Sei sulla pagina 1di 10

Available online at www.sciencedirect.

com

Acta Materialia 59 (2011) 854–863


www.elsevier.com/locate/actamat

Creep and creep modelling of a multimodal nickel-base superalloy


James Coakley a, David Dye a,⇑, Hector Basoalto b
a
Department of Materials, Imperial College, South Kensington, London SW7 2AZ, UK
b
Advanced Forming Research Centre, DMEM, University of Strathclyde, Glasgow G1 1XQ, UK

Received 4 May 2010; received in revised form 25 August 2010; accepted 26 August 2010
Available online 16 November 2010

Abstract

The creep properties of different c0 -Ni3Al distributions in the Ni115 nickel superalloy produced by heat treatment have been exam-
ined. At the stresses and temperatures employed it is shown that particle bypass cannot occur by cutting or bowing and so presumably
occurs by a climb-glide motion. The creep strength of Ni115 is very poor at 800 °C and 360 MPa compared to 700 °C because the
increase in coarsening rapidly removes the bimodal structure. The Dyson creep model is a microstructure-based climb-glide bypass model
for unimodal distributions that links microstructural evolution (e.g. evolution of the particle dispersion) to the creep rate. The creep tests
are interpreted with the aid of the model and appear to suggest that the fine c0 , when present, strongly influences the dislocation motion.
A quasi-bimodal model is developed to account for bimodal distributions and the predictions compared to experiment. The model pre-
dicts a number of the observed experimental trends, and it’s shortcomings are identified in order to identify avenues for future
improvement.
Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Nickel alloys; Creep; Modelling

1. Introduction The serious possibility of mechanical failure in power gen-


eration due to creep has resulted in extensive research in
Nickel superalloys are predominantly employed in gas- creep mechanics and creep life predictions. Modelling the
turbine hot sections due to their excellent high temperature creep behaviour to date has largely been empirical [5–8], in
stability and mechanical strength. These properties are a which model parameters are numerically fitted to a limited
result of a two-phase structure, a face-centred cubic c creep database that can be used to interpolate behaviour
matrix that remains stable to its melting point and a coher- within the database. Empirical equations are based on the
ent ordered L12 Ni3Al c0 precipitate that hinders the flow of idea that creep is asymptotic to steady-state deformation.
dislocations. The mechanisms of creep vary with tempera- However, the use of these equations for creep predictions
ture, stress, and microstructure. For example, under low of complex engineering alloys is not ideal for two reasons.
temperature, high stress creep conditions, complex disloca- First, the creep curves of these alloys do not contain a steady
tion–precipitate shear now appears to be the dominant state regime wherein the microstructure does not evolve, and
deformation mechanism [1–4], while deformation under secondly, one must question the validity of extrapolating
higher temperature, lower stress creep conditions generally short-time creep test data to long-time behaviour, due to
occurs by particle bypass, i.e. thermally activated disloca- the non-uniqueness of the fitted parameters. Over the last
tion climb and/or Orowan bowing. 60 years various creep prediction methods have been devel-
oped, but empirical fitting methods have remained common
practice, despite well-known doubts of accuracy with extrap-
olation methods.
⇑ Corresponding author. Tel.: +44 20 7594 6811. The shortcomings of the empirical fitting methods and
E-mail address: david.dye@imperial.ac.uk (D. Dye). power law rate equations have led to the development of

1359-6454/$36.00 Ó 2010 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2010.08.035
J. Coakley et al. / Acta Materialia 59 (2011) 854–863 855

the Dyson model [9], a microstructure-based damage 2. Experimental


mechanics approach to model creep. This generic constitu-
tive description of creep is composed of two parts. First, a The alloy composition is presented in Table 1. Following
creep rate equation is derived to overcome the shortcom- the standard heat treatment (1.5 h/1190 °C/AC + 6 h/
ings of power law models. The creep rate is expressed as 1100 °C/AC), 31 subsequent heat treatments were performed
a function of not only its operating conditions of stress and characterized in a temperature range of 700 °C  1000 °C
and temperature, but also the underlying microstructure and an exposure time ranging from 6 h to 7500 h followed by
(particle size, volume fraction, dislocation density, etc.) air-cooling (AC). This work and the characterization tech-
that controls deformation. Having established the relation- niques are described in detail in Ref. [21]. From the various
ship between the macroscopic strain rate and microstruc- c0 distributions, it was decided to perform creep tests on five
ture, the kinetic equations describing the evolution of the different microstructures (Table 2). The microstructures pro-
microstructure are constructed. Coupling these differential duced from the five heat treatments are shown with secondary
equations predicts the resulting strain–time trajectories of emission scanning electron microscopy (SEM) in Fig. 1. For
an evolving microstructure. This approach has been suc- SEM, the samples were ground, polished and the c phase
cessful in predicting the creep properties of some current was electro-etched with a 2.5% phosphoric acid in water at
unimodal alloys [9,10]. 2.5 V, 2 A for one second at room temperature.
A number of the current disc alloys possess multimodal The primary particle distribution functions were
c0 distributions, e.g. RR1000 and Udimet720, with large c0 obtained from thresholding SEM images in the ImageJ
being attributed to sub-solvus heat treatment, and fine c0 software package to produce a binary c and c0 image. Indi-
precipitating from solution during cooling. The pinning vidual particle areas were determined, allowing equivalent
of dislocations and the initial creep resistance are generally particle radii to be calculated. In addition, the area fraction
attributed to dislocation line pinning by the fine c0 . During was measured. At least three images were analysed from
gas-turbine operating conditions, this multimodal system different grains for each heat treatment. The area fraction
coarsens to a unimodal distribution, at which point the is representative of the alloy, thus one can assume this is
large c0 is primarily responsible for pinning of dislocations, the volume fraction of the primary particles Vf,p. Transmis-
with an associated loss in creep strength due to an increase sion electron microscopy (TEM) micrographs were used to
in inter-particle distance [13]. There are two important characterize the fine c0 in the as-delivered sample. Again,
questions that have remained largely unanswered. First, ImageJ was used to threshold the c and c0 , and equivalent
how does a multimodal distribution coarsen with time? particle radii calculated. It is difficult to calculate the vol-
There is very little experimental data published and very lit- ume fraction of fine c0 by TEM, as only a small area is
tle modelled, as all Lifshitz–Slyozov–Wagner (LSW)-based being sampled in order to resolve the particle sizes
coarsening theories have been for unimodal systems [14– accurately.
20]. Secondly, how does one then model the loss in creep There are two possible suggested values for the overall
strength as the system shifts from a multimodal to a uni- volume fraction, Vf,t, of Ni115. Porter and Ralph [22]
modal distribution? report a volume fraction of 50%. Using the ICP-OES and
In this paper five different heat treatments are employed CS results in Thermo-Calc with an appropriate thermody-
to produce five different c0 distributions in cylindrical bars namic database yields a prediction of 60% at 600 °C.
of polycrystal nickel superalloy Nimonic 115 (Ni115), Finally, SEM analysis of the unimodal samples following
which has an initial multimodal distribution. A matrix of long-term ageing gives an area fraction 60% [21]. Since
creep tests was constructed and performed to study the the calculated Vf,t from Thermo-Calc and the unimodal
effects of the various distributions on creep response at ele- SEM image analysis are in approximate agreement, the
vated temperatures and stress. The coarsening kinetics of alloy is assumed to be 60% volume fraction in this work.
this alloy have been studied previously [21], and the results The volume fraction of secondary particles is calculated
from this work are used to both interpret the creep from the assumption
response of the different samples and to appropriately alter
the Dyson model for a multimodal system. V f ;t ¼ V f ;p þ V f ;s ð1Þ

Table 1
Composition of Ni115, in wt.%, both from the literature and for the material used here, measured by inductively coupled plasma optical emission
spectrometry (ICP-OES) and a LECO carbon sulphur (CS) analyser.
Element (wt.%) Al C Cr Co Cu Fe Mo Ni Si Ti Zr
Minimum [11] 4.5 0.12 14.0 13.0 0.2 1.0 3.0 bal 0.4 3.5 0.03
Maximum [11] 5.5 0.2 16.0 16.5 max max 5.0 bal max 4.5 0.06
Minimum [12] 4.5 0.12 14 13 60.20 61.0 3.0 54.0 61.0 3.5 60.15
Maximum [12] 5.5 0.2 16 15.5 60.20 61.0 5.0 54.0 61.0 4.5 60.15
This work 4.85 0.15 15.0 15.2 0.2 0.4 3.64 bal 0.003 3.96 0.042
856 J. Coakley et al. / Acta Materialia 59 (2011) 854–863

Table 2
Heat treatments performed on Ni115 samples following the standard heat treatment (HT; the resulting c0 volume fractions Vf, mean radii R and inter-
particle distance k of the primary distribution, secondary distribution, and overall distribution are presented).
Comments Label HT time/temp. Primary distribution Secondary distribution Overall distribution
(h)/(°C)
Vf,p (%) Rp (nm) kp (nm) Vf,s (%) Vf,seff (nm) Rs (nm) ks (%) Vf,t (nm) Rt kt (nm)
As delivered – bimodal A –/– 42 90 53 18 31 5 5 60 7 1.7
Bimodal B 6/1000 42 110 65 18 31 5 5 60 7 1.7
Bimodal C 12/1000 42 150 88 18 31 5 5 60 7 1.7
Bimodal D 48/1000 42 210 123 18 31 5 5 60 7 1.7
Unimodal E 48/900 60 114 26 0 0 – – 60 114 26

R1
Rx f ðRx ; tÞdR
A B Rx ¼ 0R 1 ð2Þ
0
f ðRx ; tÞdR
The inter-particle distance kx is taken as the square lattice
distance between two spheres in a system where all particles
have radius R ¼ Rx
1=2
kx ¼ 1:6Rx ½ðp=4V f ;x Þ  1Þ ð3Þ
For the inter-particle distance between secondary particles,
the effective volume fraction of the secondaries is first cal-
culated, Vf,seff = [Vf,s/(1  Vf,p)]  100, i.e. this would be
C D the volume fraction of particles if one were to take an im-
age of just the channels between the primary c0 precipitates.
ks is then calculated with Eq. 3, but using Vf,seff and Rs .
By performing heat treatments at 1000 °C at different
exposure times (HT samples B–D, Table 2), it was possible
to produce samples with different primary distributions,
but identical secondary distributions, as the secondary par-
ticles precipitate from dissolution on cooling from 1000 °C.
Thus it is possible to determine the strengthening contribu-
tions of each distribution during creep by using the heat
E treatments described in Table 2. A unimodal distribution
(HT sample E) was also subjected to creep testing for com-
parison with the bimodal tests.
The creep test matrix is shown in Table 3. All creep tests
1µm were performed under constant load conditions, and are
assumed to be representative of constant stress, in line with
common practice. All five samples were tested at 360 MPa
and 800 °C (creep condition I). Two further creep tests
were performed on identical bimodal samples (sample D)
at 300 MPa and 250 MPa to represent creep at lower stres-
Fig. 1. The five different Ni115 microstructures used for creep testing. The ses, tests II and III respectively. Finally, two different
heat treatments performed and distributions achieved are presented in bimodal samples (HT samples B and C) were tested at
Table 2 with corresponding labels.
360 MPa and 700 °C to investigate the creep response at
a lower temperature, where the coarsening kinetics are
for the bimodal samples. It was assumed that the secondary much slower [21].
distribution is identical for all bimodal samples in this
work, as they are precipitating on cooling from above Table 3
1000 °C. The creep tests performed (under constant load conditions) on Ni115
The cumulative distribution function for each sample which have had different prior heat treatments.
heat treatment was smoothed using a Weibull function. Heat treatment Test condition Stress (r) (MPa) Temp. (T) (°C)
The particle distribution function (PDF) is obtained from A–E I 360 800
the derivative of the Weibull. The mean radius of each dis- D II 300 800
tribution, Rx , where x = p, s, t corresponds to primary, sec- D III 250 800
B–C IV 360 700
ondary and total distributions respectively, is given by
J. Coakley et al. / Acta Materialia 59 (2011) 854–863 857

3. The Dyson creep model cations being punched out at the precipitate/matrix
interface. (vii) A continuum damage mechanics (CDM) ap-
The high-temperature creep behaviour of precipitate proach is adopted in order to couple the relevant micro-
strengthened alloys is a kinetic phenomenon. The proposed structural rearrangements during deformation to the
modelling approach of Dyson [9] seeks to derive an expres- creep rate [10,26,27].
sion for the creep rate as a function of the microstructure Dyson [9] arrives at a microstructurally explicit creep
by taking into account the dislocation–particle interac- strain rate equation given by
tions. To complete the framework, one requires knowledge " 1=2 #  2 
of the evolution of the dominant microstructure parame- 1 p rb k
_ c ¼ 1:6DC j M qm V f ð1  V f Þ  1 sinh
ters that influence creep. The benefits of a physics based 4V f MkT
microstructure creep approach are: (i) one can account ð5Þ
for varying operating conditions; (ii) one can predict the
loss of creep strength due to alterations in microstructure; where D is the coefficient of diffusion, Cj is the jog density
and (iii) by relating the creep strength to microstructural coefficient, M ¼ r=s ¼ dc=d is the Taylor factor, r is the
properties, the model may accelerate the development of applied stress, s is the shear stress, c is the shear strain rate,
new materials. Dyson [9] recently published his most thor- Vf is the volume fraction, b is the Burgers vector, k is
ough derivation of the minimum creep rate for thermally Boltzmann’s constant and T is temperature. The definition
activated climb-glide creep deformation. The theory of of inter-particle distance (Eq. 3) has been substituted into
the model is briefly reviewed here to illustrate the assump- the pre-sinh term in Eq. (5).
tions made in developing the unimodal model, while the It is important to note that the inter-particle distance k
derivation of the minimum creep rate is omitted as it is appears within the sinh term, thus the creep strain rate _ c
published elsewhere [9]. A clear understanding of the is sensitive to the inter-particle distance, and the associated
assumptions made in deriving the unimodal model is changes in inter-particle distance as the precipitates coar-
required in order to develop a physically faithful multi- sen. A climb-dominated model [24,25] would instead focus
modal creep model. on the particle size, concluding that large particles would
The model of Dyson [9,10] stems from the work of inhibit creep whereas this elaboration comes to the oppo-
Kocks et al. [23] for jerky glide, where jerky glide refers site conclusion.
to dislocation climb-glide motion through a precipitate The model can be adjusted to predict primary creep.
strengthened alloy at stresses below yield. As the model is Primary creep is described as a back-stress rk from the
based upon the climb-glide mechanism, it is only applicable plastically deforming creeping matrix to hard elastically
at stresses below those required for particle shear or deforming precipitates. rk is simply subtracted from the
Orowan bowing. applied stress r. The back-stress approaches a saturated
The Dyson model describes dislocation creep by the fol- limit rk , assumed to be caused by dislocation punching at
lowing assumptions. (i) Dislocation creep in precipitate- the precipitate/matrix interface. To complete the frame-
strengthened alloys occurs by the jerky glide mechanism. work, the evolution of the dominant microstructure
Mobile dislocations are either gliding or pinned at precipi- parameters that influence creep are required. CDM is cou-
tates, where they can be released by thermal activation. pled with the creep rate equation to quantify the rate of
evolution of each relevant microstructural feature and to
qm ¼ qg þ qc ð4Þ predict the associated loss in creep strength [10,26,27].
i.e. the mobile dislocation density qm is composed of the The complete creep rate equation is given by
gliding dislocation density qg and the climbing dislocation  
rð1  H Þ
density qc. (ii) Most dislocations are pinned. (iii) The mac- _ c ¼ ð1 þ D0d Þ_0 sinh ð6Þ
r0 ð1  Dp Þð1  Dc Þ
roscopic strain rate is provided only by those dislocations
gliding, thus the gliding dislocation generation and trap- where
ping rates need to be derived. (iv) There is always a certain h 1=2 i
fraction of dislocations in a position to escape the pinning _ 0 ¼ 1:6DC j M 1 qm V f ð1  V f Þ p=4V f 1 ð7Þ
from precipitates. This is in contrast to the derivation of
is the pre-sinh term, H = rk/r is a dimensionless hardening
Ansell and Weertman [24,25], which describes climb and
parameter, the saturated hardening limit H  ¼ rk =r; H  ¼
glide as occurring in series, leading to the (controversial)
2V f =ð1 þ 2V f Þ and a stress constant r0 ¼ ðMkT =b2 kÞ. Using
prediction that the creep strain rate is determined by the
a platen model, the evolution of the hardening parameter to-
climbing process. (v) The escape frequency is dependent
wards the saturation limit is given by
on the free-energy change associated with the escape pro-
 
cess and the probability that the energy fluctuation re- V fE H
quired occurs. (vi) Primary creep is described via a back- H_ ¼ 1   _ c ð8Þ
r H
stress from a plastically deforming creeping matrix to hard
elastically deforming precipitates. The back-stress ap- Three damage mechanisms observed in nickel
proaches a saturated limit, assumed to be caused by dislo- superalloys are incorporated into the model: cavitation
858 J. Coakley et al. / Acta Materialia 59 (2011) 854–863

Dc = 1  (Ac,i/Ac), particle coarsening Dp = 1  (ki/k), and where cAPB is the anti-phase boundary energy. A range of
an unbounded mobile dislocation multiplication term D0d ¼ bowing and cutting stresses for the secondary particles
ðqm;i =qm Þ  1, where Ac,i,ki and qm,i are the initial cavitated are calculated in Table 4, corresponding to the cAPB values
area, inter-particle distance and dislocation density respec- quoted in literature [35–40] and two Vf,t values, as Porter
tively, and Ac,k and qm are the current cavitated area, inter- and Ralph quote a different value to that measured and cal-
particle distance and dislocation density. The evolution of culated in this work [22].
the damage parameters are defined as D_ 0d ¼ C _c , where C As one would expect for very fine c0 particles in a high vol-
is a material constant termed the dislocation multiplication ume fraction alloy, rc(cutting)  rc(bowing). The shear modu-
coefficient [10]. Ashby and Dyson [26] derive, based upon lus, G = E/2(1 + l), is approximated from CMSX-4 at
K 4
LSW theory, D_ p ¼ 3p ð1  Dp Þ , where Kp is a kinetic rate 700 °C, where E  100 MPa is Young’s modulus and
constant. Dyson [27] defines D_ c ¼ kf c;u , where f,u is the uni- l  0.45 is Poisson’s ratio [41]. These values are used due
axial strain at fracture and kc has an upper bound 0.33, to a lack of data for Ni115, and are taken to be representative
when all transverse boundaries are cavitated. of high volume fraction superalloys. The range in rc(cutting) is
vast, from 86 to 759 MPa. The anti-phase boundary energy
4. Creep test results and discussion values quoted above are for pure NixAl100x, where x =
74–78, and also samples alloyed with a ternary element, i.e.
The 1000 °C 6/12/48 h + AC heat treatments produce 63 at.% Hf, Ta, Pd, B [35]. The upper bound is given by
samples of differing primary distributions but identical sec- the ternary alloyed samples. The c0 phase of Ni115 is solid-
ondary distributions to the as-delivered samples. Previ- solution strengthened with 3.5 wt.% Ti, so one would expect
ously, the strength attributed to each distribution has the anti-phase boundary energy to be in the region of the
been deduced by performing mechanical tests on bimodal upper bound, thus it is fair to assume that particle shear can-
samples and unimodal samples with the secondary particles not occur at macroscopic stresses r 6 360 MPa.
removed by heat treatment [13]. To the authors’ knowl- The full creep curves to failure are ambiguous (Fig. 2a).
edge, the heat treatments and creep testing described in this First, the ductility scatter ranges from 6% in the uni-
paper represent the first time the primary distribution has modal sample (sample E) to 24% in sample D, a bimodal
been altered while keeping the secondary particles con- sample. Nimonic 115 shows a decrease in ductility when
stant, which may show conclusively whether the secondary M23C6 grain-boundary film is present [42]. Sample A does
particles determine the creep strength of the alloy, in the not contain grain boundary carbide film; however, samples
absence of particle shear. B–E do. Thus the scatter in ductility is not due to the grain
The creep curves for the five different heat treated sam- boundary morphology, but may be due to the uniqueness
ples (A–E), at 800 °C 360 MPa, are shown in Fig. 2a–c. In of void growth in each individual sample. The authors
order to interpret the creep results of these microstructures, acknowledge that further creep tests are required under
one must first assess the creep deformation mechanisms: is identical creep conditions in order for the results and
deformation occurring primarily by dislocation particle by- drawn conclusions in this work to be definitive.
pass by climb, by Orowan bowing or by particle shear? At The following observations are made regarding the rela-
intermediate stresses it is well established that particle shear tive creep strengths of each sample: (i) Samples B–D only
of primary-size particles does not occur [28–31]. The fixed- differ in primary c0 distribution, and the creep curves of
line tension model, derived from the energy per unit length samples B–D in Fig. 2a imply that a larger Rp, and a cor-
of dislocation, can be used to estimate the critical stress for responding larger kp, lowers the creep rate. This is an unex-
Orowan bowing [32–34] pected result, as an increase in c channel size would be
Gb expected to decrease the dislocation curvature and hence
scðbowingÞ ¼ ð9Þ the pinning stress. (ii) The coarsening studies of Ni115
L
[21] aid the interpretation of the creep responses of the dif-
where G is the shear modulus, b is the Burgers vector and ferent distributions. A graph of Ni115 mean radius R
L ¼ k þ 2R. against time t1/3 at 700 °C and 800 °C, including model pre-
For dislocation shear of fine c0 , it is assumed that the dis- dictions, is reproduced for this purpose in Fig. 3. The
locations are weakly coupled. This means that the disloca- bimodal distribution mean radius initially increases as the
tions travel in dislocation pairs, the leading dislocation fine c0 dissolve and the system coarsens. A plateau is
creating an anti-phase boundary (APB) in the c0 that the observed as the distribution shifts to unimodal. In this
trailing dislocation removes. The dislocations are spaced region, there is no reduction in the number of particles,
a distance v apart, and are never simultaneously within simply a transfer of c0 from the smaller primary particles
the same fine c0 particle. Brown and Ham derive an expres- to the larger primary particles (the fine c0 have fully dis-
sion for the shear stress to cut a particle from a balance of solved). The distribution essentially broadens and the mean
forces on the dislocation pair [34] radius remains constant. The plateau region ends when
" 1=2 #
cAPB 8cAPB V f R particles begin to dissolve, and the unimodal mean radius
scðcuttingÞ ¼ Vf ð10Þ increases at a rate proportional to t1/3. Thus, for the
2b pGb2
as-delivered bimodal sample, sample A, the fine c0 are fully
J. Coakley et al. / Acta Materialia 59 (2011) 854–863 859

26 o D o
24 (a) I: 360MPa 800 C (b) I: 360MPa 800 C 1.4
22 C
1.2
20
18 B 1.0
16 A
14 B D 0.8
12 C
10 0.6
E
8
E 0.4
6
A
4 0.2
2
0 0.0
0 40 80 120 160 200 240 0 10 20 30 40 50 60 70

(c) I: 360MPa 800 C


o
(d) I: 360MPa 700 C
o 1.0
0.14 D
B C 0.9
0.12 0.8
C
0.10 0.7
B
E 0.6
0.08
0.5
A
0.06 0.4
0.3
0.04
0.2
0.02 0.1
strain (%)

0.00 0.0
0 .0 1 .0 2 .0 3 .0 4 .0 5 .0 6 .0 0 1000 2000 3000 4000
28
o
13 (e) Model Predictions (f) Sample D - 1000 C 48h HT III 26
o
12 I: 360MPa 800 C 24
11 I 22
10 II 20
9 18
8 16
7 C B E A 14
D
6 12
5 10
4 8
3 6
2 4
1 2
0 0
0 100 200 300 400 500 0 400 800 1200 1600

14 (g) Model Predictions II


o
Sample D - 1000 C 48h HT
12 III

10
I
8

0
0 50 100 150 200 250 300 350
time (hrs)
Fig. 2. The creep tests and associated model predictions for Ni115 with different initial microstructures, labelled A–E (Table 2). (a–c) Creep curves from
test I: 360 MPa and 800 °C presented on different axis in each graph for samples A–E. (d) Test IV: 360 MPa and 700 °C on samples B and C. (e) The
Dyson bimodal creep model predictions at 360 MPa and 800 °C, corresponding to the experimental results of (a)–(c). (f) Tests I–III, 360 MPa and 800 °C,
300 MPa and 800 °C, 250 MPa and 800 °C respectively. These three creep tests were performed on sample D. (g) The Dyson bimodal creep prediction for
the creep curves in (f).
860 J. Coakley et al. / Acta Materialia 59 (2011) 854–863

Table 4 there is still considerable fine c0 in the first 6 h, the plateau


Calculation of the bowing and shear stresses, rc(bowing) and rc(cutting), for region signifying a unimodal distribution. Thus, at 800 °C
Ni115 secondary particles.
one can infer that the similar creep responses are due to
cAPB (J/m2) 0.1 0.1 0.2 0.2 0.3 0.3 the fine c0 being present in all bimodal samples at short
Vf,t (%) 50 60 50 60 50 60
Vf,s (%) 8 18 8 18 8 18
times. It is unclear why the unimodal sample (sample E)
Vf,seff (%) 16 31 16 31 16 31 possesses the highest creep strength in the first 4 h. Thus,
ks (nm) 10 5 10 5 10 5 the experimental data at 800 °C 360 MPa suggest that the
sc(bowing) (MPa) 445 593 445 593 445 593 fine c0 , when present in significant amounts, has a strong
rc(bowing) (MPa) 1378 1847 1378 1847 1378 1847 influence on dislocation motion. This argument is rein-
sc(cutting) (MPa) 28 21 104 111 213 245
rc(cutting) (MPa) 86 66 323 344 660 759
forced by the creep curves at 360 MPa 700 °C (Fig. 2d).
Again, the coarsening data in Fig. 3 aid the interpretation
A range of values have been calculated to account for the range in values
quoted for the anti-phase boundary energy, cAPB, and the disputed Ni115
of the creep curves. Fig. 2d shows that samples B and C
volume fraction. The other constants used were Rs ¼ 5 nm; M ¼ 3:1; have identical creep strength up to 4000 h. There is simi-
G ¼ 35 GPa and b = 2.54 °Å. lar fine c0 present in both samples in this time period, while
the primary particles of the B and C samples are different.
The experimental evidence suggests that fine c0 , when pres-
ent in large volume fractions, is significant in controlling
the creep rate, and when r < rc(cutting).
(vi) A possible hypothesis for the order of the bimodal
strengths shown in Fig. 2a is as follows. The fine c0 in sam-
ple A fully dissolves in the shortest time, then sample B,
then sample C and finally sample D (this is predicted by
the applied mean field coarsening theory). As the fine par-
ticles dissolve, Vf,s and ks decrease, while Vf,p and kp
increase. Thus the strengthening effect due to the larger
particles is increasing during this transition while that of
the fine particles is decreasing. The fine c0 of sample A dis-
solve in the earliest time frame, and strengthening is due to
the primary particles. However, sample A remains stron-
gest due to the strong distribution of primary particles
(smallest kp). Sample B is weaker than A for, once its dis-
Fig. 3. As-delivered Ni115 particle coarsening experimental data and
tribution coarsens to unimodal, the primary particles
coarsening model predictions at 700 °C and 800 °C. The results are plotted
in t1/3, as is convention for diffusion-based coarsening theory. This graph impart less strength than in A; even at short times
is reproduced from Ref. [21]. (Fig. 2b and c) the primaries in A impart more strength
than the secondaries in B. Samples C and D are stronger
than B even though they have a larger kp, as the strength-
dissolved within t  40 h (t1/3  3.4 h1/3), so the long-time ening secondary particles are present for a longer times. C
behaviour is not controlled by the fine particles at 800 °C and D are still weaker than sample A for, although the sec-
and 360 MPa. It is not possible to draw the same conclu- ondary particles may be present in samples C and D for
sion for samples B–D, as the fine particle dissolution time longer, the strength imparted from the primary particles
is likely to be a function of kp. (iii) Sample A has a much in sample A is larger. Thus it appears that the creep
lower creep rate than sample B. These samples have similar strength evolution during the transition from multimodal
secondary distributions; however, A has finer primary par- to unimodal is a delicate function of, and highly dependent
ticles, with lower Rp and kp. (iv) Samples A and E show on, both kp, ks and the time taken to move from a multi-
lower creep rates than B–D. The lower strength imparted modal to unimodal system.
by sample A’s coarser primary particles compared to sam- The above interpretation of the creep results and the
ple E is compensated by fine secondary particles. As the corresponding creep model is based upon the assumption
fine particles in sample A dissolve, the sample then presum- that shear of the secondary particles is not occurring (see
ably has a similar microstructure to sample E, so they creep Table 4). However, it may be argued that the secondary
similarly. Therefore, A and E indicate that a larger kp is particles are so fine that they are being sheared. This offers
detrimental to the material, as expected. an alternative interpretation to the creep data at 800 °C,
(v) Analysing the short-time-scale creep responses 360 MPa. If secondary shear of the secondary particles
proves informative, Figs. 2b and c. For the first hour, all occurs, it is the primary particles alone that act as pinning
bimodal samples (A–D) show the same creep response points, strengthening the alloy. Thus sample E, with the
(Fig. 2c). Samples B–D show identical responses to 6 h, higher volume of primary particles and the smallest mean
with B and D showing identical responses to 40 h. Refer- radius of primary particles, is expected to be strongest, as
ring to the coarsening study performed at 800 °C (Fig. 3), is seen in experiment (Fig. 2a). Sample A, with a similar
J. Coakley et al. / Acta Materialia 59 (2011) 854–863 861

primary radius to sample E but a lower volume fraction of As a first attempt to model the coarsening data in
primary particles, has a shorter creep life than sample E but Fig. 2a, a quasi-bimodal model is formulated as follows:
a longer creep life than samples B–D, which contain larger (i) The initial mean radius of the overall distribution was
primary particles and the same primary particle volume found from Eq. 2 and the initial inter-particle distance
fraction. If particle shear of secondary precipitates is occur- was calculated from the mean radius of the overall distribu-
ring, one would expect sample B to be stronger than sam- tion and using the overall volume fraction, according to
ples C and D in creep as sample B contains smaller primary Eq. 3. (ii) An LSW-based coarsening rate does not apply
K
particles; however, this is not seen in experiment. Sample C to a bimodal system: D_ p ¼ 3p ð1  Dp Þ4 does not hold for
is expected to be stronger than D if shear is occurring; how- all t. The coarsening rate of Ni115 at 800 °C is shown in
ever, this is also not seen in experiment. If particle shear is Fig. 3. The coarsening rates of the other bimodal systems
occurring, the bimodal creep model reduces to a unimodal are similar (as this is a mean field model and therefore does
creep model. not capture local environment correctly), but with the pla-
teau region occurring at a larger mean radius and with the
plateau occurring for a longer time. The radius at which the
5. Accounting for bimodal precipitate distributions in the plateau region occurs is approximately equal to the initial
creep model, and comparison to experiment mean radius of the primary distribution. The mean radius
3
coarsening rate K p ¼ R13 dRdt ¼ 8:68  105 s1 prior to the
The alteration of the Dyson unimodal creep model to a plateau region and Kp = 0 s1 when R  Rp;i for each bimo-
0

bimodal model is not a trivial task; each distribution will dal system, where R0 is the initial overall mean radius and
contribute to the overall strength of the alloy, but the Rp;i is the initial mean radius of the primary distribution.
extent of each contribution needs to be determined. Two The constants required for Ni115 creep predictions are
limits exist for the strengthening contributions of each dis- given in Table 5. These constants are identical for all initial
tribution. Initially, when there is significant amount of fine Ni115 microstructures and creep conditions. The only ini-
c0 , it is generally agreed that the fine c0 controls creep tial microstructural input that changes for the initial distri-
strength, and this has been demonstrated in Section 4. Fur- butions is the initial mean radius R, which was different for
thermore, at long times the distribution will become uni- the bimodal and unimodal models. All the bimodal sam-
modal and the creep model should reduce to the Dyson ples have an equivalent initial mean radius R due to the
unimodal model. The contributions in the transient regime high number of secondary particles compared to primary.
between these two limits require formulation. The initial coarsening rate is the same for all bimodal dis-
For the short-time-frame small-particle-dominated behav- tributions; however, the time at which the coarsening rate
iour, two assumptions are made: first, that the small particles plateaus (RðtÞ ¼ Rp;i ) increases with Rp;i . The unimodal
control the deformation rate, and secondly, that no particle coarsening rate is Kp = 2.84  1007 s1 from the long-
shear occurs. The k term within the sinh is due to forces term coarsening data at 800 °C following the plateau
exerted on the trapped dislocation to escape pinning, in this region. The final inputs are test temperature and stress.
case between the small particles, such that r0 ¼MkT =ðb2 ks Þ. The microstructural inputs are estimates for a nickel
The pre-sinh term is also affected, and Eq. 7 is rewritten as superalloy and are not necessarily specific to Ni115, as a
h 1=2 i number of the required material specific parameters are
_ 0 ¼ 1:6DC j M 1 qm V f ;s ð1  V f ;t Þ p=4V f ;s 1 ð11Þ unknown. The sources of these estimates are as follows:
(i) The value of D0 is the lower bound of self-diffusion in
where the first Vf term is from the probability of a disloca- nickel [43]. (ii) Equating Cj to 1 is an assumption that all
tion–particle encounter, the second Vf term is from the
assumption that plastic deformation occurs in the matrix Table 5
alone and the final Vf term is from substitution of the in- Microstructural parameters used in the Ni115 creep predictions using the
altered Dyson model for creep of a bimodal system.
ter-particle distance. If one assumes that the kinematic
back stress is a contribution from both distributions, H_ Volume fraction Vf 0.6
Taylor factor M 3.1
and H* remain unchanged, and are given by the total Vf,t. Pre-exponential diffusivity D0 (M2 s1) 1.9  104
The transient coarsening regime remains an area of Jog density Cj 1.0
future work. It is possible to define an effective inter-parti- Mobile dislocation density qm (M2) 1010
cle distance k12 ¼ k12 þ k12 ; however, relating the change in Boltzmann’s constant k (m e2kg s2 K1) 1.38  1023
eff p s Burgers vector b (Å) 2.54
Vf,s and Vf,p as the system coarsens into the pre-sinh term Gas constant Rg (J mol1K1) 8.314
is difficult. This is required to keep the physical meaning Activation energy Q (kJ mol1) 305
of the model correct, although it may be argued that the Young’s modulus E (GPa) 160
Dislocation multiplication coefficient C 100
creep rate is essentially dominated by the sinh term, so it
Geometrical constant for equiaxed grains kc 0.33
is the definition of keff and the redefined coarsening Minimum creep failure strain f 0.1
damage rate D_ p that will determine the accuracy of the The other variables used are the initial overall mean radius Ri , the mean
model. radius coarsening rate Kp, macroscopic stress r and temperature T.
862 J. Coakley et al. / Acta Materialia 59 (2011) 854–863

encounters will form jogs. (iii) The mobile dislocation den- reflected in experiment (Fig. 2b). Thus the similar creep
sity is that of nickel [44,45]. (iv) The activation energy for responses of samples A and E seen in experiment is
diffusion and jog formation is an estimate from an exten- reflected by the model. However, the model over-predicts
sive literature review [21]. The creep rate is extremely sen- the extent of primary creep in the unimodal distribution.
sitive to the activation energy due to the exponential With regard to the other bimodal distributions, samples
term. (v) The elastic constant is that of Ni115 at 800 °C B–D, sample A is predicted to be stronger than samples
[11]. (vi) The dislocation multiplication coefficient C is B–D, reflected in experiment. Samples B–D take longer
unknown, and to date has been estimated by empirical fit- to coarsen to the respective plateau regions than sample
ting. Sondhi et al. [13] used a value C = 20 for creep predic- A, as these plateau regions lie at a larger mean radius
tions of IN100 at 700 °C and Basoalto et al. [10] used a due to the different initial primary distributions. It is pre-
value C = 67 for IN738LC from 750°C to 950 °C, while dicted that sample B reaches the plateau at 100 h and
Manonukul et al. [46] estimated C = 496 at 700 °C, C = the coarsening stops, while samples C and D are still rap-
435 at 750 °C and C = 10 at 800 °C for the polycrystal idly coarsening. This is why sample B is predicted to be
superalloy C263. Thus C = 100 is a rough estimate for stronger than C and D, while C and D do not reach the
Ni115 at 800° C. (vii) The geometrical constant kc  0.33 respective plateau regions in this lifetime prediction; they
is an upper limit when all transverse boundaries have cav- are thus predicted to have identical strengths. Experimen-
ities [27]. (viii) The material specific minimum creep failure tally (Fig. 2a), sample B has the lowest creep strength,
strain was written to one decimal place, Fig. 2a. which is unexpected. Samples C and D have almost identi-
The time-independent variables, i.e. the diffusion coeffi- cal creep strengths, as predicted by the model, and this is
cient D = D0exp(Q/RgT), the initial inter-particle dis- explained by the coarsening kinetics experienced by the
tance ki, and thus the pre-sinh term _ 0 , the saturated samples before the coarsening plateau region is reached.
hardening parameter H* and the stress constant r0, are first The primary creep regime and creep rate at short time
calculated,. The initial creep strain rate _ c ð0Þ is calculated frames is underestimated by the model. This is due to the
at t = 0,c = 0, Dd = 0, Dp = 0, Dc = 0, H = 0, as well as model shortcoming in the very short time frame. It is
the evolution rates of the hardening and damage parame- assumed that the strengthening from both distributions
ters D_ d ; D_ p , D_ c ; H_ . An Euler method is then implemented can be represented by an average radius with total volume
by specifying a time step dt for all the rate equations, ulti- fraction. This may be applicable in the transient regime,
mately predicting the creep strain c against time t. and in the unimodal model; however, when there is a lot
Three creep tests were performed, at 360 MPa and of very fine c0 , this distribution essentially dominates the
800 °C, 300 MPa and 800 °C, 250 MPa and 800 °C, on creep rate. Thus the effective volume fraction will be much
identical bimodal samples (sample D) to examine the effect lower than the total volume fraction, resulting in a predic-
of lower creep stresses (Fig. 2f). The microstructural tion of creep strength that is greater than what is actually
parameters in Table 5, along with the appropriate stresses occurring. This remains an area for future development
and temperature (360 MPa 800 °C, 300 MPa 800 °C, of the bimodal creep model.
250 MPa 800 °C), the initial radius and the radius at which
the plateau region occurs, were specified in the creep 6. Conclusions
model. The model predictions are shown in Fig. 2g. The
effect of decreasing stress was reproduced, i.e. the creep Five different initial distributions of 60% Vf Nimonic
life-times increased appropriately. The lifetime predictions, 115 were produced from differing heat treatments. Four
although they are short by up to a factor of 5, are quite rea- of these distributions were bimodal, with primary distribu-
sonable considering how many of the microstructural tions differing from sample to sample (Rp ¼ 90; 110;
parameters were estimated. The parameters require optimi- 150; 210 nm), but identical secondary distributions
zation; however, there are not enough creep data to do this (Rs ¼ 5 nm). The fifth sample was a unimodal sample with
without over-fitting the model parameters. R ¼ 114 nm for comparison. A number of identical creep
Fig. 2e shows the predictions for all the samples creep tests were performed on different samples and the results
tested at 360 MPa, 800 °C (Fig. 2a–c). The predictions were compared to a multimodal microstructure-based
show some agreement with experiment trends. The model creep model developed in this work.
predicts that all the bimodal samples have the same Under low to intermediate stress conditions, an influ-
short-time-scale creep rate, which is dominated by the fine ence of the fine c0 on dislocation motion is the only way
c0 . The experimental data can be interpreted with the aid of to rationalize the short-time-scale behaviour observed.
the model predictions. Based upon the work of Brown and Ham [34], it is shown
In the model, the as-delivered sample (A) shows the that particle bypass occurs by dislocation climb in Ni115 at
highest creep strength, followed by the initial unimodal 360 MPa. The coarsening kinetics of Ni115 are far more
sample (E). After the bimodal sample coarsens rapidly to rapid at 800 °C than at 700 °C. The bimodal microstruc-
a unimodal system (50 h with the above coarsening rate), tures exist for thousands of hours at 700 °C but only tens
the inter-particle distance of sample E is slightly larger than of hours at 800 °C, with associated creep lifetimes decreas-
sample A, thus A is predicted to be slightly stronger. This is ing from thousands of hours to hundreds of hours as
J. Coakley et al. / Acta Materialia 59 (2011) 854–863 863

temperature increases from 700 to 800 °C at r = 360 MPa. [7] Meric L, Poubanne P, Cailletaud G. Trans ASME 1991;113:162.
Thus it is the coarsening rate of secondary particles which [8] Ion J, Barbosa A, Ashby M, Dyson BF, McLean M. The modelling of
creep for engineering design-I. Technical Report, National Physical
may be the life-limiting factor for the Ni115 alloy. Laboratory 1986.
Inspection of the complete creep curves of samples A, B [9] Dyson B. Mater Sci Technol 2009;25:213.
and E at 800 °C and 360 MPa suggests that a smaller value [10] Basoalto HC, Sondhi SK, Dyson BF, McLean M. A generic
of kp lowers the creep rate, since under these conditions the microstructure-explicit model of creep in nickel-base superalloys. In:
longer-time-scale creep properties are dominated by the Green KA, et al. editors. Superalloys 2004. Pennsylvania: Minerals,
Metals and Materials Soc; 2004.
primary particles due to the rapid rate of dissolution of [11] Betteridge W, Heslop J. The Nimonic alloys. 3rd ed. London: Edward
the fine particles. This result is expected. The only way to Arnold; 1959.
interpret the opposite results of samples B–D is via the [12] Special Metals Wiggin. Nimonic alloy 115 properties. Publication
influence of the fine particles. According to the mean field number SMC-094 2004.
coarsening model, the fine particles are present for longest [13] Sondhi SK, Dyson BF, McLean M. Acta Mater 2004;52:1761.
[14] Lifshitz IM, Slyozov VV. J Phys Chem Solids 1961;19:35.
in sample D, then C, then B, which appears to result in a [15] Wagner C. Z Elektrochem 1961;65:581.
greater effective strength in sample D than in sample C, [16] Brailsford AD, Wynblatt P. Acta Metall 1979;27:489.
which, in turn, is greater than in sample B. [17] Voorhees PW, Glicksman ME. Acta Metall 1984;32:2001.
The creep model of Dyson is a climb-glide bypass model [18] Voorhees PW, Glicksman ME. Acta Metall 1984;32:2013.
for unimodal distribution which accounts for tertiary creep [19] Marqusee JA, Ross J. J Chem Phys 1984;80:536.
[20] Tokuyama M, Kawasaki K. Physica A 1984;123:386.
within a CDM framework. This model was altered and [21] Coakley J, Basoalto HC, Dye D. Acta Metall 2010;58:4019.
implemented as a quasi-bimodal model. It was not possible [22] Porter A, Ralph B. J Mater Sci 1981;16:707.
to accurately predict the creep time to failure, primarily [23] Kocks UF, Argon AS, Ashby MF. Prog Mater Sci 1975;19:1.
because a number of required microstructural parameters [24] Ansell GS, Weertman J. Trans Metall Soc AIME 1959;215:838.
for Ni115 were unknown and thus were estimated. How- [25] Weertman J. Trans Metall Soc AIME 1960;218:207.
[26] Special Metals Wiggin. Nimonic alloy 115 properties. Publication
ever, the model did give order–magnitude agreement, and number SMC-094 2004.
displayed a number of important trends, giving promise [27] Dyson BF. J Press Vessel Technol 2000;122:281.
to this creep modelling approach for bimodal systems: (i) [28] Pollock TM, Argon AS. Acta Metall Mater 1992;40:1.
at 800 °C, in the short time frame when fine c0 was present, [29] Matan N, Cox D, Carter P, Rist MA, Rae CMF, Reed RC. Acta
all four bimodal samples show the same creep response; (ii) Mater 1999;47:1549.
[30] Caron P, Ohta Y, Nakagawa YG, Khan T. Creep deformation
the model fails to predict the correct order of creep anisotropy in single crystal superalloys. In: Reichman S et al., editors.
responses for the bimodal samples, samples A–D. The Superalloys 1988. Pennsylvania: Minerals, Metals and Materials Soc;
implication is that the strengthening from both the primary 1998.
and secondary particles needs to be considered separately. [31] Sass V, Glatzel U, Feller-Kniepmeier M. Creep anisotropy in the
Furthermore, the local effects need to be considered in the monocrystalline nickel-base superalloy CMSX-4. In: Kissinger R
et al., editors. Superalloys 1996. Pennsylvania: Minerals, Metals and
coarsening model. These conclusions should improve Materials Soc; 1996.
agreement between model and experiment in the future. [32] Friedel J. Dislocations. Oxford: Pergamon; 1964.
[33] Cottrell A. Dislocations and plastic flow in crystals. Oxford: Oxford
Acknowledgements University Press; 1953.
[34] Brown L, Ham R. Dislocation–particle interactions. In: Kelly A,
Nicholson R, editors. Strengthening methods in crystals. Lon-
J.C. would like to acknowledge funding by EPSRC and don: Applied Science Publishers; 1971.
QinetiQ under the Industrial CASE scheme. The material [35] Kruml T, Conforto E, Lo Piccolo B, Caillard D, Martin JL. Acta
used in the study was supplied by QinetiQ plc. The authors Mater 2002;50:5091.
would also like to acknowledge fruitful discussions on [36] Kruml T, Martin JL, Bonneville J. Philos Mag A 2000;80:1545.
applications of the model with Prof. B.F. Dyson, the [37] Hemker KJ, Mills MJ. Philos Mag A 1993;68:305.
[38] Karnthaler HP, Muhlbacher ET, Rentenberger C. Acta Mater
machining and testing performed at Incotest UK, and the 1996;44:547.
testing performed by W. Mitten at QinetiQ plc. [39] Sun J, Lee CS, Lai JKL, Wu JS. Intermetallics 1999;7:1329.
[40] Bontemps-Neveu C. Ph.D. thesis Université Paris Sud.; 1991.
References [41] Dye D, Coakley J, Vorontsov VA, Stone HJ, Rogge RB. Scripta
Mater 2009;61:109.
[1] Rae CMF, Matan N, Reed R. Mater Sci Eng A 2001;300:125. [42] Decker R. Strengthening mechanisms in nickel-base superalloys. New
[2] Rae CMF, Matan N, Cox DC, Rist MA, Reed RC. Metall Mater York: The International Nickel Company; 1970.
Trans A 2000;31:2219. [43] Brown AM, Ashby MF. Correlations for diffusion constants. Acta
[3] Rae CMF, Zhang L. Mater Sci Technol 2009;25:228. Metall 1980;28:1085.
[4] Rae CMF, Reed RC. Acta Mater 2007;55:1067. [44] Lin TL, McLean D. Metal Sci J 1968;2:108.
[5] Walles KFA, Graham A. J Iron Steel Inst 1955;179:105. [45] Muller T. Acta Metall 1971;19:691.
[6] Evans RW, Wilshire B. Mater Sci Technol 1987;3:701. [46] Manonukul A, Dunne FPE, Knowles D. Acta Mater 2002;50:2917.

Potrebbero piacerti anche