Effect of powder preparation technology on the hot deformation behavior of HIPed P/M nickel-base superalloy FGH96
M.J. Zhang,F.G. Li,S.Y. Wang,C.Y. Liu
DOI: https://doi.org/10.1016/j.msea.2011.01.118
2011-01-01
Abstract:Research highlights ► HIPed P/M nickel-base superalloy FGH96 prepared by plasma rotation electric pole (PREP) powder exhibits high flow stress at lower temperatures (<1050 °C). ► The effect of preparation technology on hot deformation behavior is slight at higher temperatures (>1050 °C) or lower strain rates (<0.01 s −1 ). ► The activation energy decreases with the increase of strain. ► The activation energy of the alloy prepared by PREP powder is higher than that prepared by argon atomization (AA) powder. ► The heat treatment time has a significant influence on recrystallized grain size for the alloy prepared by AA powder. Abstract In this paper, the hot deformation characteristics of P/M nickel-base superalloy FGH96 prepared by different powder preparation technologies were studied in the deformation temperature range from 1000 °C to 1100 °C and the strain rate range from 0.001 s −1 to 1 s −1 using hot compression tests. The peak stress vs. deformation temperature curves and the peak stress vs. strain rate curves were established, respectively. The results show that the specimens prepared by plasma rotation electric pole (PREP) powder were more sensitive to deformation temperature and strain rate. On the basis of the dynamic material model, the processing maps for hot working were developed. The activation energies and Zener–Hollomon parameters were obtained by linear statistical regression method. For the specimens prepared by PREP powder, the peaks of power dissipation mainly located in lower temperature domain (1000–1030 °C), and the efficiencies of power dissipation (EPD) obtained in the strain range from 0.1 to 0.7 were essentially similar. This indicated that strain had a slight influence on processing maps. For the specimens prepared by argon atomization (AA) powder, the effects of strain on EPD and instability domains were significant. The lower activation energies and Z values indicated that the workability of the specimens prepared by AA powder is better than that prepared by PREP powder. Moreover, it was found that effects of the heat treatment time on activation energy and Zener–Hollomon parameter were significant. With the increase of heat treatment time, the dislocation density and the volume fraction of precipitation phase gradually decreased. Microstructural observation demonstrated that the phenomenon of recrystallized grains coarsening existed in the specimens prepared by longer heat treatment time. The heat treatment time of the specimens prepared by AA powder should be appropriately shortened in order to prevent recrystallized grains coarsening. Abbreviations AA argon atomization PREP plasma rotation electric pole PPBs prior particle boundaries EPD efficiency of power dissipation HIP hot isostatic pressing DRX dynamic recrystallization DRV dynamic recover DMM dynamic materials model GB grain boundary Keywords P/M nickel-base superalloy FGH96 Preparation technology Processing map Peak stress Activation energy Zener–Hollomon parameter 1 Introduction The nickel-base superalloy FGH96 (nominal composition Ni–15Cr–4W–3Mo–2Al–3Ti) is a type of damage-tolerant nickel-base superalloy prepared by power metallurgy (PM). In aviation industry, the uniform microstructure, high working temperature and low crack growth rate make it widely applicable for turbine disks [1,2] . However, microstructural instability and cracking may be generated during deformation owing to higher strain rate sensitivity index and narrower forging temperature range. In order to improve the workability of this alloy, the further investigation must be conducted focused on the microstructural deformation mechanism. It has been accepted that the initial microstructure obtained by different powder metallurgy technologies and heat treatment regimes significantly affected the workability of the materials during hot working. At present, two types of powder preparation technologies applied to the P/M nickel-base superalloy FGH96 are plasma rotation electric pole (PREP) and argon atomization (AA). PREP technology uses plasma arc to melt high-speed rotating alloy bar, and makes liquid drops cooled in inert gases. Owing to centrifugal force, the differences between two types of powders are mainly in particle shape, density and size range. In general, PREP powder exhibits uniform spherical shape and has less hollow particles. Thus, the microstructure prepared by above powder preparation technologies may be influenced. The main production methods for the P/M nickel-base superalloy are as follows: hot isostatic pressing (HIP), direct hot extrusion, HIP + forging, hot extrusion + forging, HIP + heat treatment and HIP + hot extrusion. However, the prior particle boundaries (PPBs) were annihilated during hot extrusion and forging because the dynamic recrystallization (DRX) would occur when deformation was enough large, which lead to inexplicit effect of powder preparation technologies. By comparing, the P/M nickel-base superalloy prepared by HIP could preserve PPBs [3] . Thus, the investigation was conducted based on HIP in this paper. For optimizing microstructure and hot working technology, it is very important to investigate the effect of initial microstructure on the deformation behavior of the P/M nickel-base superalloy FGH96 based on different powder preparation technologies. It is well known that processing map is beneficial for optimizing hot working processes and controlling microstructure in this type of material. Reddy and Srinivasan [4] investigated the hot deformation behavior of spray formed + HIPed Al–Li alloy using processing map, and found that the grain boundary cavitations and flow localization may cause cracking processes in instability domains. Kim and Kwon [5] studied on the surface defect prediction in the hot bar rolling based on processing map, and the processing maps were derived based on two available approaches. Gronostajski et al. [6] investigated the processing map of CuAl9·2Fe3 aluminium bronze and pointed out that the dynamic recover (DRV) was dominating at the temperatures below 645–680 K, and above these temperatures, DRX was dominating. Venugopal and Sivaprasad [7] used a series of industrial hot forming operations to validate the predictions of the processing map of 304L. The results showed that excellent correlation existed between processing maps and microstructures. Krishna and Prasad [8] pointed out that the occurrence of superplasticity was attributed to two-phase micro-duplex structure, higher strain rate sensitivity and lower flow stress by researching the processing maps of near-α alloy 685. Liu and Hu [9] found that strain had a slight influence on processing maps of Haynes230, and Cai et al. [10] also obtained the same conclusion for a nickel-base superalloy. Wang and Lu [11] investigated the processing maps of Ti–6.5Al–3.5Mo–1.5Zr–0.3Si alloy, and pointed out that the optimum deformation parameters were at a deformation temperature of 1055 °C and a strain rate of 0.001 s −1 . Although the relation between processing map and hot deformation behavior was studied for other materials, the effect of initial microstructure on workability was less reported on the basis of different preparation technologies. In this paper, the hot deformation behavior of HIPed P/M nickel-base superalloy FGH96 was investigated based on processing map and activation energy. The basis for processing map and the problem in applying processing map theory to the material are presented. The processing map is developed on the basis of the dynamic materials model (DMM), which is recently reviewed by Prasad [12] and Gegel [13] . According to DMM, processing map is consisted of power dissipation map and instability map. The efficiency of power dissipation (EPD) is a dimensionless parameter and reflects the ratio of dissipated power used for microstructure evolution. And, it is a function of deformation temperature and strain rate in the power dissipation map [14–16] . The variation of EPD with temperature and strain rate could exhibit some specific microstructural mechanisms. In general, high EPD is usually related to DRX in hot deformation [3] . The absorbed power P in plastic deformation process can be expressed as: (1) P = σ ⋅ ε ˙ = G + J = ∫ 0 ε ˙ σ d ε ˙ + ∫ 0 σ ε ˙ d σ where σ is the flow stress (MPa) and ε ˙ is the strain rate (s −1 ). The power caused by plastic deformation is expressed by G , and the dissipated power caused by microstructure evolution is indicated by J [17,18] . The power dissipation J can be defined as: (2) J = ∫ 0 σ ε ˙ d σ = m m + 1 σ ε ˙ where m is the strain rate sensitivity index. The EPD η is a ratio of power dissipation to its maximum value. It is given by the following equation: (3) η = J J max = 2 m m + 1 According to plastic flow continuity principle, material instability may occur when the value of d D / d ε ˙ is lower than that of D / ε ˙ . Based on DMM, the stability function is as follows [19–21] : (4) ξ ( ε ˙ ) = ∂ log ( m / m + 1 ) ∂ log ε ˙ + m when ξ ( ε ˙ ) < 0 , the flow instability procedure will execute. It can be seen that the parameter ξ ( ε ˙ ) is regarded as the function of temperature and strain rate. The variation of the instability parameter ξ ( ε ˙ ) with temperature and strain rate constitutes the instability map. Generally, microstructural manifestations of flow instability are adiabatic shear bands formation, flow localization, dynamic strain aging mechanical twinning and kinking or flow rotations [8] . So, the processing map consisted by power dissipation map and instability map is useful for optimizing hot workability and producing desired microstructure. 2 Experimental The main chemical compositions (wt.%) of the P/M nickel-base superalloy FGH96 used for investigation are shown in Table 1 . The powder was produced by PREP and AA, respectively. Fig. 1 (a) shows the powder prepared by PREP. It can be seen that the powder was nearly spherical in shape, and its size was more uniform than that of the powder prepared by AA ( Fig. 1 (b)). The average sizes of PREP powder and AA powder used for HIP were 65 μm ( Fig. 2 (a) ) and 40 μm ( Fig. 2 (b)), respectively. The powder was encapsulated in stainless steel capsules, and then HIP was conducted in a single stage at the temperature of 1200 °C and the pressure of 120 MPa. The optical micrographs of the specimens prepared by PREP powder and AA powder are shown in Fig. 3 (a) and (b), respectively. Table 2 shows the heat treatment time for different kinds of the specimens. In the table, the preparation technologies for different specimens were represented by the characters as follows: PREP-3h, PREP-5h, AA-3h and AA-5h, respectively. The cylindrical specimens of 8 mm diameter and 15 mm height were machined for hot compression testing. During machining, it was ensured that the edges of the specimens were chamfered to avoid fold-over in the initial stages of compression. In order to reduce friction, concentric grooves were provided on the faces of the specimens to ensure effective lubrication during compression. A computer controlled servo-hydraulic thermecmaster-Z machine was used for hot compression tests. This machine was equipped with an exponential decay of the actuator speed so that constant true strain rates in the range from 0.001 s −1 to 1 s −1 could be imposed on the specimens. The specimens were induction heated to the test temperatures at a heating rate of 15 °C/s and held for 5 min to ensure temperature uniformity prior to deformation. Hot compression experiments were conducted under isothermal condition in the deformation temperature range from 1000 °C to 1100 °C and the strain rate range from 0.001 s −1 to 1 s −1 . During deformation, the temperature was controlled within ±2 °C. All of the specimens were deformed to the true strains of 0.5 and 0.7, and cooled to room temperature immediately. The flow stress in the isothermal compression of PREP-3h, PREP-5h, AA-3h and AA-5h as a function of strain rate and deformation temperature is given in Tables 3–6 , respectively. 3 Results and discussion 3.1 Stress–strain behavior A series of stress–strain curves of the P/M nickel-base superalloy FGH96 deformed in the deformation temperature range from 1000 °C to 1100 °C and the strain rate range from 0.001 s −1 to 1 s −1 are shown in Fig. 4 (a)–(d) . The figures show that the flow stresses exhibit single peaks at certain strains and then gradually decrease or even keep constant at the high strain zone. Initial fast increase in stress is associated with an increasing in dislocation density and the formation of poorly developed sub-grain boundaries, as a result of working hardening and DRV. For the P/M nickel-base superalloy, the dislocations piled up and tangled around strengthening phase γ ’ may more easily lead to the increasing stress. However, with the further increase of strain, high dislocation density may stimulate the occurrence of DRX once a critical strain is exceeded. Fig. 4 (a)–(d) shows that all of the specimens exhibit continuous flow softening phenomenon after peaks, which may be associated with DRX. Compared with the specimen AA-5h, the specimen PREP-3h exhibits higher flow stress at lower temperatures (<1050 °C). However, the difference among the flow stresses of the different specimens gradually reduced with the increase of deformation temperature ( Fig. 4 (b)–(d)). This indicates that their deformation behaviors tend to be uniform at higher temperatures (>1050 °C). Liu et al. [22] investigated the hot deformation behavior of strengthening phase γ ’ and pointed out that the solution temperature of the phase fluctuated in a temperature range from 1088 °C to 1125 °C. So, the identical deformation behavior in high temperature may be associated with the decreasing volume fraction of strengthening phase γ ’ . Moreover, Fig. 4 (b) shows that the flow stresses exhibit the same slight uptrend in the steady-state stage at the temperature of 1100 °C, and the specimens PREP-5h and AA-5h exhibit higher flow stresses. This may be related to the interaction of potentially mobile dislocations with other dislocations, grain boundaries, or the periodic friction of the lattice itself which determines the rate flow. 3.2 Effects of temperature and strain rate on peak stress Since continuous flow softening after peaks is associated with DRX, the value of peak stress represents the threshold value of energy used for DRX in a sense. The Fig. 5 (a) shows the curves of deformation temperature vs. peak stress at the strain rate of 0.1 s −1 . The curves of that at other strain rates are essentially similar. The curves became dispersive at lower temperatures (<1050 °C). This is because that dislocation density, grain boundaries (GBs), strengthening phase γ ’ (Ni 3 (Al,Ti)) and grain sizes significantly affect the hot deformation behavior of the nickel-base superalloy at lower temperatures. For the P/M nickel-base superalloy, preparation technology determines the dislocation density and grain size of the initial microstructure. So, it can be seen that the effect of preparation technology on DRX is significant at lower temperatures. However, owing to the increasing thermal diffusion of metal atom and the decreasing volume fraction of strengthening phase γ ’ , the effect is weakened with the increase of deformation temperature. At the temperature of 1100 °C, the overlapped curves indicate that the specimens have the identical deformation mechanism. The curves of strain rate vs. peak stress at the deformation temperature of 1000 °C are shown in Fig. 5 (b). The curves of that at other temperatures are essentially similar. It can be seen that the curves overlap at the strain rate of 0.001 s −1 and disperse at higher strains (>0.1 s −1 ). This indicates that the effect of preparation technology on peak stress is slight at lower strain rates because DRV is dominant and makes dislocation density lower. However, with the increase of strain rate, the increasing dislocation density leads to the strengthened interaction of dislocations although this may stimulate the occurrence of DRX. Thus, it can be seen that the effect of preparation technology on DRX is significant at higher strain rates (>0.1 s −1 ). Moreover, Fig. 5 (a) and (b) show that the peak stress of specimen PREP-3h exhibits higher sensitivity to deformation temperature and strain rate. This indicates the variations of dislocation density of specimen PREP-3h with temperature and strain rate are prominent. On the basis of above analysis, it can be seen that the effects of strain rate and temperature on the hot deformation behavior of various specimens are different. It is known that DRX, flow instability and cracking may be also shown flown soften behavior. So, it is not very appropriate to predict the deformation mechanism on the basis of the shapes of the flow curves alone, although they may be correlated with the mechanisms established by other methods. 3.3 Processing maps The domains of the processing maps could be interpreted based on the characteristic efficiency variation associated with the microstructural processes [23–25] . The deformation mechanisms of the safe domains include DRV, DRX and superplasticity. In fact, DRX was shown to be more efficient process to keep the flow stresses and rates of work hardening considerably low [26] , which is thought to be a preferred choice for hot working. Generally, the efficiency values associated with DRX are about 30–50%, while the values relate to superplasticity are higher [27,28] . Moreover, the EPD by cracking processes is also generally very high since the energy conversion into surface energy is most efficient [29] . 3.3.1 PREP-3h The processing maps at other strains are essentially similar, so Fig. 6 (a) shows the map obtained at the strain of 0.6. High EPD occurs in the deformation temperature range from 1000 °C to 1020 °C and the strain rate range from 0.1 s −1 to 1 s −1 . In this domain, the efficiency steeply increases with the increase of strain rate and decrease of temperature. Such a feature has been observed to be typical of a fracture process in several materials [30,31] . In the deformation temperature range from 1030 °C to 1080 °C, the EPD is sensitive to strain rate and respectively reaches the minimum at the strain rates of 0.001 s −1 and 1 s −1 . This indicates that the occurrence of DRX is insufficient in these domains. Fig. 7 (a) shows the microstructure of the specimen deformed at the temperature of 1050 °C and the strain rate of 0.1 s −1 at the strain of 0.7. It can be seen that the PPBs are annihilated, and the fine recrystallized grains generated at high-angle grain boundaries. However, many micro dark carbide particles centralized at GBs may cause instability. Fig. 6 (a) shows that an instability domain lies in the deformation temperature range from 1020 °C to 1090 °C and the strain rate range from 1 × 10 −0.75 s −1 to 1 s −1 at the strain of 0.6. With the increase of strain, the instability domain gradually increases. Combined with Figs. 2 and 3 , because the specimen PREP-3h is more sensitive to strain rate and temperature, the strain rate should be strictly controlled during hot deformation when temperature is higher than 1030 °C. 3.3.2 PREP-5h The processing map of specimen PREP-5h at the strain of 0.6 is just shown in Fig. 6 (b) owing to less effect of strain on power dissipation. It can be seen that the EPD is higher at the temperature of 1000 °C, but it gradually decreases with the increasing temperature. At the temperature of 1030 °C, the EPD decreases to 33%. This indicates that the interaction of dislocations may stimulate the occurrence of DRX and result in the high efficiency of DRX at lower temperatures. With the increase of temperature, the decreasing volume fraction of strengthening phase γ ’ leads dislocation density to decrease. Thus, it can be seen that power dissipation is significantly affected by deformation temperature at lower temperatures (<1030 °C). Moreover, the map exhibits a triangle plateau in the temperature range from 1030 °C to 1060 °C and the strain rate range from 0.01 s −1 to 0.1 s −1 . According to the relation between EDP and DRX, the dislocation density is approximately a constant in this domain. However, at temperatures beyond about 1060 °C, Fig. 6 (b) shows that the EPD is more sensitive to strain rate and the variation of EDP with strain rate is obvious. At the deformation temperature of 1100 °C, there are two lower power dissipation domains in the strain rate ranges from 1 × 10 −0.5 s −1 to 1 s −1 and from 1 × 10 −3 s −1 to 1 × 10 −2.5 s −1 . The instability domain of specimen PREP-5h mainly lies in the strain rate range from 0.1 s −1 to 1 s −1 where EPD is lower and sensitive to strain rate. Correspondingly, Fig. 4 (b) shows that specimen PREP-5h exhibits higher flow stress in the steady-state stage. Thus, this instability may be associated with flow localization and cracking caused by low efficiency of DRX and DRV. Moreover, Fig. 7 (b) shows the microstructure of the specimen deformed at the deformation temperature of 1050 °C and the strain rate of 0.01 s −1 at the strain of 0.7. The recrystallized grain boundaries become serrated as a result of subgrain boundary formation. According to above analysis, the optimum condition of specimen PREP-5h is in the temperature range from 1020 °C to 1060 °C and the strain rate range from 0.01 s −1 to 0.1 s −1 . 3.3.3 AA-3h The processing maps of specimen AA-3h deformed at the strains of 0.4 and 0.6 are shown in Fig. 8 (a) and (b) . It can be see that the effect of strain on EPD is significant. At the strain of 0.4, the high power dissipation domains are as follows: (a) deformation temperature range from 1000 °C to 1050 °C and strain rate range from 1 × 10 −2.5 s −1 to 1 × 10 −0.75 s −1 ; (b) deformation temperature range from 1080 °C to 1100 °C and strain rate range from 0.001 s −1 to 0.01 s −1 ; (c) deformation temperature range from 1080 °C to 1100 °C and strain rate range from 0.1 s −1 to 1 s −1 . The peak of EPD increases with the increase of strain and reaches to 50% which indicates the superplastic may occur. Fig. 9 (a) shows the microstructure of the specimen deformed at the deformation temperature of 1050 °C and the strain rate of 0.1 s −1 . It can be seen that the recrystallized grains are uniform. The black precipitates are uniformly dispersed in grains, which may pin GBs and dislocations to keep the size of recrystallized grains small. The variation of power dissipation with strain rate is steeply in the domains as follows: (a) deformation temperature range from 1010 °C to 1060 °C and strain rate range from 1 × 10 −0.5 s −1 to 1 s −1 ; (b) deformation temperature range from 1080 °C to 1100 °C and strain rate range from 1 × 10 −2.5 s −1 to 1 × 10 −1.5 s −1 . In these domains, the unstable microstructure may be generated. Compared Fig. 8 (a) with (b), it can be seen that the instability domain increases with the increasing of strain. This indicates that the effect of strain on instability domains is significant. Thus, the optimum condition for specimen AA-3h is in the deformation temperature range from 1020 °C to 1060 °C and the strain rate range from 1 × 10 −2.5 s −1 to 1 × 10 −1 s −1 . 3.3.4 AA-5h The variation of EPD with strain rates and deformation temperatures at the strain of 0.3 is shown in Fig. 8 (c). The processing maps obtained at other strains are essentially similar. It can be seen that the map exists two high power dissipation domains at lower temperatures (<1030 °C), where the value of EPD is close to 59%. However, the EPD gradually decreases with the increase of temperature. It can be seen that the EPD decreases to 28% at the temperature of 1100 °C and the strain rates of 0.001 s −1 and 1 s −1 . Thus, the microstructure evolution is difficult to occur at higher temperatures (>1070 °C) except in the strain rate range from 1 × 10 −2.25 s −1 to 1 × 10 −0.5 s −1 , where the EPD increases to 38%. Fig. 9 (b) shows the microstructure of the specimen deformed at the deformation temperature of 1100 °C and the strain rate of 0.01 s −1 . The recrystallized grains coarsened (average grain size is about 42.5 μm) and the grain boundaries became serrated. Moreover, Fig. 8 (c) shows that two instability domains simultaneously existing at the strain of 0.3 are as follows: (a) deformation temperature range from 1000 °C to 1028 °C and strain rate range from 1 × 10 −2.6 s −1 to 1 × 10 −1.7 s −1 ; (b) deformation temperature range from 1078 °C to 1100 °C and strain rate range from 1 × 10 −0.5 s −1 to 1 s −1 . On the basis of above analysis, the processing windows for the specimens prepared by different preparation technologies are shown in Table 7 . 3.4 Effects of activation energy With a view to identify the rate controlling process during deformation, activation analysis is conducted using the kinetic rate equation, which relates the flow stress ( σ ) to temperature ( T ) and strain rate ( ε ˙ ) by: (5) ε ˙ = A σ n exp − Q R T where A is a frequency factor, n is the stress exponent, Q is the activation energy, R is the gas constant. On the basis of the stress exponent and activation energy values, the deformation mechanisms for microstructure development during hot working are identified. For this model, the flow stress data at various temperatures and strain rates at strains of 0.1 and 0.6 have been used. For specimen PREP-3h, the variation of flow stress with strain rate at different temperatures is shown in Fig. 10 (a) on a log–log scale. Linear statistical regression methods are used to determine the value of n . A liner fit with a good correlation factor could be obtained at the temperature of 1100 °C while deviations occurred at other temperatures especially at the strain rate of 0.01 s −1 . According to Eq. (7) , the value of n is obtained by inversing the slope of the line. It can be seen that the value of n is strain rate and temperature dependent. In this paper, this dependence is neglected by averaging the values of n . The average values of n for the specimens prepared by different preparation technologies are shown in Table 8 . Arrhenius plot showing the variation of flow stress with inverse of temperature on a semi-log scale is given in Fig. 10 (b). The activation energy values were obtained from the figure and are reported in Table 8 . It is well accepted that the climb of dislocations, DRV and DRX are related with activation energy, which can reflect the deformation behavior of materials [32] . Generally, the materials exhibiting high activation energy during deformation present bad workability owing to high deformation resistance caused by the increase of dislocation density. However, the occurrence of DRX can decrease the dislocation density and lead to lower flow stress. Table 8 shows that the activation energies at the strain of 0.1 are obviously higher than those at the strain of 0.6, indicating that the effect of strain on activation energy is significant. The activation energy of specimen PREP-3h is 266.66 kJ/mol at the strain of 0.1, which is close to that for lattice diffusion in pure nickel (278.2 kJ/mol) [33] . Thus, most of the dislocations may be hindered by grains and lead to the increase of activation energy in a sense. Although the activation energy decreases with the increase of strain, the value is still higher than the activation energy for grain boundary diffusion (118.2 kJ/mol) at the strain of 0.6 [34] . So, the dislocation movements mainly occur at GBs when the flow stress reaches a steady-state value, which is a typical DRX process. Further, it is well known that recrystallized grains nucleating and growing are a process of energy dissipation. This means that DRX is difficult to occur if activation energy is higher during deformation. High activation energy also makes the mobility of dislocation decrease and result in instability phenomenon [35] . Table 8 shows that specimen PREP-3h exhibits higher activation energy at the strain of 0.6. By comparing, it can be seen that the activation energy of specimen PREP-5h is obviously lower. This indicates that to prolongate the heat treatment time can decrease the dislocation density and cause lower activation energy which is good for workability. Moreover, it can be seen that the specimens prepared by AA also exhibit lower activation energy. This may be associated with non-uniform PPBs after HIP ( Fig. 3 (b)). The dislocations may easily shift across these PPBs owing to insufficient lattice energy to block the slip of dislocations. 3.5 DRX analysis An obvious Z effect on the microstructure of deformed nickel-base superalloy has been demonstrated: finer grains are formed in deformation with higher Z values, which can be understood from the grain refinement mechanism induced by plastic deformation [36] . It is known that the relation between dynamic recrystallized grain size and Zener–Hollomon parameter is described by: (6) D DRX = K Z A − n where D DRX is the size of recrystallized grains, Z is Zener–Hollomon parameter, K , n and A are constants related to material. According to Eq. (6) , the sizes of the recrystallized grains are in inverse proportion to the parameter Z . Further, it is known that an increase in strain rate during deformation is thought to have an equivalent effect to that of a decrease in deformation temperature and vice versa. The combined effects of strain rate and deformation temperature are often represented by Zener–Hollomon parameter, as defined by: (7) Z = ε ˙ exp Q R T where R is the gas constant, Q is the related activation energy for deformation (J/mol), ε ˙ is the strain rate (s −1 ), T is the deformation temperature (K). For specimen PREP-3h, the curves of Zener–Hollomon parameter vs. the flow stress is shown in Fig. 11 . This plot shows a good linear correlation between the flow stress and the Z value. The regression coefficient R 2 was 0.91. It implies that within the experimental condition, the power law relationship for hot deformation is obeyed. By means of linear statistical regression methods, the average values of parameter Z for the specimens prepared by different preparation technologies are obtained, as shown in Table 9 . It can be seen specimen PREP-3h exhibits higher Z value in steady-state stage, which indicates that the finer recrystallized grains may be generated during deformation. The corresponding microscopy shows that fine recrystallized grains exit at high-angle GBs during deformation ( Fig. 7 (a)). But according to previous analysis, the microstructure of specimen PREP-3h may exhibit instability during deformation owing to the high activation energy and more sensitivity to temperature and strain rate. In this paper, heat treatment time was increased to 5 h in order to decrease the activation energy of the specimens. However, the results show that Z value also decreased with the increase of heat treatment time although the dislocation density and the quantity of precipitation phase decreased, and this may cause grain coarsening during deformation ( Fig. 7 (b)). Thus, the weight may be balanced between workability and strength request. Moreover, by comparing, the specimens prepared by AA powder exhibit a better workability during hot working. But, to prolongate heat treatment time made their recrystallized grains coarsening severely ( Fig. 9 (b)). Based on the above-mentioned analysis, the heat treatment time of the specimens prepared by AA powder should be appropriately shortened. 4 Conclusion 1. The relationships among peak stress, deformation temperature and strain rate indicate that preparation technology significantly influences the hot deformation behavior of the specimens at higher strain rates (>0.1 s −1 ) or lower temperatures (<1050 °C). The P/M nickel-base superalloy prepared by PREP powder exhibit higher flow stresses at lower temperatures (<1050 °C), and the peak stress of specimen PREP-3h significantly increases with increasing strain rate and decreasing temperature. This may be related to its higher activation energy, uniform PPBs and the higher dislocation density. However, the effects of preparation technologies gradually decrease with the decrease of strain rate or the increase of temperature. 2. Compared with specimen AA-3h, the effect of strain on processing maps is slight for specimens PREP-3h and PREP-5h. At higher strain rates (>0.1 s −1 ), the instability may occur in domains where EPD is lower than 30%. The instability domains at higher strain rates (>0.1 s −1 ) may be related to adiabatic heating which can cause the flow localization and cracking, and those at higher temperatures (>1080 °C) and lower strain rates (<0.01 s −1 ) may be associated with the wedge cracking. Moreover, the micrograph of specimen PREP-3h deformed at the deformation temperature of 1050 °C and the strain rate of 0.1 s −1 shows that most of the precipitation phases concentrate at GBs, which easily lead to instability. 3. Generally, activation energy is close related to dislocation density and the interaction of dislocations may cause increasing deformation resistance. Since the GBs are less hard than the matrix and strengthening phase γ ’ , the stress concentrations may be relieved if the softening rate of grain boundary exceeds the stress accumulation rate at the GBs. Otherwise cracks or pores are expected to nucleate at the GBs. Thus, it can be seen that high activation energy is not good for hot working. By comparing, specimen AA-3h exhibits lower activation energy than specimen PREP-3h during deformation. This may be associated with non-uniform PPBs after HIP. Moreover, the effect of heat treatment time on activation energy is significant, so activation energy can be obviously decreased by prolongating heat treatment time. 4. It is accepted that the sizes of recrystallized grains are related to Zener–Hollomon parameter. Owing to higher Z value, specimen PREP-3h may generate finer recrystallized grains during deformation. The micrograph of specimen PREP-5h shows that the GBs are serration and the recrystallized grains are coarsening. Thus, the weight must be considered between heat treatment time and strength request. Furthermore, both of specimens AA-3h and AA-5h exhibit lower Z values. In order to prevent recrystallized grains coarsening, the heat treatment time of the specimens prepared by AA powder should be appropriately shortened. References [1] Y.H. Liu F.G. Li S.C. Wu Acta Aeronaut. Astronaut. Sin. Chinese 24 2003 278 [2] M.J. Zhang F.G. Li S.Y. Wang C.Y. Liu Mater. Sci. Eng. A 527 2010 6771 6779 [3] Y.Q. Ning Z.K. Yao H.Z. Guo H. Li T. Yu Y.W. Zhang Mater. Sci. Eng. A 527 2010 961 966 [4] G.J. Reddy N. Srinivasan A.A. Gokhale B.P. Kashyap J. Mater. Process. Technol. 209 2009 5964 [5] H.Y. Kim H.C. Kwon H.W. Lee Y.T. Im S.M. Byon H.D. Park J. Mater. Process. Technol. 205 2008 70 [6] Z. Gronostajski J. Mater. Process. Technol 125–126 2002 119 [7] S. Venugopal P.V. Sivaprasad M. Vasudevan S.L. Mannan S.K. Jha P. Pandey Y.V.R.K. Prasad J. Mater. Process. Technol. 59 1995 343 [8] V.G. Krishna Y.V.R.K. Prasad N.C. Birla G.S. Rao J. Mater. Process. Technol. 71 1997 377 [9] Y. Liu R. Hu J.S. Li H.C. Kou H.W. Li H. Chang H.Z. Fu J. Mater. Process. Technol. 209 2009 4020 [10] D.Y. Cai L.Y. Xiong W.C. Liu G.D. Sun M. Yao Mater. Charact. 58 2007 941 [11] K.L. Wang S.Q. Lu M.W. Fu X. Li X.J. Dong Mater. Charact. 60 2009 492 [12] Y.V.R.K. Prasad H.L. Gegel S.M. Doraivelu J.C. Malas J.T. Morgan K.A. Lark D.R. Barker Metall. Trans. A 15 1984 1883 [13] H.L. Gegel J.C. Malas S.M. Doraivelu V.A. Shende Forming and Forging vol. 14 1988 ASM International Ohio [14] M.C. Somani K. Muraleedharan Y.V.R.K. Prasad V. Singh Mater. Sci. Eng. A 245 1998 88 [15] M.C. Somani N.C. Birla Y.V.R.K. Prasad V. Singh J. Mater. Process. Technol. 52 1995 225 [16] Y.V.R.K. Prasad T. Seshacharyulu Mater. Sci. Eng. A 243 1998 82 [17] J. Luo M.Q. Li W.X. Yu H. Li Mater. Sci. Eng. A 504 2009 90 [18] Y.V.R.K. Prasad S. Sasidhara Hot Working Guide: A Compendium of Processing Maps 1997 ASM International Materials Park, OH Mater. Inform. Soc. [19] S.V.S.N. Murty B.N. Rao B.P. Kashyap J. Mater. Process. Technol. 166 2005 268 [20] H.J. Frost M.F. Ashby The Plasticity and Creep of Metals and Ceramics 1982 Pergamon Press London [21] R. Raj Metall. Trans. A 12 1981 1089 [22] J.T. Liu G.Q. Liu B.F. Hu Y.P. Song Z.R. Qing S. Xiang Y.W. Zhang Rare Met. Mater. Eng. 35 2006 418 422 [23] N. Srinivasan Y.V.R.K. Prasad Metall. Trans 25A 1994 2275 2284 [24] S.C. Medeiros Y.V.R.K. Prasad W.G. Frazier R. Srinivasan Mater. Sci. Eng. A 293 2000 198 207 [25] G. Ganesan K. Raghukandan R. Karthikeyan B.C. Pai Mater. Sci. Eng. A 369 2004 230 235 [26] O. Sivakesavam I.S. Rao Y.V.R.K. Prasad Mater. Sci. Technol. 9 1993 805 810 [27] N. Srinivasan Y.V.R.K. Prasad Mater. Sci. Technol. 8 1992 206 212 [28] N. Ravichandran Y.V.R.K. Prasad Mater. Sci. Eng. A 156 1992 195 204 [29] Y. Ning Z. Yao M.W. Fu H. Guo Mater. Sci. Eng. A 527 2010 6968 6974 [30] Y.V.R.K. Prasad S. Sasidhara Hot Working Guide: A Compendium of Processing Maps 1997 ASM International Materials Park, OH [31] T. Seshacharyulu S.C. Medeiros W.G. Frazier Y.V.R.K. Prasad Mater. Sci. Eng. A 325 2002 112 125 [32] H.J. McQueen Mater. Sci. Eng. A 387–389 2004 203 208 [33] E.A. Brandes Smithells Metals Reference Book 6th ed. 1983 Butterworths London [34] I. Kaur, W. Gust, L. Kozma, Handbook of Grain and Interphase Boundary Diffusion Data, vol. 1, pp. 523 and vol. 2, pp. 1066, Ziegler Press, Stuttgart, 1989. [35] J.C. Malas S. Venugopal T. Seshacharyulu Mater. Sci. Eng. A 368 2004 41 47 [36] Y.Q. Ning Z.K. Yao H.Z. Guo M.W. Fu H. Li X.H. Xie Mater. Sci. Eng. A 527 2010 6794 6799