Hot deformation behavior and processing workability of a Ni-based alloy

Hot deformation behavior and processing workability of a Ni-based alloy

Accepted Manuscript Hot deformation behavior and processing workability of a Ni-based alloy Zhipeng Wan, Lianxi Hu, Yu Sun, Tao Wang, Zhao Li PII: S0...

5MB Sizes 0 Downloads 49 Views

Accepted Manuscript Hot deformation behavior and processing workability of a Ni-based alloy Zhipeng Wan, Lianxi Hu, Yu Sun, Tao Wang, Zhao Li PII:

S0925-8388(18)32884-6

DOI:

10.1016/j.jallcom.2018.08.010

Reference:

JALCOM 47098

To appear in:

Journal of Alloys and Compounds

Received Date: 16 April 2018 Revised Date:

1 August 2018

Accepted Date: 2 August 2018

Please cite this article as: Z. Wan, L. Hu, Y. Sun, T. Wang, Z. Li, Hot deformation behavior and processing workability of a Ni-based alloy, Journal of Alloys and Compounds (2018), doi: 10.1016/ j.jallcom.2018.08.010. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Graphical Abstract

ACCEPTED MANUSCRIPT Hot deformation behavior and processing workability of a Ni-based alloy Zhipeng Wan1, 2, 3, Lianxi Hu1, 2*, Yu Sun1, 2**, Tao Wang3, Zhao Li3 1 National Key Laboratory for Precision Hot Processing of Metals, Harbin Institute of Technology, Harbin 150001, P.R. China

Harbin 150001, China

RI PT

2 School of Materials Science and Engineering, Harbin Institute of Technology,

3 Science and Technology on Advanced High Temperature Structural Materials

China

M AN U

ABSTRACT

SC

Laboratory, AEEC Beijing Institute of Aeronautical Materials, Beijing 100095, P.R.

In the present work, the high-temperature deformation behavior of a U720LI alloy was investigated by means of hot compression tests at temperatures of 1060-1180oC, strain rates of 0.001-10s-1 under true strain of 0.8. The Arrhenius-type model, considering the dissolution of γ' precipitates, was established on the basis of the friction

TE D

and adiabatic heating corrected flow stress data. The hot deformation activation energies for quasi-γ phase and γ+γ' dual-phase microstructures were determined as 417 kJ·mol-1 and 687 kJ·mol-1, respectively. The high activation energy for γ+γ' dual-phase

EP

microstructures was mainly attributed to the precipitation hardening effect of γ' (Ni3(Al,Ti)) particles which hindered dislocation slip via the pinning effect during hot

AC C

deformation. The processing map was developed based on the dynamic material model. According to the detailed microstructural investigations, the optimum hot deformation domains of U720LI alloy were identified as 1080-1100oC/0.01-0.3s-1 for γ+γ' dual-phase microstructure and 1140-1160oC/0.01-0.1s−1 for quasi-γ phase microstructure, respectively. * Corresponding author. ** Corresponding author. Tel.: +86 451 86418613. Fax: +86 451 86418613. E-mail addresses: [email protected] (L.X. Hu), [email protected] (Y. Sun). 1

ACCEPTED MANUSCRIPT Keywords: U720LI alloy; Hot deformation behavior; Processing map; Workability.

1. Introduction U720LI alloy is a typical cast-wrought nickel-based alloy used for turbine discs and

RI PT

blades up to 730oC [1]. Compared with the Udimet720 alloy, U720LI alloy was developed by decreasing the chemistry of interstitials, i.e. C and B in Udimet720 alloy, and then the term ‘‘low interstitial” abbreviated as ‘‘LI”. Chromium content has also been reduced from 18% to 16% to prevent the formation of sigma phase to long-term

SC

high-temperature exposure [2]. U720LI alloy possesses excellent mechanical properties at high temperature [3, 4] and is quite stable in the long-term (as long as

M AN U

1000h) exposed at 760oC [5]. Generally, the mechanical properties of materials rely on the microstructure which, in turn, are controlled by hot deformation parameters such as temperature, strain and strain rate. Thus, the Arrhenius-type equation and processing map play important roles in characterization of flow stress behavior, optimization of process parameters and microstructure control for metals and alloys during hot

TE D

deformation. Tan et al. [6] have studied the hot deformation behavior of fine grained Inconel 718 superalloy. The flow stress values predicted by the developed Arrhenius-type constitutive equation were in reasonable agreement with the

EP

experimental values, and optimum process parameters were determined based on the established processing maps. Zhang et al. [7] have investigated the hot working

AC C

characteristics of Ni-Cr-Mo-based C276 superalloy based on the flow stress curves and microstructure. Two peaks of efficiency corresponding to the occurrence of DRX were determined to be deformed at 1125oC and 0.1s-1 as well as 1025oC and 0.001s-1. Wu et al. [8] have developed the Arrhenius-type equation and processing map in the temperature range of 1050-1150oC and strain rates of 0.001-1 s-1 for hot isostatic pressed spray formed FGH100 alloy. Pu et al. [9] have studied the flow stress behavior and hot workability of nickel-based alloy UNS10276. The flow instability behavior, induced by microbands, deformation twins and restrained dynamic recrystallization (DRX) behavior, was identified at temperatures lower than 1180oC and strain rates 2

ACCEPTED MANUSCRIPT higher than 5s-1. Moreover, the roles of glide or climbing of dislocations, microtwins, stacking faults and atomic diffusion in various hot deformation conditions have been revealed [10-13]. Over the past few years, lots of efforts have been made on the U720LI alloy. Chang

RI PT

et al. [14] have studied the solidification behavior of U720LI superalloy. It demonstrated that the formation of η-phase, which was harmful to the mechanical property, was mainly attributed to the limited solubility of Ti in primary γ phase and γ/γ' Eutectic. However, increasing boron content can significantly retard the γ matrix

SC

solidification, which would increase the eutectic (γ+γ') precipitation and reduce the formation of η-phase phase and Zr-rich particles [15]. The interactions of slip bands

M AN U

with grain boundaries for U720LI alloy have been investigated by SEM and EBSD techniques during tensile tests [16]. The twist angle between projections of slip bands and the boundary plane was identified to govern slip transfer processes. Németh et al. [17] have identified a relationship between tensile ductility and oxidation kinetics for

TE D

U720LI superalloy. The results suggested that the internal intergranular oxidation along the γ-grain boundaries, and in particular, at incoherent interfaces of the primary γ' precipitates and γ matrix may cause the embrittlement of the alloy. Furthermore, the flow stress behavior and microstructural changes during hot deformation of U720LI

EP

alloy have been proposed in some previous studies. Liu et al. [18] have studied the hot deformation behavior of U720LI alloy with fine, coarse and mixed grains,

AC C

respectively. Dynamic softening mechanism and the effect of γ' precipitates on microstructure evolution were identified. For the coarse grain, the nucleation of DRX grains was influenced by the γ′ interparticle spacing [19]. Qu et al. [20] have established the constitutive equation and the processing map of GH4720Li alloy based on the flow stress during hot compression tests. However, the effect of phase transformation during hot compression tests was not considered and the validity of the optimum process parameters were not testified by the microstructure. In this study, the hot deformation behavior of U720LI nickel-based alloy is investigated according to 3

ACCEPTED MANUSCRIPT hot compression tests. The Arrhenius-type model considering the dissolution of γ' precipitates is established using the corrected stress-strain data and the optimum process parameters for hot deformation are determined with the validation of

2. Experimental materials and procedures

RI PT

deformed microstructure in both quasi-γ phase and γ+γ' dual-phase regions.

The main chemical composition (wt %) of U720LI alloy used in this research is as follows: C 0.018, Cr 15.94, Co 14.5, W 1.25, Mo 2.94, Al 2.56, Ti 4.91, Zr 0.03, and

SC

balance Ni. The microstructure of the initial material from an as-forged bar is shown in Fig. 1 and its average grain size is 21µm. The specimen with a diameter of 10 mm

M AN U

and a height of 15 mm was cut by electric-discharge machining and were deformed using a Gleeble 3800 thermal simulation machine. High quality graphite sheet with thickness of 0.05 mm was utilized between anvils and specimen surface to reduce the friction effect during hot deformation. As shown in Fig. 2, the specimens were heated up to deformation temperatures of 1060oC, 1080oC, 1100oC, 1120oC, 1140oC, 1160oC

TE D

and 1180oC at a rate of 10oC/s and held for 5min to minimize temperature gradient. Compression tests were conducted to a true strain of 0.8 at strain rates ( ε& ) of 0.001s-1, 0.01s-1, 0.1s-1, 1s-1 and 10s-1, and then immediately quenched in water to freeze the

EP

microstructure. All of the compressed specimens were cut parallel to the compression axis by wire-electric discharge machine. After mechanical polishing, the specimens

AC C

were chemical-etched using 5g CuCl2+100ml C2H5OH+100ml HCl, and optical microscope (OM) was employed for microstructure observations.

Fig. 1. Initial microstructure of as-forged U720LI alloy. 4

RI PT

ACCEPTED MANUSCRIPT

Fig. 2. Schematic diagram of hot compression test in this work.

SC

3. Results and discussion

M AN U

3.1. Corrections of friction and adiabatic heating on flow stress curves 3.1.1. Correction of the friction effect

During uniaxial hot compression tests, the presence of friction can lead to the formation of dead zones and inhomogeneous deformation, and flow stress deviates from their true values. Compared with the sliding friction, the effect of sticking one at

TE D

higher temperature would increase the flow stress significantly. Therefore, in order to evaluate the effect of friction and determine the accurate flow stress during hot deformation, researchers employed the simplified relations to eliminate the effect of

EP

friction on flow stress curves at high deformation temperatures. The corrected flow stress considering the friction effect can be written as [21, 22]:

AC C

σf =

(

)

σ

1 + 2 3 3 f ( r0 h0 ) exp ( 3ε 2 )

(1)

where σf is the corrected flow stress, σ is the measured flow stress, ε is the measured strain and f is the friction factor. r0 and h0 are the initial radius and height of samples, respectively. It can be noted from the Eq. 1 that the friction factor (f) plays an important role in determining the value of corrected flow stress. The friction factor f is always taken as a constant [22, 23], while the experimental evidence demonstrates that the friction factor f is related to the deformation temperature and strain rate [21]. Based on the theory proposed by Ebrahimi and Najafizadeh, the friction factor can be 5

ACCEPTED MANUSCRIPT determined by [21, 24]: f =

(r h)b

(2)

( 4 3 ) − ( 2b 3 3 )

where b is the barrel parameter, h and r are height and average radius of cylinder after

b=4

∆r h r ∆h

r = r0

h0 h

(3)

SC

as:

RI PT

deformation. The barrel parameter (b) and average radius(r) in Eq. 2 can be expressed

(4)

M AN U

where ∆h and ∆r are reduction of height and difference between maximum and top radius of cylinder after deformation. According to Eqs. 2-4, the friction factor can be determined by measuring the height, maximum radius and top radius of cylinder after deformation. However, it is difficult to measure the top radius of cylinder after deformation accurately. Therefore, with approximation of the profile of the barreled

TE D

samples with an arc of a circle, top radius of cylinder after deformation can be calculated by following equation[21]:

h0 2 r0 − 2rM2 h

(5)

EP

rT = 3

where rM and rT are maximum and top radius of cylinder after deformation, as shown

AC C

in Fig. 3. Therefore, the friction factor (f) can be calculated by measuring the maximum radius (rM) as well as the height of cylinder after deformation (h) according to the Eqs. 2-5. The predicted frictionless flow stress curves derive from the Eq. 1 are shown in Fig. 4a. It can be seen from the figure that the corrected flow stress is lower than measured flow stress. The difference between corrected and uncorrected flow stress curves increases with increasing strain, which is mainly attributed to the increased contact surface between the specimen and die. In additions, compared with the uncorrected flow stress curves, the work hardening behavior is eliminated after correction of the friction effect. 6

RI PT

ACCEPTED MANUSCRIPT

M AN U

SC

Fig. 3. Schematic diagram of parameters in Eqs. 1-5.

Fig. 4. Corrected flow stress curves due to the friction and adiabatic heating effects of

TE D

U720LI alloy under various temperatures and strain rates: (a) 1100 oC and (b) 10s-1. 3.1.2. Correction of the adiabatic heating effect The decrease of flow stress after reaching a peak stress is a familiar appearance for

EP

various alloys attributing to the flow softening effect. The softening effect can be arisen from microstructural change and increase of temperature in alloys during hot

AC C

deformation, both or either alone. Microstructural changes mainly indicate the occurrence of continuous dynamic recrystallization (CDRX) and discontinuous dynamic recrystallization (DDRX) behavior in nickel-based alloy. The increase of temperature in alloys during hot deformation is related to the stain rate. At high strain rate, the hot compression tests are considered as the adiabatic condition, and temperature of the samples increases significantly. While temperature variation caused by deformation is negligible at low strain rate (isothermal condition). Therefore, to remove the effect of temperature rise caused by adiabatic heating, an expression for temperature correction should be performed on the flow stress curves. The flow stress 7

ACCEPTED MANUSCRIPT ( σ ) corresponding to the actual temperature T can be expressed as follows [22, 23, 25, 26]:

σ =σ f +

0.95 ρC p

( ∫ σ dε ) B  ddσT  ε

f

(6)

f

0

ε& ,ε

RI PT

where ρ and Cp are the density and specific heat of deformed material, respectively. T is the absolute temperature (K) during hot deformation. The adiabatic correction factor (B) in Eq. 6 can be calculated using the following equation:

−1

(7)

SC

  ε B = 1 +   ( ( xw K w ) + (1 HTC ) + ( xD K D ) ) xw ρ C pε&   

where xw is half height of the deformed sample, Kw is the thermal conductivity of

M AN U

deformed material, HTC is the interface heat-transfer coefficient, xD is the distance from the die surface to the die interior where temperature is constant and KD is the thermal conductivity of die. The parameters in Eqs. 6 and 7 are shown in Table 1. In this study, the effect of adiabatic heating is ignored because of the negligible

TE D

temperature rise effect for the strain rates lower than 1s-1, and the similarity conditions are performed on 904L austenitic stainless steel and powder metallurgy Ti-47Al-2Cr-2Nb-0.2W alloy [24, 27]. The corrected flow stress curves due to the adiabatic heating and friction effects are shown in Fig. 4b. These curves exhibit a peak

EP

in the flow stress at a strain and reach a steady state, which indicates that the corrected flow stress curve displays the characteristic of dynamic recrystallization (DRX).

AC C

Table 1 Parameters in Eq. 3 and 4[25, 28, 29].

Parameter ρ Cp Kw KD HTC xD

Value 8140 (kg·m-3) 750 (J·kg-1·K-1) 29 (W·mK-1) 21 (W·mK-1) 25000 (W·m-2·K-1) 0.015 (m)

3.2. Arrhenius-type constitutive equation Generally, the relationship among the flow stress, deformation temperature and 8

ACCEPTED MANUSCRIPT strain rate for a give strain can be described by the Arrhenius-type equation. The Zener-Holloman parameter (Z) which represents the effects of strain rate and temperature on flow stress can be expressed as [30]:

 A sinh ασ n ( )     Q n Z = ε& exp( ) = f (σ ) =  A1σ 1 RT  A exp βσ ( )  2 

RI PT

(8)

where Q is the activation energy (kJ·mol-1), R is the universal gas constant (8.314

SC

kJ·mol-1·K-1). As shown in Eq. 8 that correlationship between Z and the flow stress can be determined by exponential, power and hyperbolic type equations. A, A1, A2, n1,

M AN U

α and β are material constants and n is stress exponent. The peak stress level is employed to depict the hot deformation behavior. Taking the natural logarithm of Eq. 8:

ln ( A ) + n ln sinh (ασ )  Q  1   ln ( ε& ) +   = ln ( A1 ) + n1 ln (σ ) R T   ln ( A2 ) + βσ

TE D

(9)

Material constants β, n1 and n can be determined by taking partial differential on

EP

both sides of Eq. 9 for a given deformation temperature, as expressed in the following equations [31, 32].

AC C

   ∂ ln ε&   ∂ ln ε&  ∂ ln ε&   β =  ; n1 =   ; n= σ σ ∂ ∂ ln    ασ ∂ ln sinh T T ( p )  p  p     

(10) T

The values of β and n1 can be determined from the Eq. 10 by employing the

corrected flow stress. Hence, the material constant α can be obtained (α=β/n1), and the stress exponent n is deduced by linear regression. Taking partial differential on both sides of Eq. 9 and rearranging yields:

  ∂ ln ε&  Q = R  ∂ ln sinh (ασ p )     T  9

 ∂ ln sinh (ασ p )      ∂ 1 T   ( )   ε&

(11)

ACCEPTED MANUSCRIPT In Eq. 11, the activation energy can be determined by the slope of ln [sinh(ασ)] versus 1/T. Based on the quantitative metallography, the majority amount of the γ' precipitates are dissolved at a temperature of 1140oC and a hold time of 5min, and the mass percent of the γ' precipitates is no more than 5% according to the phase

RI PT

diagram[18]. The precipitation hardening effect induced by γ' precipitates reduces significantly. Thus the samples deformed at the temperature of 1140oC can be approximately considered as in the single γ phase region. Then the deformation temperature over 1140oC is termed as quasi-γ phase region. The plots used to derive

SC

average values of n1, β, α, n, Q and InA are shown in Fig. 5. Based on the experimental data (shown in Fig. 5), the average values of n1, β, α, n, Q and InA for

M AN U

U720LI alloy during hot compression tests are shown in Table 2. The microstructure of γ+γ' dual-phase and quasi-γ phase is modelled using two linear ranges from which the activation energy is determined separately for the two phase regions, as shown in Fig. 5c. It can be seen from Table 2 that the value of stress exponent is 3.571, which is similar to other nickel-based alloys [33-35]. The activation energies Q for quasi-γ

TE D

phase and γ+γ' dual-phase microstructures of U720LI alloy are 417 kJ·mol-1 and 687 kJ·mol-1, respectively. They are higher than the self-diffusion energy of pure nickel alloy (270 kJ·mol-1) [36]. This suggests that the pinning effect of γ' precipitates and

EP

hardening effect induced by alloying element are able to enhance the deformation activation process significantly. However, the activation energies for U720LI alloy are

AC C

lower than that of Udimet720 alloy in the range of 470-743 kJ·mol-1 [37], which is mainly attributed to the high content of interstitials (carbon, C and boron, N) and alloying element (chromium, Cr) compared with the U720LI alloy [38]. Consequently, by substituting the parameters listed in Table 2, the constitutive equations at peak stress level can be expressed as:

ε& = 9.535 × 1024 sinh ( 0.00687σ p )  for γ+γ' dual-phase microstructure, and:

10

3.571

exp(−

687000 ) RT

(12)

ACCEPTED MANUSCRIPT ε& = 5.988 × 1014 sinh ( 0.00687σ

p

)

3.571

exp(−

417000 ) RT

(13)

for quasi-γ phase microstructure. Comparison between peak stresses experimentally determined and calculated by the established Arrhenius-type constitutive equation is shown in Fig. 6. It can be seen from the figure that the calculated peak stresses agree

model prediction can be identified at a strain rate of 10s-1.

RI PT

well with the experimental results at strain rates of 0.001-1s-1, while a low accuracy of

U720LI alloy. n1

β

4.544

0.0312

γ+γ' dual-phase

n

Q (kJ·mol-1)

lnA

0.00687

687

57.517

417

34.026

3.571

AC C

EP

TE D

Quasi-γ phase

α

M AN U

Phase region

SC

Table 2 Kinetic parameters for γ+γ' dual-phase and quasi-γ phase microstructures of

Fig. 5. Plots of ln ε& -ln σ p (a) and ln ε& -ln[sinh(α σ p )] (b) at different temperatures, (c) Plot of ln[sinh(α σ p )]-1000/T for quasi-γ phase and γ+γ' dual-phase microstructures and (d) relationship between lnZ and ln[sinh(α σ p )] of U720LI alloy. 11

SC

RI PT

ACCEPTED MANUSCRIPT

Fig. 6. Comparison between peak stresses experimentally determined and calculated by

M AN U

the established Arrhenius-type constitutive equation. 3.3. Establishment of processing map

To predict the optimum processing conditions of U720LI alloy during hot deformation, the processing map is presented in this section. A processing map is developed based on dynamic materials model (DMM) which is proposed by

TE D

Prasad[39]. In the DMM, the total power energy P can be divided into two main parts, one is the power dissipation (G) indicating the dissipation energy value arising from plastic deformation, and the other one is metallurgical processes (J) which is related to

EP

the energy concerning the microstructural changes, like dynamic recovery and dynamic recrystallization. The correlationship among P, G and J can be expressed

AC C

mathematically as [40, 41]:

σ

ε

0

0

P = G + J = ∫ ε& dσ + ∫ σ d ε&

(13)

The partitioning of power between J and G is given by the following equation:

& d ln σ ∆ log σ dJ ε& dσ εσ = = ≈ =m dG σ d ε& σ ε&d ln ε& ∆ log ε&

(14)

where m is the strain rate sensitivity. For a certain temperature, J can be simplified to: J max =

12

mσ ε& m +1

(15)

ACCEPTED MANUSCRIPT Eq. 15 indicates that the value of J at any given temperature and strain rate can be determined according to the flow stress ( σ ), strain rate ( ε& ) and strain rate sensitivity (m). A maximum value of J can be obtained when m equals to 1, and the process is considered as an ideal linear dissipator, thus:

σ ε&

(16)

RI PT

J max =

2

The efficiency of power dissipation (η), which indicates the power dissipation

dimensionless parameter [31]: 2m m +1

(17)

M AN U

η = J J max =

SC

capacity of the sample, for a non-linear dissipator can be defined in terms of a

Domains with high η in power dissipation map can be identified as the optimum processing conditions, which should be further supported by the microstructure of deformed samples. The flow instability characterized by adiabatic shear bands, slip localization and crack formation based on the grain boundary cavitation should also be

TE D

considered when the optimum processing conditions are determined [42]. Therefore, the flow instability map based on a continuum instability criterion is established. Further, the extremum principles of irreversible thermodynamics as applied to continuum mechanics of large plastic deformation are employed to identify the

AC C

EP

occurrence of flow instability. The instability criteria ξ ( ε& ) is expressed by [43]:

ξ ( ε& ) =

∂ ln  m ( m + 1)  +m≤0 ∂ ln ε&

(18)

It can be deduced from the Eq. 18 that the instability domains are characterized by

the negative value of ξ ( ε& ) . In this study, the corrected flow stress curves are used to determine m, η and ξ, and then the processing map is established according to the superimposition of power dissipation map and instability map at a strain of 0.8, as shown in Fig. 7. The processing map is a three dimensional plot with isocontour lines which is dependent on deformation temperature and strain rate at a given strain. It can 13

ACCEPTED MANUSCRIPT be seen from the Fig. 7 that the gray colored regions represent the unstable zones and the white colored domain denotes stable zones. The processing map in this study is divided into three main domains, and the workability of the sample in each domain is discussed with the help of the deformed microstructure. The domain C with the η of

RI PT

10-25% and negative ξ ( ε& ) exists in the temperature range of 1060-1100oC and strain rates of 1-10s-1 for a true strain of 0.8. In general, the flow instability of the samples during hot deformation could be probably associated with adiabatic shear

SC

bands, localized plastic flow or cracking[44]. It could be seen from the Fig. 7 that the main instability domain, identified from the microstructure in Domain C, is

M AN U

characterized by the adiabatic shear bands. It has been widely acknowledged that the adiabatic heating during hot deformation is occurred at higher strain rates due to insufficient deformation times [43]. Domain B indicates the unstable zone with high η (the maximum η is about 25%) deformed in the temperature range of 1140-1180oC and strain rates of 1-10s-1 for a true strain of 0.8. The microstructure in Domain B

TE D

corresponds to the instability mechanisms characterized by cracking along the compressed grain boundary, as shown in Fig. 7. At high strain rates and temperature, grain boundary length per unit area decreases with the increase of deformation

EP

temperature and the stress concentration at grain boundaries could not be released instantly due to insufficient time [43], and then the intergranular cracking initiates. In

AC C

addition, the microstructure shown in domain B of Fig. 7 exhibits the typical incomplete DRX behavior with necklace structure, which suggests that initiation of DRX nuclei distributes along the large size of deformed grain boundaries. The deformation condition with high power dissipation efficiency in the stable

region are termed as domain A, as seen from Fig. 7. Domain A is mainly composed of two stable regions with high η found to be nearly 31-43%, which corresponds to microstructure representing a homogeneous and completed dynamic recrystallization structures without deformation defect. Moreover, it could be seen from the processing map shown in Fig. 7, the high efficiency of power dissipation (about 49%) deformed 14

ACCEPTED MANUSCRIPT at 1180oC/0.01s-1 is presented in the stable zone, while the average grain size of the deformed sample exhibits extraordinary grain growth behavior depended on its high temperature and low strain rate, as shown in Fig. 8. Therefore, based on Eqs. 17 and 18 as well as validation of micrographs, optimum processing conditions for the alloy in

TE D

M AN U

SC

RI PT

this study are 1080-1100oC/0.01-0.3s-1 and 1140-1160oC/0.01-0.1s−1.

Fig. 7. Processing map of U720LI alloy at ε=0.8 with microstructure validation in

EP

different deformation regions.

The microstructure evolution of samples deformed at a strain of 0.8 and various

AC C

deformation conditions is shown in Fig. 8. It can be seen from the figure that the growth of dynamic recrystallized grain is promoted by the increase of deformation temperature and decrease of strain rate. Generally, the correlationship between Z and DRX grain size can be expressed by an exponential function[45]. However, the presence of precipitates corresponding to the chemical composition of the alloy and the adiabatic shear bands because of the inhomogeneous deformation would affect the deformed microstructure of the materials significantly. Thus, the correlationship between Z and DRX grain size (dDRX) is not always possible to be established or 15

ACCEPTED MANUSCRIPT presents poor correlation coefficient for some specific alloys [36, 46, 47]. The distribution of DRX grain size under various deformation conditions as well as the variation of Z with respect to DRX grain size are shown in Fig. 9. It can be seen from Fig. 9a that the average DRX grain size is not sensible to the deformation temperature

RI PT

at high strain rate due to the occurrence of complete DRX behavior and the restrained grain growth behavior. Hence, as shown in Fig. 9b, it is impossible to establish a

AC C

EP

TE D

M AN U

SC

relation between Z and DRX grain size in the present investigation.

Fig. 8. Microstructure of U720LI alloy at the strain of 0.8 with different hot compression conditions.

16

RI PT

ACCEPTED MANUSCRIPT

SC

Fig. 9. (a) Distribution of DRX grain size under various deformation conditions and (b) relationship between Z and DRX grain size of U720LI alloy.

M AN U

4. Conclusions

Hot deformation behavior and workability of U720LI alloy were investigated by using the hot compression tests on Gleeble 3800 thermal-mechanical simulator, and following conclusions can be drawn:

TE D

(1) The experimental flow stress for all deformation conditions was modified by the correction of friction effect, and the influence of adiabatic heating effect on the flow stress was also discussed for strain rates of 10s-1 and 1s-1. The dynamic recrystallization was identified as the dominant softening mechanism for U720LI

EP

alloy based on the corrected flow stress curves due to the friction and adiabatic heating effects.

AC C

(2) The Arrhenius-type model considering the dissolution of γ' precipitates was

established using corrected flow stress. The dependence of flow stress on strain rate and temperature during hot deformation can be expressed as: 3.571

exp(−

687000 ) RT

3.571

exp(−

417000 ) RT

ε& = 9.535 × 1024 sinh ( 0.00687σ ) 

for γ+γ' dual-phase microstructure,

ε& = 5.988 × 1014 sinh ( 0.00687σ )  for quasi-γ phase microstructure.

(3) Activation energies Q for quasi-γ phase and γ+γ' dual-phase microstructuress of 17

ACCEPTED MANUSCRIPT U720LI alloy were 417 kJ·mol-1 and 687 kJ·mol-1, respectively. The activation energy of U720LI alloy was higher than the self-diffusion energy of pure nickel alloy and lower than Udimet720 alloy in the range of 470-743 kJ·mol-1. (4) Based on the analysis of power dissipation efficiency and microstructure

RI PT

evolution, the optimum hot deformation processing conditions for γ+γ' dual-phase and quasi-γ phase microstructures were determined be 1080-1100oC/0.01-0.3s-1 and

AC C

EP

TE D

M AN U

SC

1140-1160oC/0.01-0.1s−1, respectively.

18

ACCEPTED MANUSCRIPT Acknowledgment This research did not receive any specific grant from funding agencies in the public,

AC C

EP

TE D

M AN U

SC

RI PT

commercial, or not-for-profit sectors.

19

ACCEPTED MANUSCRIPT Data availability The raw/processed data required to reproduce these findings can be shared, and data

AC C

EP

TE D

M AN U

SC

RI PT

will be available on request.

20

ACCEPTED MANUSCRIPT

References [1] K. Gopinath, A.K. Gogia, S.V. Kamat, et al., Dynamic strain ageing in Ni-base superalloy 720Li, Acta Mater. 57 (2009) 1243-1253. [2] M.P. Jackson, R.C. Reed, Heat treatment of Udimet 720Li: the effect of

RI PT

microstructure on properties, Mater. Sci. Eng. A 259 (1999) 85-97. [3] M. Marchionni, G.A. Osinkolu, G. Onofrio, High temperature low cycle fatigue behaviour of Udimet 720 Li superalloy, Int. J. Fatigue 24 (2002) 1261-1267.

[4] W.J. Evans, J.P. Jones, S. Williams, The interactions between fatigue, creep and

SC

environmental damage in Ti 6246 and Udimet 720Li, Int. J. Fatigue 27 (2005) 1473-1484.

M AN U

[5] P.W. Keefe, S.O. Mancuso, G.E. Maurer, Effect of heat treatment and chemistry on the long-term phase stability of a high strength nickel-based superalloy, Superalloys 1992 (1992) 487-496.

[6] Y.B. Tan, Y.H. Ma, F. Zhao, Hot deformation behavior and constitutive modeling of

TE D

fine grained Inconel 718 superalloy, J. Alloys Compd. 741 (2018) 85-96. [7] C. Zhang, L.W. Zhang, W.F. Shen, et al., The processing map and microstructure evolution of Ni-Cr-Mo-based C276 superalloy during hot compression, J. Alloys Compd. 728 (2017) 1269-1278.

EP

[8] H.X. Wu, C.C. Ge, Q.Z. Yan, et al., Plastic deformation behavior of spray formed superalloy FGH100, Mater. Sci. Eng. A 699 (2017) 156-164.

AC C

[9] E. Pu, W. Zheng, Z. Song, et al., Hot deformation characterization of nickel-based superalloy UNS10276 through processing map and microstructural studies, J. Alloys Compd. 694 (2017) 617-631. [10] Y.H. Liu, Y.Q. Ning, X.M. Yang, et al., Effect of temperature and strain rate on the workability of FGH4096 superalloy in hot deformation, Mater. Des. 95 (2016) 669-676. [11] Y.P. Li, X.J. Ye, J.X. Li, et al., Influence of cobalt addition on microstructure and hot workability of IN713C superalloy, Mater. Des. 122 (2017) 340-346. 21

ACCEPTED MANUSCRIPT [12] M. Azarbarmas, M. Aghaie-Khafri, J.M. Cabrera, et al., Dynamic recrystallization mechanisms and twining evolution during hot deformation of Inconel 718, Mater. Sci. Eng. A 678 (2016) 137-152. [13] M. Fahrmann, A. Suzuki, Effect of cooling rate on Gleeble hot ductility of Udimet

RI PT

alloy 720 billet, Superalloys 2008 (2008) 311-316. [14] L.T. Chang, H. Jin, W.R. Sun, Solidification behavior of Ni-base superalloy Udimet 720Li, J. Alloys Compd. 653 (2015) 266-270.

[15] G.D. Zhao, L.X. Yu, G.L. Yang, et al., The role of boron in modifying the

SC

solidification and microstructure of nickel-base alloy U720Li, J. Alloys Compd. 686 (2016) 194-203.

M AN U

[16] J. Genée, L. Signor, P. Villechaise, Slip transfer across grain/twin boundaries in polycrystalline Ni-based superalloys, Mater. Sci. Eng. A 701 (2017) 24-33. [17] A.A.N. Németh, D.J. Crudden, D.E.J. Armstrong, et al., Environmentally-assisted grain boundary attack as a mechanism of embrittlement in a nickel-based superalloy,

TE D

Acta Mater. 126 (2017) 361-371.

[18] F.F. Liu, J.Y. Chen, J.X. Dong, et al., The hot deformation behaviors of coarse, fine and mixed grain for Udimet 720Li superalloy, Mater. Sci. Eng. A 651 (2016) 102-115. [19] Q.Y. Yu, Z.H. Yao, J.X. Dong, Deformation and recrystallization behavior of a

EP

coarse-grain, nickel-base superalloy Udimet720Li ingot material, Mater. Charact. 107 (2015) 398-410.

AC C

[20] J.L. Qu, Z.N. Bi, J.H. Du, et al., Hot Deformation Behavior of Nickel-Based Superalloy GH4720Li, J. Iron Steel Res. Int. 18 (2011) 59-65. [21] R. Ebrahimi, A. Najafizadeh, A new method for evaluation of friction in bulk metal forming, J. Mater. Process. Tech. 152 (2004) 136-143. [22] A. Gholamzadeh, A. Karimi Taheri, The prediction of hot flow behavior of Al–6%Mg alloy, Mech. Res. Commun. 36 (2009) 252-259. [23] M.A. Mostafaei, M. Kazeminezhad, Hot deformation behavior of hot extruded Al–6Mg alloy, Mater. Sci. Eng. A 535 (2012) 216-221. 22

ACCEPTED MANUSCRIPT [24] Y. Han, G.J. Qiao, J.P. Sun, et al., A comparative study on constitutive relationship of as-cast 904L austenitic stainless steel during hot deformation based on Arrhenius-type and artificial neural network models, Comp. Mater. Sci. 67 (2013) 93-103.

RI PT

[25] S.I. Oh, S.L. Semiatin, J.J. Jonas, An analysis of the isothermal hot compression test, Metall. Trans. A 23 (1992) 963-975.

[26] P. Dadras, J.F. Thomas, Characterization and modeling for forging deformation of Ti-6Ai-2Sn-4Zr-2Mo-0.1 Si, Metall. Trans. A 12 (1981) 1867-1876.

SC

[27] X.P. Liang, Y. Liu, H.Z. Li, et al., Constitutive relationship for high temperature deformation of powder metallurgy Ti–47Al–2Cr–2Nb–0.2W alloy, Mater. Des. 37

M AN U

(2012) 40-47.

[28] R.L. Goetz, S.L. Semiatin, The adiabatic correction factor for deformation heating during the uniaxial compression test, J. Mater, Eng. Perform. 10 (2001) 710-717. [29] Academic Committee of the Superalloys, CMS, China Superalloys Handbook:

TE D

Part Ⅰ , China Zhijian Publishing House, Standards Press of China, 2012, pp. 870-872. [30] Y.T. Wu, Y.C. Liu, C. Li, et al., Deformation behavior and processing maps of Ni3Al-based superalloy during isothermal hot compression, J. Alloys Compd. 712 (2017) 687-695.

EP

[31] S.A. Sajjadi, A. Chaichi, H.R. Ezatpour, et al., Hot Deformation Processing Map and Microstructural Evaluation of the Ni-Based Superalloy IN-738LC, J. Mater, Eng.

AC C

Perform. 25 (2016) 1269-1275. [32] H.R. Ezatpour, A. Chaichi, S.A. Sajjadi, The effect of Al2O3-nanoparticles as the reinforcement additive on the hot deformation behavior of 7075 aluminum alloy, Mater. Des. 88 (2015) 1049-1056. [33] X.M. Chen, Y.C. Lin, D.X. Wen, et al., Dynamic recrystallization behavior of a typical nickel-based superalloy during hot deformation, Mater. Des. 57 (2014) 568-577. [34] H.B. Zhang, K.F. Zhang, Z. Lu, et al., Hot deformation behavior and processing map of a γ′-hardened nickel-based superalloy, Mater. Sci. Eng. A 604 (2014) 1-8. 23

ACCEPTED MANUSCRIPT [35] Y.H. Liu, Y.Q. Ning, Z.K. Yao, et al., Plastic deformation and dynamic recrystallization of a powder metallurgical nickel-based superalloy, J. Alloys Compd. 675 (2016) 73-80. [36] N.R. Jaladurgam, A.K. Kanjarla, Hot deformation characteristics and

RI PT

microstructure evolution of Hastelloy C-276, Mater. Sci. Eng. A 712 (2018) 240-254. [37] T. Sakai, M. Ohashi, Dislocation substructures developed during dynamic recrystallisation in polycrystalline nickel, Mater. Sci. Tech. 6 (1990) 1251-1257.

[38] T. Seshacharyulu, S.C. Medeiros, W.G. Frazier, et al., Hot working of commercial

SC

Ti–6Al–4V with an equiaxed α–β microstructure: materials modeling considerations, Mater. Sci. Eng. A 284 (2000) 184-194.

M AN U

[39] Y.V.R.K. Prasad, H.L. Gegel, S.M. Doraivelu, et al., Modeling of dynamic material behavior in hot deformation: Forging of Ti-6242, Metall. Trans. A 15 (1984) 1883-1892.

[40] K.A. Babu, S. Mandal, A. Kumar, et al., Characterization of hot deformation

TE D

behavior of alloy 617 through kinetic analysis, dynamic material modeling and microstructural studies, Mater. Sci. Eng. A 664 (2016) 177-187. [41] Z.X. Shi, X.F. Yan, C.H. Duan, Characterization of hot deformation behavior of GH925 superalloy using constitutive equation, processing map and microstructure

EP

observation, J. Alloys Compd. 652 (2015) 30-38. [42] Y.M. Kim, Y.B. Song, S.H. Lee, et al., Characterization of the hot deformation

AC C

behavior and microstructural evolution of Ti–6Al–4V sintered preforms using materials modeling techniques, J. Alloys Compd. 676 (2016) 15-25. [43] S.W. Zeng, A.M. Zhao, H.T. Jiang, et al., Flow behavior and processing maps of Ti-44.5Al-3.8Nb-1.0Mo-0.3Si-0.1B alloy, J. Alloys Compd. 698 (2017) 786-793. [44] X.N. Peng, H.Z. Guo, Z.F. Shi, et al., Study on the hot deformation behavior of TC4-DT alloy with equiaxed α+β starting structure based on processing map, Mater. Sci. Eng. A 605 (2014) 80-88. [45] V.A. Kumar, R.K. Gupta, S.V.S.N. Murty, et al., Hot workability and 24

ACCEPTED MANUSCRIPT microstructure control in Co20Cr15W10Ni cobalt-based superalloy, J. Alloys Compd. 676 (2016) 527-541. [46] G. Meng, B.L. Li, H.M. Li, et al., Hot deformation behavior of an Al–5.7wt.%Mg alloy with erbium, Mater. Sci. Eng. A 516 (2009) 131-137.

AC C

EP

TE D

M AN U

SC

superalloy 718, Mater. Sci. Eng. A 528 (2011) 3218-3227.

RI PT

[47] Y. Wang, W.Z. Shao, L. Zhen, et al., Hot deformation behavior of delta-processed

25

ACCEPTED MANUSCRIPT

Figure captions Fig. 1. Initial microstructure of as-forged U720LI alloy. Fig. 2. Schematic diagram of hot compression test in this work. Fig. 3. Schematic diagram of parameters in Eqs. 1-5.

RI PT

Fig. 4. Corrected flow stress curves due to the friction and adiabatic heating effects of U720LI alloy under various temperatures and strain rates: (a) 1100 oC and (b) 10s-1.

Fig. 5. Plots of ln ε& -ln σ p (a) and ln ε& -ln[sinh(α σ p )] (b) at different temperatures, (c)

SC

Plot of ln[sinh(α σ p )]-1000/T for quasi-γ phase and γ+γ' dual-phase microstructures and (d) relationship between lnZ and ln[sinh(α σ p )] of U720LI alloy.

M AN U

Fig. 6. Comparison between peak stresses experimentally determined and calculated by the established Arrhenius-type constitutive equation.

Fig. 7. Processing map of U720LI alloy at ε=0.8 with microstructure validation in different deformation regions.

TE D

Fig. 8. Microstructure of U720LI alloy at the strain of 0.8 with different hot compression conditions.

Fig. 9. (a) Distribution of DRX grain size under various deformation conditions and

AC C

EP

(b) relationship between Z and DRX grain size of U720LI alloy.

26

ACCEPTED MANUSCRIPT

Table captions Table 1 Parameters in Eq. 3 and 4[25, 28, 29]. Table 2 Kinetic parameters for γ+γ' dual-phase and quasi-γ phase microstructures of

AC C

EP

TE D

M AN U

SC

RI PT

U720LI alloy.

27

ACCEPTED MANUSCRIPT Highlights  Hot deformation behavior was investigated in a wide range of processing conditions.  Arrhenius-type model was established considering the dissolution of γ’

RI PT

precipitates.  Hot workability was discussed according to processing map and microstructure.

 Optimal processing conditions were determined in quasi-γ and γ+γ’ phase

AC C

EP

TE D

M AN U

SC

regions.