Sie sind auf Seite 1von 14

The Effect of Sheet Processing on the Microstructure, Tensile, and Creep Behavior of INCONEL Alloy 718

C.J. BOEHLERT, D.S. DICKMANN, and N.C. EISINGER The grain size, grain boundary character distribution (GBCD), creep, and tensile behavior of INCONEL*
*INCONEL is a registered trademark of Special Metals Corp., Huntington, WV.

alloy 718 (IN 718) were characterized to identify processing-microstructure-property relationships. The alloy was sequentially cold rolled (CR) to 0, 10, 20, 30, 40, 60, and 80 pct followed by annealing at temperatures between 954 C and 1050 C and the traditional aging schedule used for this alloy. In addition, this alloy can be superplastically formed (IN 718SPF) to a significantly finer grain size and the corresponding microstructure and mechanical behavior were evaluated. The creep behavior was evaluated in the applied stress (a) range of 300 to 758 MPa and the temperature range of 638 C to 670 C. Constant-load tensile creep experiments were used to measure the values of the steady-state creep rate and the consecutive load reduction method was used to determine the values of backstress (0). The values for the effective stress exponent and activation energy suggested that the transition between the rate-controlling creep mechanisms was dependent on effective stresses (e a 0) and the transition occurred at e 135 MPa. The 10 to 40 pct CR samples exhibited the greatest 650 C strength, while IN 718SPF exhibited the greatest room-temperature (RT) tensile strength (1550 MPa) and ductility (f 16 pct). After the 954 C annealing treatment, the 20 pct CR and 30 pct CR microstructures exhibited the most attractive combination of elevated-temperature tensile and creep strength, while the most severely cold-rolled materials exhibited the poorest elevated-temperature properties. After the 1050 C annealing treatment, the IN 718SPF material exhibited the greatest backstress and best creep resistance. Electron backscattered diffraction was performed to identify the GBCD as a function of CR and annealing. The data indicated that annealing above 1010 C increased the grain size and resulted in a greater fraction of twin boundaries, which in turn increased the fraction of coincident site lattice boundaries. This result is discussed in light of the potential to grain boundary engineer this alloy.

I. INTRODUCTION

THE potential for improving the bulk properties of both structural and functional materials by manipulating the frequency of special grain boundaries, defined here as low angle boundaries (LABs; boundary misorientations less than 15 degrees) and coincident site lattice boundaries (CSLBs), was first introduced by Watanabe[1] and is considered to be the basis for most current grain boundary engineering (GBE) studies of metallic alloy systems. Over the past two decades, GBE efforts have demonstrated significant improvements in stress corrosion cracking (SCC), fatigue, weldability, and creep of pure Ni and Nibased superalloys.[215] In terms of Ni-based superalloys, efforts have been primarily focused on nonage-hardenable alloys such as alloys 600 and 625, where the latter alloy has exhibited significantly lower creep rates when it is processed to contain a higher fraction of special boundaries. Although creep rates of alloy 738 have also exhibited slower strain rates with increased
C.J. BOEHLERT, Assistant Professor, is with the Department of Chemical Engineering, Michigan State University, East Lansing, MI 48824. Contact e-mail: boehlert@egr.msu.edu D.S. DICKMANN, Manufacturing Engineer, is with Electronic Ceramics, Ferro Corporation, Penn Yan, NY. N.C. EISINGER, Metallurgist, is with the Special Metals Corporation, Huntington, WV. This article is based on a presentation made in the symposium entitled Processing and Properties of Structural Materials, which occurred during the Fall TMS meeting in Chicago, Illinois, November 912, 2003, under the auspices of the Structural Materials Committee.
METALLURGICAL AND MATERIALS TRANSACTIONS A

special boundary fractions,[2] few GBE studies have involved the age-hardenable series of Ni-based superalloys. Krupp et al.[16,17] have shown that the fraction of special boundaries can be significantly increased for IN 718 using cold rolling deformation followed by recrystallization annealing where increased special boundary fractions reduced the sensitivity to oxygen-induced intergranular brittle fracture (dynamic embrittlement). In order to identify processing-microstructure-property relations, the current work evaluated the GBCD of IN 718 as a function of sheet processing using 10 pct cold-rolling increments from 0 to 80 pct followed by annealing temperatures ranging between 954 C to 1050 C, and the elevated-temperature tensile and tensile-creep properties were measured. The approach taken was to characterize the creep behavior within the low-stress creep regimes where diffusional creep and grain boundary sliding may be dominating the strain-rate response. The creep mechanisms that operate during elevated-temperature deformation of pure metals and solid-solutionstrengthened alloys have been related to the value of the stress exponent, na, and the apparent activation energy, Qa, in the Dorn steady-state creep rate equation:[18] # [1] ss AD0 exp (Qa> RT ) mb> kT ( sa > m )na where T is the creep temperature in degrees Kelvin, R is the gas constant, A is a constant, b is the Burgers vector, D0 is the pre-exponential factor, is the shear modulus, and k is the Boltzmann constant. However, the na and Qa values measured for alloys containing dispersed second-phase particles
VOLUME 37A, JANUARY 200627

are generally considerably greater than those observed in pure metals and solid-solution-strengthened alloys.[1931] The na values for precipitation-hardened alloys have ranged between 5 and 15[1924] while, for dispersion-hardened alloys, including thoria-dispersed Ni-20Cr(wt pct), na values have ranged between 9 and 75.[27,30,31] The Qa values have ranged from 1 to 3 times those of the activation energy for self-diffusion.[29,31,32] These variations in the observed values of na and Qa have been rationalized by introducing the concept of a back stress, 0, which is an internal stress opposing the dislocation motion. In multiphase alloys such as IN 718, which contains an austenitic fcc phase matrix () and fine and strengthening precipitates, the applied stress during steadystate creep deformation is opposed by a backstress resulting from the presence of these strengthening particles and a defect structure within the material.[19,27,3338] Therefore, the creep deformation results from an effective stress (e a 0). As a result, the steady-state creep rate can be represented by # ss A*( sas0 )ne [2] where ne is the effective stress exponent. Although traditionally used to measure the back stress during high-stress dislocation power-law creep, it has been shown that the consecutive stress reduction method can also be applied to lowstress diffusional creep for IN 718, where low ne values (2) are observed.[33,34] II. EXPERIMENTAL The IN 718 sheets used in this study were processed at Special Metals Corporation (Huntington, WV). The heats were produced by vacuum induction melting followed by electroslag remelting. The material was hot worked using conventional practices, and the as-processed condition included mill annealing at 1066 C for all hot-rolling procedures that preceded the final CR and 982 C anneal. Subsequent thermomechanical processing (TMP) treatments included CR between 10 and 80 pct. The CR steps were performed on separate sheets each designated with 10, 20, 30, 40, 60, and 80 pct deformation. Metallographic samples were prepared from the sections of the CR sheets prior to annealing, after annealing then water quenching, and after annealing then aging. The annealing treatment was performed at one of the following temperatures: 954 C, 1010 C, or 1050 C. In addition, one set of CR samples was heat treated below the recrystallization temperature at 871 C. The aging treatment, used to precipitate out the and strengthening phases, consisted of 718 C/8 h/furnace cool to 621 C then hold at 621 C for a total aging time of 18 hours. In addition, a separate heat of IN 718SPF was produced in a similar fashion; however, the sheet cold working procedure, estimated to total between 55 and 80 pct deformation, was altered to assure the production of an ultrafine grain size product.[3942] Electron backscattered diffraction (EBSD) orientation maps, obtained using an accelerating voltage of 25 keV and a step size of 0.5 m on a PHILIPS 515 SEM* with a LaB6 filament,
*PHILIPS is a trademark of Philips Electronic Instruments Corp., Mahwah, NJ.

preparatory polishing step included greater than 20 minutes using 0.06-m colloidal silica. EDAX-TSL, Inc. (Draper, UT) manufactured the EBSD hardware and software. Brandons criteria[43] were used to distinguish between the grain boundary types. The reported fractions of random general high-angle boundaries (GHABs), LABs, CSLBs, and twins (3) were the averaged values taken from several orientation maps, performed on the cross sections, rolling faces, or longitudinal sections of the processed sheets, of areas typically greater than 1.4 mm by 0.7 mm. Flat dogbone-shaped tensile and creep specimens with a cross section of approximately 1 12 mm and a gage length of 25 mm were machined, using either a mill or an electrodischarge machine, with the tensile axis parallel to the rolling direction. The tensile experiments were performed in air at room temperature (RT) or 650 C using an Instron 8562 (Norwood, MA) machine and a strain rate of approximately 1.3 104s1. In most cases, multiple tests were performed and the reported strength and elongation values were averaged. Constant-load creep experiments were performed on Applied Test Systems Incorporated (Butler, PA) lever-arm creep apparati, using a 20:1 load ratio, in air at temperatures ranging between 638 C and 670 C and applied stresses ranging between 300 and 758 MPa. The creep strain was monitored during the tests using a linear variable differential transformer attached to the gage section. The specimen temperature, monitored by three thermocouples located within the gage section during the creep experiments, was maintained within 3 C using a single-zone furnace. The 0 values were determined at 638 C by the consecutive stress reduction method.[19,22,26] When the creep rate for a given a remained constant for at least 5 hours, it was assumed the steady-state creep rate had been achieved. Thereafter, the sample was subjected to a small stress reduction (approximately 5 pct a). This resulted in an elastic contraction of the sample, followed by an incubation period with a zero creep rate. After a period of time, creep began again at a lower rate. Once steady state was reached, another stress reduction was performed. The time of the incubation period following each stress reduction was recorded. The remaining stress vs the cumulative incubation time was plotted, and 0 was determined by taking the asymptotic value of the remaining stress when the cumula ss tive incubation time appeared to be infinite. The 0 and values proved to be repeatable as duplicate samples were ss tested at the same temperature and a and the measured values were within 5 pct of each other. In addition, the 0 ss values were not dependent on strain history for total and creep strains less than 0.5 pct as several temperature/applied stress conditions were performed, some in duplicate, before that of the backstress condition, and in each case, similar 0 values were recorded. Creep rupture experiments were performed in air at 758 MPa and 649 C. All the creep and 650 C tensile experiments were initiated after soaking the samples at the desired testing temperature for a minimum of 0.5 hours. III. RESULTS A. Microstructure The chemical composition range of the IN 718 heats used is shown in Table I. The annealed microstructures contained an equiaxed -phase (fcc) austenitic matrix and after aging
METALLURGICAL AND MATERIALS TRANSACTIONS A

were obtained for the CR samples as well as CR samples which were subsequently annealed then water quenched. The final
28VOLUME 37A, JANUARY 2006

Table I. Composition Range for the IN 718 Heats Used in This Study Element Wt Pct Element Wt Pct Ni 53.48 to 53.68 C 0.03 Ti 1.01 to 1.06 Fe 17.99 to 18.3 Mo 2.99 Cu 0.02 to 0.17 Co 0.03 to 0.12 Si 0.01 to 0.17 Cr 18.1 to 18.4 Mn 0.04 to 0.12 Al 0.46 to 0.48 Nb 5.07 to 5.11 P 0.009 to 0.012 S 0.001

Table II. Grain Size of IN 718 as a Function of CR and Annealing 954 C Annealed CR, Pct 0 10 20 30 40 60 80 SPF d, m 36.2 37.5 30.6 28.9 28.7 23.0 20.0 12.0 Standard Deviation 6.6 8.2 6.5 3.4 2.5 4.0 3.2 1.4 1050 C Annealed d, m 92.0 77.5 89.1 76.4 83.9 101.8 99.8 83.9 Standard Deviation 11.0 14.0 15.3 10.1 8.9 13.3 13.3 5.7

Table III. RT Hardness of 0 to 40 Pct CR then 954 C Annealed-Then-Water-Quenched IN 718 CR, Pct Hardness, Rb 95.1 94.0 93.7 94.6 94.2 Equivalent Hardness, Hv 220 215 214 218 216

(a)

0 10 20 30 40

Table IV. RT Properties of the 871 C Heat-TreatedThen-Aged Samples CR, Pct 0 10 20 30 40 YS, MPa 1167 1193 1346 1446 1497 UTS, MPa 1435 1433 1496 1557 1606 RT f , Pct 20.7 22.5 17.0 15.1 12.9 Hardness, Hv 434 432 460 470 478

(b)
Fig. 1(a) Low-magnification and (b) high-magnification SEM photomicrographs of the cross section of a 0 Pct cold rolled then 871 C annealed and aged microstructure illustrating the austenitic -phase matrix, fine and precipitates, and -phase precipitates (white).

fine (coherent spherical fcc (L12)) and (coherent ordered disc-shaped body-centered tetragonal (DO22)) precipitated throughout (Figure 1). The average grain size was measured through the line-intercept method (Table II).
METALLURGICAL AND MATERIALS TRANSACTIONS A

1. Annealing temperature effects Annealing temperature had a significant effect on grain size. Above 1010 C, grain growth occurs[44] and the 1050 C, 1-h annealed samples exhibited significantly larger grain sizes than those annealed at 954 C (Table II). The 1050 C annealed microstructures also exhibited an increased volume fraction of twin boundaries compared with the 1010 C and 954 C annealed microstructures; this will be discussed in Section 2. It is noteworthy that the RT hardness values significantly increased with CR deformation for the 871 C heat-treated samples, while hardness remained almost constant after 954 C annealing (Tables III and IV). This indicated that 871 C is below the austenite (-phase) recrystallization temperature for 0-40 pct CR IN 718, which is consistent with previous findings.[45] The relatively constant hardness values for the 954 C annealed samples (Table II) indicated that the annealing temperature was above the recrystallization temperature and the
VOLUME 37A, JANUARY 200629

quenching treatment successfully avoided formation of a significant volume fraction of the strengthening precipitates.[44] Overall, the hardness values agreed well with the RT tensile strength of the 871 C heat-treated-then-aged samples according to the established hardness-tensile strength relationship for IN 718[46] (data in Table IV). 2. Grain boundary character distribution Table V lists the GBCD parameters, including total fraction of CSLBs, LABs, GHABs, and 3 boundaries, as a function of CR and annealing temperature. The crystallographic grain texture, measured using pole figure analysis taken from EBSD orientation maps (Figure 2), appeared to increase slightly for increased rolling deformation where the maximum intensity for the preferential orientation of the 40 pct CR condition was approximately twice that for the baseline 0 pct CR condition. This corresponded to an increase in the LAB fraction of 0.03 in the 0 pct CR condition to 0.36 in the 40 pct CR condition (Table V). However, after annealing, the LAB fractions decreased to less than 0.3 for the 40 pct CR condition, while the LAB fractions were even lower for the 0 to 30 pct CR conditions. The maximum intensity for the preferred orientation of the IN 718SPF 954 C annealed material was within 1.5 times that of the baseline condition. In addition, no significant GBCD differences were evident with respect to the sheet orientation, and therefore, the average CSLB, LAB, GHAB, and 3 boundary fractions used were taken from each of the three sheet orientations. It is noted that GHABs dominated the GBCD of the as-cold-rolled samples as well as the samples annealed at
Table V. GBCD Parameters of the Rolled and RolledThen-Annealed Samples CR, Pct As-rolled 0 10 20 30 40 954 C 0 10 20 30 40 60 80 IN 718SPF 1010 C 0 20 30 40 1050 C 0 10 20 30 40 60 80 IN 718SPF CSLBs 0.415 0.496 0.215 0.195 0.158 0.459 0.387 0.342 0.362 0.321 0.304 0.213 0.327 0.454 0.252 0.213 0.152 0.350 0.476 0.549 0.412 0.417 0.543 0.502 0.563 GHABs 0.552 0.478 0.665 0.564 0.483 0.516 0.565 0.563 0.601 0.636 0.653 0.755 0.626 0.516 0.594 0.614 0.606 0.616 0.496 0.430 0.556 0.553 0.426 0.447 0.399 LABs 0.033 0.026 0.120 0.241 0.359 0.025 0.048 0.095 0.037 0.043 0.043 0.032 0.047 0.030 0.154 0.173 0.241 0.034 0.028 0.021 0.032 0.030 0.031 0.051 0.038 3 0.279 0.382 0.194 0.141 0.109 0.350 0.287 0.261 0.261 0.223 0.209 0.1 0.232 0.344 0.170 0.120 0.073 0.233 0.333 0.440 0.312 0.328 0.435 0.348 0.342

954 C and 1010 C (Table V). The LAB fractions were the least prevalent and always less than 0.1. For the 0 to 80 pct CR materials annealed at 954 C, increased CR increased the fraction of GHABs and decreased the fraction of CSLBs, which were dominated by twins (Table V and Figure 3). The GHABs ranged from 0.5 for the baseline 0 pct CR microstructure to 0.75 for the 80 pct CR microstructure. Other than for 3 boundaries, no particular x (1 x 29) boundary fraction was greater than 0.07. The annealing temperature had a significant effect on GBCD. The fraction of twin boundaries increased significantly for the 1050 C annealed samples compared with the 954 C and 1010 C annealed samples. For example, the twin boundary fractions increased from 0.21 to 0.26 (954 C annealing) to 0.44 (1050 C annealing) for the 20 pct CR and 60 pct CR microstructures (Table V). Opposite to that of the 954 C annealed samples, the GHABs decreased and the CSLBs and twin boundary fractions increased with increased CR for the 1050 C annealed samples (Figure 4). Representative EBSD data are illustrated in Figure 5, which highlights twin boundaries, LABs, and GHABS for a 1050 C annealed 60 pct CR material. Figure 6 illustrates the distribution of CSLBs in this microstructure. It is noted that the GBCD was not affected by aging or creep exposure, as such samples exhibited a similar distribution of GHABs, LABs, and CSLBs as those of the annealed-then-quenched samples. B. Mechanical Behavior 1. Tensile behavior Figure 7 illustrates the tensile stress/strain behavior of IN 718SPF in the 954 C annealed and 954 C annealed-thenaged conditions, where the significant strengthening effect of the fine precipitates is evident (Table VI). For both the annealed and annealed-then-aged conditions, half of the samples represented in Figure 7 were machined in an orientation 90 deg with respect to the rolling direction, indicating that strength was not significantly effected by sheet orientation. This may be expected based on the lack of a strong texture in these severely deformed microstructures, as previously described. With aging, the tensile strength increased dramatically at the expense of elongation-to-failure (f), though the f values were always greater than 12 pct, and a ductile fracture was evident for all the samples tested (Figure 8). The RT strengths of IN 718SPF approached those exhibited by the 20 to 30 pct CR deformation-hardened material, which was not annealed but heat treated at 871 C prior to aging (compare Tables IV and VI). The 650 C yield strength (YS), ultimate tensile strength (UTS), and f values are presented in Table VII for the annealed-then-aged samples as a function of CR. The strength of the IN 718SPF material decreased significantly from RT to 650 C. For the 954 C annealing treatment, the 40 pct CR material exhibited the greatest YS value, and the 80 pct CR and IN 718SPF samples exhibited significantly lower YS values. Thus, CR increases the elevated-temperature strength; however, the limit to this strengthening appears to be 40 pct CR. For the 1050 C annealed-then-aged samples, this was less evident as the YS values ranged within a tighter band (865 to 921 MPa), and the 10 pct CR materials YS was only slightly greater than that for the IN 718SPF material. The 650 C f values exceeded 8 pct for nearly all the samples tested.
METALLURGICAL AND MATERIALS TRANSACTIONS A

30VOLUME 37A, JANUARY 2006

Fig. 2Pole figures taken from EBSD data for a 1050 C annealed 60 pct CR sample. Strong texturing was not evident even after a large amount of CR deformation.

Fig. 3The fraction of GHABs, LABs, CSLBs, and 3 boundaries as a function of cold rolling for the 954 C annealed samples. The IN 718SPF data were input as 55 pct CR.

Fig. 4The fraction of GHABs, LABs, CSLBs, and 3 boundaries as a function of cold rolling for the 1050 C annealed samples. The IN 718SPF data were input as 55 pct CR.

2. Creep behavior During the creep experiments, the strain-time plots illustrated the three stages of creep: primary, secondary, and tertiary. The dependence of the steady-state creep rate on a is
METALLURGICAL AND MATERIALS TRANSACTIONS A

illustrated in Figure 9. The na values were similar to those observed previously for IN 718 by Han and Chaturvedi,[33] whose data were interpolated to 638 C and included in Figure 9(a) and Table VIII. Their data were for an as-processed
VOLUME 37A, JANUARY 200631

(a)

(b)
Fig. 5EBSD orientation map for a 1050 C annealed 60 pct CR sample. (a) Normal direction EBSD inverse pole figure map where the colors represent the samples normal direction indexed to the fcc unit triangle. (b) Image quality map highlighting twin boundaries (yellow), LABs (red), and GHABs (blue).

Fig. 7Stress vs strain plot for RT tensile tested IN 718SPF samples, which were either 954 C annealed then water quenched or 954 C annealed then aged. Fig. 6CSLB distribution chart for the 1050 C annealed 60 pct CR sample represented in Figs. 2 and 5.

Table VI. RT Tensile Properties of IN 718SPF Heat Treatment 954 C anneal then water quenched 954 C annealed then aged

condition, and they closely resembled the 0 pct CR data in the current work. Note that the na values, which ranged between 8 and 40, were similar to those measured for other particlestrengthened alloy systems[1931] and are considerably larger than those generally observed for pure metals. For the 954 C annealed-then-aged samples, it was apparent that there were two clusters of data, where the highest strain rates were exhibited by the samples CR to more than 40 pct and lower strain rates were exhibited by the 0 to 40 pct CR samples. This was not the case for the 1050 C annealed-then-aged samples.
32VOLUME 37A, JANUARY 2006

y, MPa
682 1333

UTS, MPa
1063 1554

f, MPa
982 1496

f, Pct 33.6 15.8

a. Creep backstress Figure 10 illustrates plots used to determine 0, while # Figure 11 illustrates the corresponding ss vs e plots. The # values of ne, which are listed along with ss and 0 in
METALLURGICAL AND MATERIALS TRANSACTIONS A

Table VII. 650 C Tensile Properties of IN 718 CR Deformation, Pct 954 C anneal* 0 10 20 30 40 60 80 SPF 1050 C anneal* 0 10 20 30 40 60 80 SPF

0.2 Pct YS, MPa


988 957 1002 959 1054 1050 948 991 899 921 878 906 888 865 876 913

UTS, MPa
1145 1115 1184 1161 1153 1130 1057 1056 1028 1041 1044 1054 1024 1036 1054 1088

f, Pct 13.5 10.8 11.4 12.6 12.7 6.1** 19.6 4.8** 10.1 8.1 7.4 14.0 11.8 12.7 15 12

(a)

*All samples were aged according to the aging treatment 718 C for 8 h followed by furnace cooling to 621 C and holding for total aging time of 18 h. **Sample broke out of the gage section.

ited the greatest 0 value, 630 MPa, while the 30 pct CR condition exhibited the lowest 0 value, 505 MPa (Figure 10(b)). Similar to that for the 954 C annealed-then-aged condition, there was a transition in the ne values at e 135 MPa. Thus, this trend occurred independent of both annealing temperature and CR deformation amount. This result suggests that the active creep mechanisms may be more dependent on e than on a. It is noteworthy that for most samples, the 0 value was over half of the YS at 650 C, and the 0 value was as high as 66 pct for the 30 pct CR 954 C annealed-then-aged condition. b. Creep rupture The creep rupture data, listed in Table X, indicated that increased CR tended to increase Tr. For the 954 C annealedthen-aged samples, the greatest Tr and f values were exhibited by the 30 and 40 pct cold-rolled samples, where both the Tr and f values were greater than twice those for the baseline 0 pct and 10 pct CR conditions. Cold rolling below 30 pct did not offer as significant of an increase in the creep rupture properties and both the 10 pct and 20 pct CR conditions resulted in lower f values than that for the baseline 0 pct CR condition. Each of the ruptured samples exhibited ductile dimpling throughout the fractured surface (Figure 13). The 871 C annealed samples also exhibited increased Tr values with increased CR, and the Tr values were significantly greater than the 954 C annealed samples. This is considered to be a result, in part, of the greater tensile strengths exhibited by the 871 C heat-treated materials, which, as previously discussed in Section III-A-1, were always maintained below the recrystallization temperature. IV. DISCUSSION A. Microstructure The frequency of special boundaries can be significantly increased in Ni-based superalloys having a wide variety of
VOLUME 37A, JANUARY 200633

(b)
Fig. 8Fractographs for RT tensile tested IN 718SPF: (a) 954 C annealed and (b) 954 C annealed-then-aged samples.

Tables VIII and IX, were determined from Eq. [2]. For the 954 C annealing treatment, the 20 pct CR and 30 pct CR samples exhibited the greatest 0 values, 645 and 630 MPa, respectively, while the 60 pct CR, 80 pct CR, and IN 718SPF conditions exhibited the lowest 0 values, which were less than half those of the highest 0 values. Correspondingly, # the ss values for a given a were the lowest for the 20 pct CR and 30 pct CR materials (Table VII). For e less than 135 MPa, the ne values were between one and two, while for e greater than 135 MPa, ne was greater than 3.5 (Table VIII and Figure 11). The Qa value (304 10 kJ/mol) was determined for a 20 pct CR 954 C annealed-then-aged sample using the ln ss vs 1/T plot shown in Figure 12. For the 1050 C annealed-then-aged condition, the 0 values fell in a narrower band, 450 to 630 MPa, than those for the 954 C annealed-then-aged condition. The IN 718SPF sample exhibMETALLURGICAL AND MATERIALS TRANSACTIONS A

(a)

(b)
# Fig. 9ss vs e plot for the (a) 954 C annealed-then-aged creep samples tested at 638 C. Also included are data from Han and Chaturvedi,[33] # which has been interpolated to 638 C. (b) ss vs e plot for the 1050 C annealed-then-aged creep samples tested at 638 C.

compositions and strengthening mechanisms (i.e., precipitation, solution, etc.). For alloys 625 and 738, TMP has increased the special boundary fractions from nominally 0.1 to 0.2 to 0.5 to 0.7, and the resulting microstructures have exhibited sufficient thermal stability to sustain prolonged exposure to service operating temperatures of 850 C.[2,8] In Alloy 600, TMP has increased the special boundary fraction from 0.15 to 0.7 without significantly affecting grain size or adding texture.[9,47,48] In the current work, TMP increased the fraction of special boundaries for IN 718 from less than 0.4 to 0.6 without significantly adding texture, where both the annealing temperature and cold work played significant roles. 1. Grain boundary character distribution For the as-rolled samples in the 10 to 40 pct CR deformation range, increased CR tended to decrease the fraction of 3 and CSLBs. Recrystallization annealing at temperatures of 1010 C and 954 C did not change this trend, and GHABs maintained the majority volume fractions for all the CR conditions. The overall result was that the GHABs increased and
34VOLUME 37A, JANUARY 2006

the special boundary fractions subsequently decreased with increased CR. On the other hand, the 1050 C annealing treatment, which is above the -phase grain growth temperature (1020 C[40]), had a different effect on the microstructure and in particular grain size and GBCD. The fraction of twin boundaries tended to increase with increased CR deformation for the 1050 C annealed samples (Figure 4), and this resulted in increased CSLBs and decreased GHABs with increased CR deformation. It is noted that some variability existed within this trend as the 20 pct CR data exhibited similar GHABs as the 60 pct CR and IN 718SPF materials; these were the three microstructures that exhibited the greatest fractions of special boundaries (0.6). Independent of annealing treatment the LAB fraction was always less than 0.1 and relatively unaffected by CR amount. Achieving a majority of special boundaries therefore may be achieved for IN 718, where annealing temperature is a vital factor. It is noted that the corresponding GBCD, and in particular twin boundary fractions, compared favorably to that measured by Krupp et al.[16,17] for cold-worked then 1050 C annealed IN 718. Overall, increasing the grain size through higher-temperature annealing resulted in increased twin boundary fractions, which is consistent with observations for other Ni-based superalloys, including Waspaloy.[49] As most of the twins exhibited were complete twins (Figure 5), rather than island, edge, or incomplete twins,[50] this is an understandable result, as increased twin boundary fractions will result not only from the increased number of twins but also the increased boundary area of the twins in the microstructure. For other GBE-processed Ni-base superalloys such as alloys 625 and 738, special boundary fractions similar to those found in the current work have led to significantly improved creep resistance including lower creep rates and increased Tr, without a deterioration in tensile strength or ductility.[2] The creep data for the 1050 C annealed-thenaged samples indicate significantly decreased creep rates and increased back stress values for the IN 718SPF condition. Thus, this processing condition may be optimal for creep, indicating there may be fertile ground to GBE IN 718 as well as other age-hardenable Ni-base superalloys. It is noted that the creep deformation did not alter the GBCD significantly, as deformed samples exhibited similar GHABs, LABs, CSLBs, and 3 values as the undeformed microstructures. Although it has been shown here that IN 718 can be TMP to a majority of special boundary fractions similar to other Ni-based superalloys systems, the grain boundary statistics indicate that these microstructures do not compare well with the twin-limited ideal microstructure, which is composed entirely of special grain boundaries (i.e., 2/3 are twins) as discussed by Davies and Randle.[51] The twin-limited microstructure has been approached in other investigations of Ni-based alloys and austenitic steels but never achieved,[52,53] while the current results, where the twin boundary fractions saturated at 0.44, are similar to most findings including those for oxygenfree electronic (OFE) Cu and IN 600 alloy.[47] Thus, the desired maximum value of twins in the microstructure may not be achievable through commercially practical TMP treatments. B. Mechanical Behavior 1. Tensile behavior The measured RT strength and f values of IN 718SPF were intermediate to those published by Smith and Yates[39] for two
METALLURGICAL AND MATERIALS TRANSACTIONS A

Table VIII. The 638 C Creep Data of IN 718 Samples Annealed at 954 C Then Aged CR Deformation, Pct 0

a, MPa
574 574 594 595 609 611 632 649 674 591 596 610 634 666 694 666 676 683 648 668 680 578 593 596 598 610 613 613 627 378 399 417 436 455 471 490 501 537 564 598 612 375 397 417 457 489 523 620 673 696 720 334 374 405

0, MPa
540 540 540 540 540 540 540 540 540 570 570 570 570 570 570 645 645 645 630 630 630 550 550 550 550 550 550 550 550 310 310 310 310 310 310 310 310 310 310 310 310 300 300 300 300 300 300 524 524 524 524 305 305 305

a 0, MPa
34 34 54 55 69 71 92 109 134 21 26 40 64 96 124 21 31 38 28 38 50 28 43 46 48 60 63 63 77 68 89 107 126 145 161 180 191 227 254 288 302 75 97 117 157 189 223 96 149 172 196 29 69 100 2.1 1.9 3.5 3.8 4.4 4.6 6.0 9.1 1.1 1.0 1.1 2.8 5.1 1.1 1.9 3.8 7.3 1.0 1.5 5.6 8.4 2.4 2.9 2.8 3.3 4.7 5.2 4.8 6.6 3.3 4.2 5.2 7.2 9.9 1.2 2.0 3.3 5.8 7.9 1.2 1.3 2.3 3.0 3.8 6.0 9.7 2.4 5.9 1.2 1.7 2.3 2.1 3.3 6.0

ss 109 109 109 109 109 109 109 109 108 109 109 109 109 108 108 109 109 108 109 109 109 109 109 109 109 109 109 109 109 109 109 109 109 109 108 108 108 108 108 107 107 109 109 109 109 109 108 109 108 108 108 109 109 109

na 10.8 10.8 10.8 10.8 10.8 10.8 10.8 10.8 10.8 18.3 18.3 18.3 18.3 18.3 18.3 39.8 39.8 39.8 36.0 36.0 36.0 13.7 13.7 13.7 13.7 13.7 13.7 13.7 13.7 4.6 4.6 4.6 10.4 10.4 10.4 10.4 10.4 10.4 10.4 10.4 10.4 4.5 4.5 4.5 10.3 10.3 10.3 9.0 9.0 9.0 9.0 5.3 5.3 5.3

ne 1.2 1.2 1.2 1.2 1.2 1.2 1.2 1.2 1.2 1.7 1.7 1.7 1.7 1.7 1.7 1.7 1.7 1.7 1.7 1.7 1.7 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.0 1.0 1.0 3.6 3.6 3.6 3.6 3.6 3.6 3.6 3.6 3.6 1.1 1.1 1.1 3.9 3.9 3.9 1.9 1.9 1.9 1.9 1.1 1.1 1.1

10

20 30 40

60

80

Han and Chatervedi[34] interpolated to 638 C

SPF

chemically different heats of IN 718SPF. However, similar to their results, the RT tensile strengths of the IN 718SPF 954 C annealed-then-aged samples were significantly greater than those for conventionally processed IN 718,[44] and this is most likely a result of the ultrafine grain size. In fact, the IN 718SPF material exhibited the greatest RT properties of all the recrystallization-annealed materials studied. The only processed samMETALLURGICAL AND MATERIALS TRANSACTIONS A

ples that exhibited greater strengths were the 30 pct and 40 pct CR deformation-hardened samples, which were heat treated at 871 C prior to aging (Table IV). However, the elevatedtemperature properties of IN 718SPF degrade more rapidly with temperature than the 0 to 40 pct CR samples as the strength of IN 718SPF decreased 30 pct from RT to 650 C, compare the data in Table VI and VII.
VOLUME 37A, JANUARY 200635

(a) (a)

(b) (b)
Fig. 10Remaining stress vs cumulative incubation time for the cold-rolled and (a) 954 C annealed-then-aged and (b) 1050 C annealed-then-aged samples. # Fig. 11ss vs e plot for the cold-rolled and (a) 954 C annealed-thenaged and (b) 1050 C annealed-then-aged samples. The data were used to calculate the listed ne values.

Increasing the annealing temperature was also not beneficial to the elevated-temperature strengths, as the 1050 C annealed samples tended to exhibit lower strengths than the 954 C annealed samples. This is expected to be a result of the -phase grain-size discrepancy. Based on the similar YS values between the 10 pct and 40 pct CR conditions, it is expected that 10 to 40 pct CR does not have a significant effect on elevated-temperature strength, and therefore, the creep behavior for the 10 to 40 pct CR conditions may be compared without considering YS effects. It is noted, however, that CR 10 to 40 pct does significantly increase the elevated-temperature strength from the baseline 0 pct CR condition. 2. Creep behavior a. Effective stress exponent Han and Chaturvedi[33,34] observed an incubation time after stress reduction during creep in both the power-law and the diffusional-creep regimes for IN 718. According to Harris[54]
36VOLUME 37A, JANUARY 2006

and Burton,[55] for diffusional creep to continue, the stress concentration at the precipitate-matrix interface created by the entrapment of diffusing vacancies can only be relieved by the formation of prismatic dislocation loops. However, Ansel and Weertman[56] have suggested that diffusional creep in two-phase alloys can only occur by the process of dislocation climb over the precipitate particles. Therefore, diffusional creep in two-phase alloys may not only involve vacancy diffusion in the matrix and dislocation motion in the grain boundary region but also dislocation creation and motion within grains. Han and Chaturvedi[33] observed dislocation networks within grains, where it was suggested that the observed incubation time could be due to the activation of BardeenHerring sources.[57] Therefore they concluded that the consecutive stress reduction method can be used to determine 0 during power-law creep as well as diffusional creep. The experimental results of the current work indicate two separate regimes based on the effective stress level. The ne values were measured to be between one and two for e less than 135 MPa, and ne was 3.0 to 4.3 for e greater than
METALLURGICAL AND MATERIALS TRANSACTIONS A

Table IX. 638 C Creep Data of IN 718 Samples Annealed at 1050 C Then Aged CR Deformation, Pct 0

a, MPa
614 638 658 668 591 616 639 642 627 648 668 696 715 579 580 601 627 597 617 617 638 643 681 691 733 598 619 620 639 682 707 740 592 615 634 656 678 655 661 682 688

0, MPa
530 530 530 530 560 560 560 560 590 590 590 590 590 505 505 505 505 550 550 550 550 550 550 550 550 520 520 520 520 520 520 520 570 570 570 570 570 630 630 630 630

a 0, MPa
84 108 128 138 31 56 79 82 37 58 78 106 125 74 75 96 122 47 67 67 88 93 131 141 183 78 99 100 119 162 187 220 22 45 64 86 108 25 31 52 58

ss 4.2 109 5.2 109 8.8 109 1.1 108 2.6 109 9.1 109 1.8 108 1.6 108 3.1 109 4.2 109 9.0 109 2.3 108 3.5 108 4.8 109 5.2 109 7.7 109 9.3 109 5.2 109 7.2 109 7.1 109 9.7 109 8.7 109 1.6 108 1.7 108 3.7 108 4.3 109 5.0 109 4.8 109 6.6 109 1.8 108 3.6 108 6.8 108 2.35 109 3.78 109 6.22 109 8.05 109 1.09 108 3.5 109 5.1 109 1.5 108 1.6 108

na 11.4 11.4 11.4 11.4 22.7 22.7 22.7 22.7 19.5 19.5 19.5 19.5 19.5 8.1 8.1 8.1 8.1 8.4 8.4 8.4 8.4 8.4 8.4 13.1 13.1 6.3 6.3 6.3 6.3 16.0 16.0 16.0 11.4 11.4 11.4 11.4 11.4 31.6 31.6 31.6 31.6

ne 1.9 1.9 1.9 1.9 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 1.3 1.3 1.3 1.3 1.1 1.1 1.1 1.1 1.1 1.1 3.0 3.0 1.0 1.0 1.0 1.0 4.3 4.3 4.3 1.0 1.0 1.0 1.0 1.0 1.3 1.3 1.3 1.3

10

20

30

40

60

80

SPF

135 MPa. Based on this result, the creep mechanism may be dependent on e, and the low-stress regime may be dominated by diffusional creep or grain boundary sliding. The ne values of the current work are in agreement with the findings of Ansel and Weertman[56] and Han and Chaturvedi[34] and promote the suggestion that differing creep mechanisms may be active and dependent on effective stress for particle-strengthened alloys. b. Activation energy The measured Qa value (304 kJ/mol) was greater than both the activation energy for self-diffusion (265 to 280 kJ/mol) and the activation energy for the creep process (276 kJ/mol) of pure Ni[58] and the creep process of Ni-Cr in solid solution (295 kJ/mol).[59] However, the creep rate expression used to calculate these values considers neither the influence of temperature on the value of G nor the concept of back stress. Using the back stress, lower activation energies than those calculated using the applied stress have been calculated,[34] and such values lie within the range expected for lattice selfMETALLURGICAL AND MATERIALS TRANSACTIONS A

diffusion. Thus, a similar result would be expected based on the Qa measured here, and self-diffusion is considered to be more likely than grain boundary diffusion for creep of IN 718 in the temperature range of 638 C to 670 C. c. Backstress For the 954 C annealed-then-aged samples, the significant drop in 0 with increased CR deformation beyond 40 pct was correlated with a decreasing average grain size. It also corresponded to a significant decrease in the special boundary fractions (Figure 3). It has been hypothesized that an increase in the special boundary fraction decreases the annihilation rate of dislocations in the grain boundary, leading to an increase in 0, a decrease in the e, and therefore a reduction in the overall creep strain rate.[10,11] The CSLBs, and in particular twin boundaries, are not considered to be effective sources and sinks, and a boundary close to a CSL may require a finite excess concentration of vacancies before it will reabsorb them.[60,61] The closer the orientation to that of a CSL, the
VOLUME 37A, JANUARY 200637

values observed in the current work suggest that 0 may be dependent on both special boundary fractions and grain size. It has been suggested that the backstress arises from the inability of dislocations to bow between adjacent particles; i.e., at stresses below the Orowan bowing stress, dislocations will not be able to bow between adjacent particles and positive strain will not occur. For example, according to Evans and Wilshire,[63] Ni20Cr2ThO2 single crystals will not creep at stresses below its Orowan bowing stress, and polycrystalline Ni20Cr2ThO2 will not creep until its Orowan bowing stress is exceeded locally. These results suggest that an effective stress, which is equal to the applied stress minus the Orowan bowing stress, governs the creep behavior. The Orowan bowing stress, bow, can be approximated as tbow Gb> l
# Fig. 12In ss vs 1/T plot used to calculate Qa for a 20 pct CR 954 C annealed-then-aged sample.

[3]

Table X. Creep Rupture Properties at 649 C/758 MPa for the 954 C* and 871 C** Annealed Aged Samples CR Deformation, Pct 0 10 20 30 40 Tr ,* h 24.9 27.5 38.5 64.8 57.9 f,* Pct 9.3 6.7 7.9 28.9 38.0 Tr ,** h 73.6 65.7 99.8 103.8 128.9 f,** Pct 26.1 25.2 22.0 18.2 23.8

where G is the shear modulus, b is the burgers vector, and is the mean interparticle spacing. It is important to note that varies with particle size. Using G 64 GPa,[44], b 0.3 nm,[64] and 44 to 56 nm[65], bow values range from 340 to 440 MPa. The significantly higher back stress values observed in this study (up to 645 MPa) suggest that there must be other defect/dislocation interactions or diffusion barriers that contribute to the back stress in addition to the Orowan bowing stress. d. Creep rupture Increased CR tended to increase Tr and f for 954 C annealed-then-aged samples. The IN 718SPF material was not evaluated in creep in this study, but based on previous creep rupture data,[39] its Tr value is expected to be similar to that exhibited by the 20 pct CR condition. Thus, there appears to be a limit to the amount of CR deformation that will result in increased creep rupture life and f. This limit appears to be near 30 pct CR, as this condition exhibited the maximum Tr value and a decrease in Tr was observed at 40 pct CR. It is noted that a significant decrease in the grain size occurred with increased CR from the 40 pct CR condition to the IN 718SPF condition, and this may be a significant factor in the creep rupture discrepancy. Although previous findings have correlated a significant increase in creep rupture properties with increased special boundary fractions,[2] further investigation of the effect of GBCD on creep rupture properties is necessary to indicate the importance of grain boundaries to the creep resistance under these creep stress and temperature conditions. Comparing the creep strain rate and rupture properties, it appears that the 30 pct CR condition for the 954 C annealed-then-aged samples results in the most attractive overall creep behavior. V. SUMMARY AND CONCLUSIONS

Fig. 13Fractograph for a creep rupture sample which was 0 pct CR then 954 C annealed-then-aged sample.

higher this excess, and this may slow the rate of creep.[62] For the 1050 C annealed-then-aged samples, the IN 718SPF condition exhibited both the highest 0 value and special boundary fraction (0.6) of all the conditions studied. The 0
38VOLUME 37A, JANUARY 2006

IN 718 was processed through sequential increments of CR between 0 to 80 pct followed by annealing between 954 C to 1050 C then aging to evaluate processing-microstructureproperty relationships. For annealing temperatures of 954 C and 1010 C, increased CR led to decreased twin boundary fractions. This resulted in decreased CSLBs and increased GHABs with increased CR deformation. An opposite trend
METALLURGICAL AND MATERIALS TRANSACTIONS A

was observed for 1050 C annealed microstructures, which exhibited increased twin boundary fractions and CSLBs and decreased GHABs with increasing CR deformation. This observation was correlated to -phase grain size as the grain size increased significantly with increased annealing temperature from 954 C to 1050 C. The lowest GHAB fraction observed was 0.4, and therefore, the greatest special boundary fraction recorded was 0.6. This indicates the potential to GBE IN 718 as a majority of special boundaries can be obtained through commercially practical TMP processing treatments. The steady-state creep rate, backstress, and creep-rupture properties were measured for the 954 C annealed-then-aged samples. For the 954 C annealed-then-aged microstructures, the greatest backstress and lowest creep rate values were exhibited by the 20 pct and 30 pct CR microstructures, and the 30 pct CR microstructure exhibited the greatest Tr value. For the 1050 C annealed-then-aged microstructures, the greatest backstress and lowest creep rate values were exhibited by the IN 718SPF microstructure. The effective stress exponent values, which incorporated the backstress, suggested that the creep deformation mechanism is dependent on effective stresses (e) where the transition point occurs at e 135 MPa. This was independent of both CR deformation and annealing temperature. The significantly higher backstress values in relation to the estimated Orowan bowing stress suggests there must be other defect/dislocation interactions or diffusion barriers that contribute to the backstress. The RT tensile results revealed the exceptional strength and adequate elongation exhibited by the fine-grained IN 718SPF material. For the 954 C annealed-then-aged samples, the 10 to 40 pct CR microstructures exhibited the greatest 650 C tensile strengths, while no clear trend in 650 C tensile strength was exhibited with CR deformation for the 1050 C annealedthen-aged condition. Regardless of processing condition, the tensile and creep specimens exhibited a ductile fracture. Overall, the 20 pct and 30 pct CR samples exhibit the best elevated-temperature properties in the 954 C annealed condition, while the IN 718SPF exhibited exceptional strength and creep resistance in the 1050 C annealed condition, on par with the 20 to 30 pct CR 954 C annealed samples. ACKNOWLEDGMENTS This work was supported by the NSF Division of Materials Research (Grant No. DMR-0533954). The IN 718SPF sheets were processed under the supervision of Gaylord Smith (SMC), who along with James Crum (SMC) offered helpful guidance for this work. The authors are grateful to Dr. Dingqiang Li and Messieurs Serkan Civelekoglu and Jay Spike for their technical assistance. REFERENCES
1. T. Watanabe: Res Mechanica, 1984, vol. 11, pp. 47-84. 2. E.M. Lehockey, G. Palumbo, and P. Lin: Metall. Mater. Trans. A, 1998, vol. 29A, pp. 3069-79. 3. B. Alexandreanu, B.M. Capell, and G. Was: Mater. Sci. Eng., 2001, vol. 300A, pp. 94-104. 4. E.M. Lehockey, G. Palumbo, P. Lin, and A.M. Brennenstuhl: Scripta Mater., 1997, vol. 36 (10), pp. 1211-18. 5. C. Cheung, U. Erb, and G. Palumbo: Mater. Sci. Eng., 1994, vol. 185A, pp. 39-43.

6. D. Crawford: Ph.D. Thesis, University of Michigan, Ann Arbor, MI, 1991. 7. G. Palumbo and K.T. Aust: Materials Interfaces: Atomic Level Structure and Properties, D. Wolf and S. Yip, eds, Chapman and Hall, New York, NY, 1989, pp. 190-211. 8. G. Palumbo, E.M. Lehockey, and P. Lin: J. Met., 1998, Feb., pp. 40-43. 9. P. Lin, G. Palumbo, U. Erb, and K.T. Aust: Scripta Mater., 1995, vol. 33 (9), pp. 1387-92. 10. G.S. Was, V. Thaveeprungsriporn, and D.C. Crawford: J. Met., 1998, Feb., pp. 44-49. 11. V. Thaveeprungsriporn and G. Was: Metall. Mater. Trans. A, 1997, vol. 28A, pp. 2101-12. 12. E.M. Lehockey and G. Palumbo: Mater. Sci. Eng., 1997, vol. A237, pp. 168-72. 13. G. Palumbo and K.T. Aust: Acta Metall. Mater., 1990, vol. 38 (11), pp. 2343-52. 14. B. Alexandreanu, B.H. Sencer, V. Thaveeprungsriporn, and G. Was: Acta Mater., 2003, vol. 51, pp. 3831-48. 15. H. Guo, M.C. Chaturvedi, N.L. Richards, and G.S. McMahon: Scripta Mater., 1999, vol. 40 (3), pp. 383-88. 16. U. Krupp W.M. Kane, C. Laird, and C.J. McMahon, Jr.: Mater. Sci. Eng., 2003, vol. A349, pp. 213-17. 17. U. Krupp W.M. Kane, X. Liu, O. Ducber, C. Laird, and C.J. McMahon, Jr.: Mater. Sci. Eng., 2004, vols. A387A89, pp. 409-13. 18. A.K. Mukherjee, J.E. Bird, and J.E. Dorn: Trans. ASM, 1969, vol. 62, pp. 155-79. 19. K.R. Williams and B. Wilshire: Met. Sci. J., 1973, vol. 7, pp. 176-79. 20. D. Sidney and B. Wilshire: Met. Sci. J., 1969, vol. 3, p. 56. 21. J.D. Parker and B. Wilshire: Met. Sci. J., 1975, vol. 9, p. 248. 22. P.W. Davies, G. Nelmes, K.R. Williams, and B. Wilshire: Met. Sci. J., 1973, vol. 7, pp. 87-92. 23. W.J. Evans and G.F. Harrison: Met. Sci. J., 1976, vol. 10, p. 307. 24. R. Lagneborg and B. Bergman: Met. Sci. J., 1976, vol. 10, p. 20. 25. D.D. Sherby and P.M. Burke: Prog. Mater. Sci., 1968, vol. 13, pp. 323-90. 26. J.D. Parker and B. Wilshire: Met. Sci., 1978, Oct., pp. 453-58. 27. R. Lund and W.D. Nix: Acta Metall., 1976, vol. 24, pp. 469-81. 28. C.N. Ahlquist and W.D. Nix: Acta Metall., 1971, vol. 19, pp. 373-85. 29. C.L. Meyers, J.C. Shyne, and O.D. Sherby: Aust. Inst. Met., 1963, vol. 8, p. 171. 30. B.A. Wilcox and A.H. Clauer: Acta Mater., 1972, vol. 20, pp. 743-57. 31. A.H. Clauer and B.A. Wilcox: Met. Sci. J., 1967, vol. 1, p. 86. 32. B.A. Wilcox and A.H. Clauer: Met. Sci. J., 1969, vol. 3, p. 26. 33. Y. Han and M.C. Chaturvedi: Mater. Sci. Eng., 1987, vol. 85, pp. 59-65. 34. Y. Han and M.C. Chaturvedi: Mater. Sci. Eng., 1987, vol. 89, pp. 25-33. 35. W. Chen and M.C. Chaturvedi: Mater. Sci. Eng., 1994, vol. A183, pp. 81-89. 36. J.C. Gibeling and W.D. Nix: Mater. Sci. Eng., 1980, vol. 45, pp. 123-35. 37. S. Purushothman and J.K. Tien: Acta Mater., 1978, vol. 26, p. 519. 38. J.H. Hausselt and W.D. Nix: Acta Mater., 1977, vol. 25, pp. 595-607. 39. G.D. Smith and D.H. Yates: Proc. Advancements in Synthesis and Processes, Society for the Advancement of Material and Process Engineering, Covina, CA, 1992, pp. M207-M218. 40. G.D. Smith and H.L. Flower: Proc. Superalloys 718, 625, 706 and Various Derivatives, TMS, Warrendale, PA, 1994, pp. 355-64. 41. B.A. Baker: INCO Alloys International Technical Investigation Report No. BAB1323093, INCO, Huntington, WV, Sept. 1993. 42. Y. Huang and P.L. Blackwell: Mater. Sci. Technol., 2003, vol. 19, pp. 461-66. 43. D.G. Brandon: Acta Metall., 1966, vol. 14, pp. 1479-84. 44. INCONEL Alloy 718 Bulletin 4th ed., Special Metals Corporation, Huntington, WV, 1985, pp. 1-25. 45. W.C. Liu, Z.L. Chen, and M. Yao: Metall. Mater. Trans., 1999, vol. 30A, pp. 31-40. 46. Special Metals Corporation, Huntington, WV, unpublished research, 1964. 47. M. Kumar, W.E. King, and A.J. Schwartz: Acta Mater., 2000, vol. 48, pp. 2081-91. 48. A.J. Schwartz and W.E. King: J. Met., 1998, Feb., pp. 50-55. 49. M. Qian and J.C. Lippold: Acta Mater., 2003, vol. 51, pp. 3351-61. 50. V. Randle: The Role of Coincidence Site Lattice in Grain Boundary Engineering, The Institute of Materials, London, 1996, pp. 1-104. 51. P. Davies and V. Randle: Mater. Sci. Technol., 2001, vol. 17, pp. 615-26. 52. D. Lin, G. Palumbo, and K.T. Aust: Scripta Mater., 1997, vol. 36, pp. 1145-49.

METALLURGICAL AND MATERIALS TRANSACTIONS A

VOLUME 37A, JANUARY 200639

53. V.Y. Gertsman and J.A. Szpunar: Scripta Mater., 1998, vol. 38, pp. 1399-1404. 54. J.E. Harris: Met. Sci. J., 1973, vol. 7, p. 1. 55. B. Burton: Mater. Sci. Eng., 1973, vol. 11, pp. 337-43. 56. G.S. Ansel and J. Weertman: Trans. TMS-AIME, 1959, vol. 215, p. 838. 57. D. Hull: Introduction to Dislocations, Pergamon, Oxford, United Kingdom, 1975, pp. 188-90. 58. J.P. Dennison, R.J. Llewellyn, and B. Wilshire: J. Inst. Met., 1967, vol. 95, p. 115.

59. 60. 61. 62. 63.

D. Sidey and B. Wilshire: Met. Sci. J., 1969, vol. 3, p. 56. M.F. Asby: Scripta Mater., 1969, vol. 3, pp. 837-42. H. Gleiter: Acta Metall., 1979, vol. 27, pp. 187-92. F.R.N. Nabarro: Metall. Mater. Trans. A, 2002, vol. 33A, pp. 213-18. R.W. Evans and B. Wilshire: Creep of Metals and Alloys, A The Institute of Metals, New York, NY, 1985. 64. C. Mercer, A.B.O. Soboyejo, and W.O. Soboyejo: Mater. Sci. Eng., 1999, vol. A270 (2), pp. 308-22. 65. J.M. Oblack, D.F. Paulonis, and D.S. Duvall: Metall. Trans., 1974, vol. 5, pp. 143-53.

40VOLUME 37A, JANUARY 2006

METALLURGICAL AND MATERIALS TRANSACTIONS A

Das könnte Ihnen auch gefallen