manson Correlation
Alloy Relationship Coefficient Relationship Coefficient
12 p Nf0.3730  6.96 0.9995 p Nf 1.774  2.6 105 0.9832
29 p Nf0.4524  14.29 0.9381 p Nf 1.854
 1.0 106
0.9624
60 p Nf0.4871  18.83 0.9950 p Nf 1.978
 4.0 106
0.9505
100 p Nf0.6613  64.55 0.9847 p Nf 1.516
 6.4 104
0.9466

2676—VOLUME 36A, OCTOBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


the specimens fatigued at low cyclic strain amplitude
(t /2  0.4 pct, Figure 7(a)). However, multiple crack ini-
tiation sites could be observed in the specimens fatigued at
the high cyclic strain amplitude (t/2  1.0 pct, Fig-
ure 7(b)). The typical fractographs in the region of stable
crack propagation are shown in Figures 8(a) through (d).
Fatigue crack propagation in all four alloys was observed
to be characterized primarily by a large number of well-
defined fatigue striations, which was taken as evidence of
transgranular crack growth at RT. A large number of sec-
ondary cracks were observed on the fracture surfaces of
the specimen of the alloy with the lowest (12 ppm) and the
highest (100 ppm) concentrations of B (Figures 8(a) and
(d)). As the B content increased from 12 to 29 ppm, the
number of secondary cracks decreased (Figure 8(b)). Fig-
ure 8(c) depicts a typical fracture surface of the alloy with
60 ppm B. Almost no secondary cracks are observed, but
Fig. 6—A comparison of the effect of B concentration on the LCF life-
plastic deformation traces can be discerned on the facets of
time of IN 718 at RT and 650 °C. the fracture surface. The number of secondary cracks in the
alloy with 100 ppm B (Figure 8(d)) were, however, more
than those observed in the alloy with 29 and 60 ppm B
(Figures 8(b) and (c)).
Figure 9 shows other distinguishing features between the
fracture surfaces of 12 and 29 ppm alloys. It is seen that
characteristic fatigue striations formed on the fracture sur-
face of IN 718 with 29 ppm B (Figure 9(a)), but planar plas-
tic deformation traces were observed on the fracture surface
of IN 718 with 12 ppm B (Figure 9(b)). These traces were
observed to be associated with the intersection of planar slip
bands with the fracture surface or the grain boundaries.

E. Deformation Microstructures
The deformation microstructures produced by LCF in the
specimens with different concentrations of B and fatigued
at similar cyclic plastic strain amplitudes (about 0.45 to
0.60 pct) were examined via TEM. Regularly spaced arrays
of planar deformation bands on {111} slip planes were
observed in the fatigued specimen with 12 ppm B cyclically
deformed at p /2  0.447 pct (Figure 10). They were char-
acterized by the presence of two groups of planar deforma-
tion bands lying along the traces of intersection of (111 ) and
(11 1) slip planes with the (111) thin foil surface, respec-
tively. Trace analysis showed that (11 1) and (1 1 1) slip sys-
tems were activated simultaneously, giving rise to the
saturated diamond-shaped deformation structure in the
fatigued specimens. In addition, some slip bands were
observed to be concentrated in local band regions in the
LCF specimen with the lower B concentration (12 ppm)
(Figure 11). As a result, concentration of plastic deformation
was expected to occur in these areas.
Planar dislocation bands lying on {111} planes were also
observed in the specimens with 29 ppm B at p /2 
0.449 pct (Figure 12). The TEM examination showed that
twinning also occurred in this alloy during cyclic deforma-
Fig. 7—Typical fractographs showing the crack initiation sites in IN 718 tion at RT. Interaction between slip bands and twinning was
with 29 ppm B fatigued at different cyclic strain amplitudes: (a) t /2  observed to occur, as indicated by the arrows in Figure 12.
0.4 pct and (b) t /2  1.0 pct.
It was observed that the intersection of a twin with planar
slip bands resulted in the formation of a staggered step on
in the region of the fatigue crack initiation site in the alloy the planar slip bands, as shown in Figure 12.
with 29 ppm B fatigued at different cyclic strain amplitudes. Figure 13 illustrates the typical deformation microstruc-
The fatigue crack initiation site can be clearly identified in ture that formed in the 60 ppm B alloy fatigued at p /2 

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, OCTOBER 2005—2677


Fig. 8—Typical fractographs in the region of stable crack propagation in IN 718 with different concentrations of B fatigued at t /2  0.4 pct: (a) 12 ppm B,
(b) 29 ppm B, (c) 60 ppm B, and (d) 100 ppm B.

Fig. 9—Fracture surfaces in the region of stable crack propagation in IN 718 fatigued at t /2  0.4 pct: (a) fatigue striations in the alloy with 29 ppm B
and (b) plastic traces in the alloy with 12 ppm B.

0.584 pct. Two groups of planar deformation bands, nearly tion tangles such as cells and walls, which are the typical
perpendicular to each other, were observed in a thin foil with features in fatigued copper single crystals,[18,19] were not
[001] orientation. Trace analysis suggested them to be lying observed. This suggests an absence of traditional persistent
on the (111) and (1 11) planes. Three-dimensional disloca- slip bands in this material with higher B content.

2678—VOLUME 36A, OCTOBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 10—Typical deformation substructure formed in the alloy with 12 ppm B,
p /2  0.447 pct (
g  220; incident beam || [111]).

Fig. 12—Two groups of planar dislocation bands in the alloy with 29 ppm B,
p /2  0.449 pct (
g  022; incident beam || [122]).

Fig. 11—Plastic deformation concentration in some localized band regions in


the alloy with 12 ppm B, p /2  0.276 pct (
g  200; incident beam || [012]).

As the B concentration increased from 60 to 100 ppm,


similar planar slip bands were observed in specimens
fatigued at  p /2  0.6 pct, as shown in Figure 14.
Deformation is seen to be concentrated within the planar
deformation bands in this alloy and the deformation
bands became much more pronounced. The trace analysis
showed that multiple slips on different {111} planes were
activated.
The dark-field TEM examination of   and  precipitates
in the fatigued specimens revealed that they were sheared
in all four alloys during cyclic deformation at RT. An exam-
ple is shown in Figure 15(a), which is the dark-field micro-
graph of a specimen with 100 ppm B cyclically deformed
to p /2  0.6 pct, at RT, and taken with ( 1 10) superlattice Fig. 13—Two groups of dislocation bands nearly perpendicular to each
reflection in the (111) foil (Figure 15(b)). Plastic deforma- other in the alloy with 60 ppm B, p /2  0.584 pct (
g  200; incident
tion traces could be discerned in the microstructure of the beam || [001]).

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, OCTOBER 2005—2679


fatigued specimen, as indicated by arrows in Figure 15(a). the (111) foil plane. In comparison with the nonfatigued
The trace analysis showed the shearing traces to be paral- specimen (Figure 2(d)), a reduction in the size of  
lel to the trace of intersection of the (1 1 1) slip plane with precipitates from 30 to 20 nm was observed after fatigue
deformation (Figure 15(a)). The reduction in the particle size
could be attributed to the particles being sheared repeatedly,
during cyclic deformation. The TEM examinations also
revealed the presence of precipitate-free bands in the fatigued
specimens, as shown in Figure 16, which is a dark-field
image of a fatigued specimen taken with the ( 110) super-
lattice reflection in the (111) foil. The angle between the
traces of three groups of such bands, which coincide with
the traces of (11 1), (1 11), and (1 1 1) planes, with the (111)
foil plane, is equal to 60 deg, as indicated by the arrows in
Figure 16.

IV. DISCUSSION
A. Shearing of   Precipitates, Formation of Planar Slip
Bands, and Cyclic Deformation Softening
A significant cyclic softening was observed at RT in all
four IN 718 superalloys, except when the material was
deformed at the lowest cyclic strain amplitude, t /2 
0.4 pct. This may be attributed to the reduction in the size
of   precipitates by repeated shearing during cyclic defor-
mation in the planar slip mode. The shearing mechanism of
  particles has been discussed by several authors.[16,20–25]
Sundararaman et al.[21] reported that when the size of   par-
ticles in IN 718 superalloy exceeded a critical value
(10 nm),   precipitates were sheared by the deformation
twins, and the smaller precipitates were sheared by the move-
ment of paired dislocations. Clavel and Pineau[23] showed
that the precipitates in IN 718 were sheared in the course
of cyclic straining, and plastic deformation occurred by the
propagation of planar bands, which were identified as twins.
Fig. 14—Two groups of dislocation bands in the 100 ppm B alloy, p /2  However, Oblak and co-workers[16,24] observed that shear-
0.6 pct (
g  020; incident beam || [100]). ing of   particles appeared to take place by the coupled

Fig. 15—Dark-field micrograph with  g [110] reflection and the foil in [111] orientation, showing sheared precipitates in the specimen with 100 ppm B
deformed to p /2  0.6 pct at RT: (a) dark-field image and (b) SAD pattern.

2680—VOLUME 36A, OCTOBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 16—Dark-field micrograph with  g  [110] reflection and the foil in
[111] orientation, showing precipitate-free bands on three groups of {111}
slip planes in a specimen with 100 ppm B deformed to p /2  0.6 pct
at RT.

motion of a/2 110 pairs in IN 718 deformed to 3 pct


strain at RT. Worthem et al.,[25] on the other hand, reported
that shearing of   precipitates occurred by slip only dur-
ing cyclic deformation and no significant shearing of   pre-
cipitates was observed after the monotonic strains, and no
deformation twins were identified.
As an ordered phase, the   precipitates possess an
antiphase domain boundary (APB) energy representing the
extra energy associated with the ordered atomic positions in
the crystal lattice versus the normal disordered or random
atomic positions. An antiphase boundary is created in both
  and   particles in IN 718 superalloy when they are cut
Fig. 17—A schematic illustrating the shearing of  precipitates and the
by a dislocation, and order is restored by the passage of a formation of planar slip bands in IN 718 during cyclic deformation:
second dislocation of the same character. The repulsive (a) schematic representation of the shearing process of  precipitates by
energy between the two dislocations traveling on the same dislocations, where cross-hatching of particles indicates the presence of
slip plane is balanced by the APB energy of the precipitate antiphase boundary; (b) three-dimensional representation of {111} slip planes
phase, resulting in a coupled or superdislocation. The shear- in a thin foil with [111] orientation; (c) projections of {111} planes on (111)
thin foil plane; (d) three-dimensional representation of {111} slip planes
ing of   particles during cyclic deformation is shown in a thin foil in [001] orientation; and (e) projections of {111} planes on
schematically in Figure 17(a). Since cross-slip of the cou- the (001) thin foil plane.
pled dislocations is extremely difficult, slip continues to
occur on the same slip plane during continued deformation,
resulting in the formation of planar slip bands in the fatigued of short-range order in   precipitates caused by the passage
IN 718 specimens, as shown schematically in Figures 17(b) of subsequent dislocations after the passage of first-slip dis-
and (d). It is a three-dimensional representation of all four location, as was observed by Schwander et al. regarding
{111} slip planes in a thin foil in [111] orientation in the Ni-11.2 pct Mo crystals.[26] On the basis of the short-range-
[112]-[110]-[111] space, and Figure 17(c) displays projections ordering parameter, determined by X-ray diffuse scattering
of three {111} planes, i.e., ( 111), (1 11), and (11 1) planes measurements, and the effective pair potentials, obtained by
on a (111) thin foil plane. In other words, Figure 17(c) also applying the inverse Monte Carlo technique, they calculated
represents schematically the projection of planar slip bands the change in configurational energy of Ni-11.2 pct Mo
on the (111) plane. Figure 17(d) shows a similar three-dimen- crystals caused by successive slip by a number of unit dis-
sional representation of all four {111} slip planes in a thin locations on the same {111} plane. They found that a con-
foil in [100] orientation in [1 10]-[110]-[001] space, and figurational energy change of about 25 mJ/m2 was necessary
Figure 17(e) shows a projection of all four {111} planes, i.e., for the first slip, and after five slip steps, the configurational
on a (001) thin foil plane. energy remained constant. That is, the passage of subsequent
In this alloy, an additional factor that may promote pla- dislocations after the first five did not cause any further
nar slip is a reduction in the change in configurational energy change in configurational energy, and planar slip was favored.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, OCTOBER 2005—2681


The continuation of planar slip during cyclic deformation shown in Figure 18. In addition, the cyclic softening stage
due to these two factors would make the passage of dislo- did not occur in the cyclic stress response curves at the low-
cations repeatedly cut the particles, which would cause a est cyclic strain amplitudes (Figures 3(a) and (b)). This would
reduction in their size and may also produce precipitate-free imply that the microdeformation character could be different
bands, as was observed in this investigation (Figure 16). The at the higher cyclic strain amplitudes in comparison with
reduced particle size would decrease the flow stress of the that at the lower strain amplitudes. The tendency to form
material, causing a cyclic softening of the material. planar slip bands in IN 718 was observed to increase as the
cyclic strain amplitude increased. Fracture surface exami-
nation showed that fatigue crack initiated from a single crack
B. Dual-Slope Coffin–Manson Plots of IN 718
origin in the alloy fatigued at lower cyclic strain amplitudes
Superalloys
(Figure 7(a)), whereas multiple crack origins could be
It has been observed that the log-log plots of the cyclic observed in the specimens fatigued at higher cyclic strain
plastic strain amplitude vs fatigue life for many materials are amplitudes (Figure 7(b)). Therefore, based upon the obser-
linear; i.e. they exhibit the well-known Coffin–Manson rela- vations in this investigation and the general conclusions of
tionship. However, such log-log plots of the results obtained others,[30,31] a probable cause of deviation from the single-
in this study exhibited two slopes, as shown in Figure 5(b) slope Coffin–Manson behavior in the fatigue life plot of IN
and Table III. A similar “two-slope” relationship between 718 may be due to the changes in the deformation microstruc-
the fatigue life and the cyclic plastic strain amplitude was tures and fatigue crack initiation mode as a function of the
also reported by Clavel and Pineau,[27] Merrick,[28] and Sanders plastic strain amplitude and B concentration. However, fur-
et al.[29] in IN 718 fatigued at RT. Furthermore, such a two- ther studies are needed to determine the cause of the change
slope behavior was also observed by Sanders et al.[30,31] in in the slopes of Coffin–Manson plots with certainty.
7075 and 7050 type aluminum alloys, which, like IN 718, are
also strengthened by fine precipitates. A departure from the
single-slope behavior in these aluminum alloys has been C. Effect of B on Cyclic Deformation Mechanism and
explained in terms of changes in the substructure induced by Fatigue Lifetime of IN 718
the plastic deformation at higher plastic strain amplitudes, as Boron is an essential alloying element for inducing good
compared to that observed at lower plastic strain amplitudes. creep properties in IN 718, and it was also observed that up
The strain localization in these aluminum alloys was observed to a concentration of 60 ppm, B also significantly enhanced
to occur during cyclic deformation at lower plastic strain the fatigue resistance of IN 718 at RT (Figures 5 and 6). The
amplitudes, while the deformation was more homogeneous effect of B on the LCF life may be due to its effect on pro-
at higher plastic strain amplitudes. moting deformation by planar slip. The TEM examination
In the present study, dense planar slip bands were observed revealed that although planar slip bands were observed to form
to be the typical deformation substructure (Figures 10 and in all four alloys (Figures 10 and 12 through 14), some slip
12 through 14) in specimens deformed at higher cyclic plas- concentration was also observed in some localized regions in
tic strain amplitudes (p /2  0.4 pct). In contrast, the IN 718 with the lower B concentration (Figure 11). In the
typical deformation structure of the alloy with 12 ppm B, alloys with higher B concentrations, in addition to being seg-
deformed at the lower cyclic strain amplitude of t /2  regated at the grain boundaries, as an interstitial element, B
0.4 pct (p /2  0.01 pct), was characterized by a lower may also form clusters in the matrix, which, along with the
density of dislocations on the {111} primary slip planes, as boride precipitate particles, would act as obstacles to disloca-
tion movement. This would impede dislocation slip, and may
render cross-slip at RT even more difficult owing to the extra
stress that would be required for dislocations to escape from
pinning by the clusters of B atoms, and by boride particle in
alloys with higher B concentrations. This would facilitate the
creation of favorable slip paths that requires a lower critical
stress for dislocations to continue to slip on the same slip plane,
thus promoting planar slip, as observed in this study.
It has been suggested that alloys that deform by the for-
mation of fine narrow slip bands are more resistant to fatigue
than those in which deformation is concentrated in a few
banded regions in which coarse slip bands may lead to early
crack initiation.[32,33] The deformation in the alloy with higher
B concentration, however, would be more dispersed, which
would eliminate the deleterious effect of inhomogeneous
distribution of plastic deformation by preventing an early
nucleation of cracks due to the presence of intense slip bands,
thus improving the fatigue resistance of the alloy.[33,34,35]
Since slip concentration was observed only in the alloy con-
taining 12 ppm B, its LCF fatigue life was the lowest, and
improved as the B concentration increased. This is suggested
Fig. 18—A typical deformation structure in the alloy with 12 ppm fatigued to be the reason for the improved fatigue resistance of alloys
at t /2  0.4 pct (p /2  0.01 pct) (
g  220; incident beam || [111]). with an increase in B concentration of up to 60 ppm.

2682—VOLUME 36A, OCTOBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fractographic observations shown in Figure 8 also revealed 0.8 pct). As the cyclic strain amplitude decreased to
that the grain boundary strength was enhanced with an increase t /2 0.6 pct, the period of cyclic hardening was absent
in B concentrations of up to 60 ppm, as evidenced by the and the material showed a continuous cyclic softening
decreasing amount of secondary cracks on the fracture sur- until fracture. A nearly stable cyclic stress amplitude was
face. This is also reflected in the beneficial effect of increas- observed at t /2  0.4 pct in all the alloys.
ing the concentration of B on the fatigue lifetime, as seen in 2. Among the four alloys fatigued at the same total cyclic
Figures 5(a) and (b). The slightly lower fatigue life of the strain amplitude, the cyclic saturation stress amplitude
alloy containing 100 ppm B, compared to the alloy with value was the highest in the alloy with 60 ppm B, and
60 ppm B, may be related to the observed reoccurrence of the lowest in the alloy with 29 ppm B.
secondary cracking in this alloy (Figure 8(d)). The relatively 3. Boron was observed to improve the LCF lifetime of IN
lower amount of B of 60 ppm compared to 100 ppm may 718 at RT. A noticeable improvement occurred up to
improve LCF resistance by distributing the plastic deforma- 29 ppm B, but the improvement was moderate when con-
tion uniformly. The excessive B addition, however, may lead centrations of B exceeded 29 ppm B. The positive effect
to a reduction in the fracture ductility at RT due to an increase of B concentration on the fatigue lifetime at RT was less
in the alloys’ dislocation slip resistance. As a result, the amount than that observed at 650 °C. A linear relationship between
of secondary crack would increase and the LCF life would the fatigue life and cyclic total strain amplitude was
decrease in the alloy with 100 ppm B, as was observed. observed, while a two-slope relationship between the fatigue
The relationship between the microstructure and cyclic life and cyclic plastic strain amplitude was observed to
deformation behavior is known to be complex, especially for occur with an inflection point at about p /2  0.40 pct.
high-strength commercial alloys. It is evident, however, that 4. A fractographic examination showed that fatigue crack
any microstructural feature that causes an inhomogeneous initiated from specimen surfaces, and the crack propa-
distribution of plastic strain will have a deleterious effect on gation mode was predominantly transgranular with a large
the fatigue life.[33,35] Microscopically, strong dislocation pile- number of fatigue striations on the fracture surfaces. The
ups and high stress concentration would lead to fatigue crack number of secondary cracks was observed to be higher
nucleation and favor crack propagation, and would thereby in alloys with 12 and 100 ppm B than that present in
decrease the fatigue life of IN 718 with a lower B content. alloys containing 29 and 60 ppm B.
However, it is worthy of note that research on the effect of 5. The deformation microstructures of all four alloys were
B on LCF of IN 718 at 650 °C[15] has shown that B has a characterized by the presence of groups of planar defor-
stronger influence in improving the fatigue life at 650 °C in mation bands lying along the {111} slip plane traces, and
comparison with that observed at RT (Figure 6). Microstruc- deformation occurred by planar slip. However, some slip
tural examination showed that planar slip bands are the typ- was observed to be concentrated in local band regions
ical cyclic deformation dislocation configuration at both RT in the alloy with 12 ppm B.
and 650 °C. Furthermore, the width and spacing of planar 6. The   precipitates in IN 718 were sheared by the slip of
slip bands observed at RT and 650 °C in foils at [100] ori- dislocations during cyclic deformation. After the initial shear-
entation remained almost constant at different cyclic plastic ing, the trailing dislocations on the same slip plane repeat-
strain amplitudes.[20] The spacing was about 0.70 m and the edly sheared the   precipitates and reduced their size to
width 0.06 m at both RT and 650 °C. Some plastic defor- such an extent that they offered very little or no resistance
mation concentration was observed in the alloy with 12 ppm to the further movement of dislocations. The preferential
B at 650 °C as well.[15] This indicates that the plastic defor- paths, which had fewer precipitates and lower critical stress
mation mode of IN 718 at RT remained similar to that at 650 for dislocation slip, were established. As a result, cyclic
°C. However, no substantial improvement in the fatigue life- deformation softening occurred and planar slip bands formed.
time of IN 718 was observed above 29 ppm B at RT, as com- 7. The improvement in the LCF life of IN 718 by B addi-
pared to a continuous increase in fatigue lifetime observed tion is attributed to its effect on promoting cyclic defor-
at 650 °C from 12 ppm B to 60 ppm B. This implies that mation of IN 718 at RT by planar slip, which caused a
the effect of B is enhanced by its effect on the grain bound- homogeneous distribution of deformation and delayed
ary behavior at elevated temperatures, i.e., B plays a more the initiation of microcracks, thus enhancing the fatigue
important role in improving the high-temperature mechani- resistance of the material.
cal properties of IN 718 as compared to the RT properties.
A significant improvement in the fatigue lifetime of IN 718
at 650 °C due to B microadditions results from improving ACKNOWLEDGMENT
grain boundary cohesion and stabilizing grain boundary con-
stituents via B segregation to grain boundaries and the pre- The authors thank the Natural Sciences and Engineering
cipitation of boride particles on grain boundaries.[15] Research Council of Canada for the financial support.

V. SUMMARY AND CONCLUSIONS REFERENCES


1. W.J. Pennington: Met. Progr., 1958, vol. 73, pp. 82-86.
1. At higher plastic strain amplitudes, all four IN 718 alloys 2. F.N. Darmara, J.S. Huntington, and E.S. Machlin: J. Iron Steel Inst.,
with different concentrations of B exhibited a relatively 1959, vol. 191, pp. 266-275.
3. R.T. Holt and W. Wallace: Int. Met. Rev., 1976, vol. 6, Mar., pp.1-24.
short initial cyclic hardening, with the magnitude increas- 4. Wei-Di Cao and R.L. Kennedy: in Superalloys 718, 625, 706 and
ing with increasing plastic strain amplitude, followed by Various Derivatives, E.A. Loria, ed., TMS, Warrendale, PA, 1997,
a regime of continuous cyclic softening at RT (t /2  pp. 511-20.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, OCTOBER 2005—2683


5. T.J. Garosshen, T.D. Tillman, and G.P. McCarthy: Metall. Trans. A, 20. L. Xiao, D.L. Chen, and M. C. Chaturvedi: Scripta Mater., 2005,
1987, vol. 18A, pp. 69-77. vol. 52, pp. 603-07.
6. S. Floreen and J.M. Davidson: Metall. Trans. A, 1983, vol. 14A, 21. M. Sundararaman, P. Mukhopadhyay, and S. Banerjee: Acta Metall.,
pp. 895-901. 1988, vol. 36, pp. 847-64.
7. A.K. Jena and M.C. Chaturvedi: J. Mater. Sci., 1984, vol. 19, pp. 3121-39. 22. H.F. Merrick: Metall. Trans. A, 1976, vol. 7A, pp. 505-14.
8. U. Franzoni, F. Marchetti, and S. Sturlese: Scripta Metall., 1985, vol. 19, 23. M. Clavel and A. Pineau: Metall. Trans. A, 1978, vol. 9A,
pp. 511-6. pp. 471-80.
9. X. Huang, M.C. Chaturvedi, N.L. Richards, and J. Jackman: Acta 24. D.F. Paulonis, J.M. Oblak, and D.S. Duvall: Trans. ASM, 1969, vol. 62,
Mater., 1997, vol. 45, pp. 3095-107. pp. 611-22.
10. W. Chen and M.C. Chaturvedi: Acta Mater., 1997, vol. 45, pp. 2735-46. 25. D.W. Worthem, I.M. Robertson, F.A. Leckie, D.E. Socie, and C.J.
11. W. Chen, M.C. Chaturvedi, N.L. Richards, and G. McMahom: Metall. Altstetter: Metall. Trans. A, 1990, vol. 21A, pp. 3215-20.
Mater. Trans. A, 2001, vol. 32A, pp 931-39. 26. P. Schwander, B. Schonfeld, and G. Kostorz: Phys. Status Solidi B,
12. X. Huang, M.C. Chaturvedi, and N.L. Richards: Metall. Mater. Trans. 1992, vol. 172, pp. 73-85.
A, 1996, vol. 27A, pp. 785-90. 27. M. Clavel and A. Pineau: Mater. Sci. Eng., 1982, vol. 55, pp. 157-71.
13. W. Chen, N.L. Richards, and M.C. Chaturvedi: Metall. Mater. Trans. 28. H.F. Merrick: Metall. Trans., 1974, vol. 5, pp. 891-97.
A, 2001, vol. 32A, pp. 931-39. 29. T.H. Sanders, Jr., R.E. Frishmuth, and G.T. Embley: Metall. Trans. A,
14. H. Guo, M.C. Chaturvedi, N.L. Richards, and G.S. McMahon: Scripta 1981, vol. 12A, pp. 1003-10.
Mater., 1999, vol. 40, pp. 383-8. 30. T.H. Sanders Jr. and J.T. Staley: Fatigue and Microstructure, Metals
15. L. Xiao, D.L. Chen, and M.C. Chaturvedi: Metall. Mater. Trans. A, Park, OH, 1978, pp. 467-522.
2004, vol. 35A, pp. 3477-87. 31. T.H. Sanders, Jr. and E.A. Starke, Jr.: Metall. Trans. A, 1976, vol. 7A,
16. J.M. Oblak, D.F. Paulonis, and D.S. Duvall: Metall. Trans., 1974, pp. 1407-18.
vol. 5, pp. 143-53. 32. J.C. Grosskreutz: Metall. Trans., 1972, vol. 3, pp. 1255-62.
17. I. Kirman and D.H. Warrington: Metall. Trans., 1970, vol. 1, 33. E.A. Starke, Jr. and G. Lutjering: Fatigue and Microstructure, Metals
pp. 2667-75. Park, OH, 1978, pp. 14-15.
18. C. Laird: Fatigue and Microstructure, ASM, Metals Park, OH, 1978, 34. A. Saxena and S.D. Antolovich: Metall. Trans. A, 1975, vol. 6A,
pp. 149-203. pp. 1809-28.
19. P. Lukas, M. Klesnil, and J. Krejci: Phys. Status Solidi, 1968, vol. 27, 35. L. Xiao, Y. Umakoshi, and J. Sun: Metall. Mater. Trans. A, 2001,
pp. 545-58. vol. 32A, pp. 2841-50.

2684—VOLUME 36A, OCTOBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A

You might also like