Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Journal Pre-proof

Application of an advanced mean-field dislocation creep model to


P91 for calculation of creep curves and time-to-rupture diagrams

Florian Riedlsperger , Bernhard Krenmayr , Gerold Zuderstorfer ,


Bernhard Fercher , Bernd Niederl , Johannes Schmid ,
Bernhard Sonderegger

PII: S2589-1529(20)30177-0
DOI: https://doi.org/10.1016/j.mtla.2020.100760
Reference: MTLA 100760

To appear in: Materialia

Received date: 23 March 2020


Accepted date: 28 May 2020

Please cite this article as: Florian Riedlsperger , Bernhard Krenmayr , Gerold Zuderstorfer ,
Bernhard Fercher , Bernd Niederl , Johannes Schmid , Bernhard Sonderegger , Application of an
advanced mean-field dislocation creep model to P91 for calculation of creep curves and time-to-rupture
diagrams, Materialia (2020), doi: https://doi.org/10.1016/j.mtla.2020.100760

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. 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.

© 2020 Published by Elsevier B.V. on behalf of Acta Materialia Inc.


Application of an advanced mean-field dislocation creep model to P91
for calculation of creep curves and time-to-rupture diagrams
Florian Riedlspergera,*, Bernhard Krenmayra, Gerold Zuderstorfera, Bernhard Ferchera,b,
Bernd Niederla, Johannes Schmida,c, Bernhard Sondereggera,d
* corresponding author; e-mail: florian.riedlsperger@tugraz.at
a
Institute of Materials Science, Joining and Forming (IMAT), Graz University of Technology, 8010 Graz, Austria
b
Present Addr.: Inst. of Internal Combustion Engines a. Thermodynamics, Graz Univ. of Techn., 8010 Graz, Austria
c
Present Address: Institute of Production Engineering, Graz University of Technology, 8010 Graz, Austria
d
Present Address: Materials Testing Institute (MPA), University of Stuttgart, 70569 Stuttgart, Germany

Keywords: creep; steel; simulation; precipitate kinetics; dislocation density


ABSTRACT
This work deals with the development of a comprehensive mean-field dislocation creep

model and its application to the martensitic 9% Cr-steel P91. Microstructural data from

literature, EBSD measurements and results from precipitate kinetic simulations serve as

input for the creep simulation in addition to material- and physical constants. The model

has the capability for calculating creep curves depending on their initial microstructure,

stress and temperature and is thus able to deduce time-to-rupture diagrams. Side-result

is the microstructural evolution during creep exposure (different types of dislocation

densities, subgrains and precipitates). The capability of the model is demonstrated in

P91 within the stress range of 50-110 MPa at 650°C. Simulated results agree well with

experimental data from sources such as NIMS, ASME, ECCC and industrial data.

1 INTRODUCTION
1.1 Application and properties of martensitic steels

9-12 % martensitic Cr-steels (such as P91, P92 or FB2) are state of the art materials for

use in heat-exposed components of thermal power plants. High creep strength and

oxidation resistance [1], sufficient ductility [2], good thermal conductivity and low

thermal expansion [3] plus economical manufacturing (compared to austenitic steels or

Ni-base alloys) [1] [3] qualify them to be included in turbine parts and surroundings,

pipes and heat exchangers [4]. Raising operation temperatures (to 620°C e.g.) and steam

1
pressures (to 300 bar e.g.) help to reduce CO2 emissions from fossil-fired power plants

by maximizing their degree of efficiency [1] [5]. In this work we focus on P91, a

tempered martensitic 9% Cr steel produced by heat-treatment of austenitizing, air

cooling and tempering. The microstructure contains high-angle prior-austenite grain

boundaries together with martensite packets, blocks and laths [6] and low-angle

subgrain-boundaries [7]. All named locations provide spots for precipitates to nucleate

from a supersaturated matrix: Apart from M23C6- which may show significant

coarsening-, small MX carbonitrides contribute to precipitation strengthening [8].

Precipitates represent obstacles for passing mobile dislocations and they may also pin

subgrain boundaries [9] [10], preventing recrystallization [11]. One of the biggest

dangers for P91 creep strength lies in the diffusion of Cr into small MX precipitates (20-

40 nm [12] [7]), causing gradual transformation into the more stable modified Z-phase.

These particles reach 250- 750 nm in size [12], thus contributing less to precipitation

strengthening compared to MX, and are suspected of leading to a breakdown in creep

strength [13]. There are contradicting reports in literature on the role of Laves phase

precipitates (Fe,Cr)2(Mo,W) which form in 9% Cr steels. It has often been argued that

Laves phase depletes the matrix of Mo or W, which might otherwise provide solid

solution strengthening [12] [14] [15]. However, backstress calculations have shown this

effect to be negligible [5]. Coarse Laves phase particles in P92 were suspected

elsewhere of being starting point for creep cavities [16].

1.2 Creep modelling

Creep models can be divided into I.) phenomenological models and II.) (semi-) physical

models. Phenomenological models are not designed to depict microstructural evolution,

but instead provide analytical functions for deformation, creep rate or life time. Larson-

2
Miller [17] and Monkman-Grant [18] are amongst the most wide-spread methods for

determining the material’s life time. Models indicating the strain rate involve McVetty

[19] and Conway [20] for the primary creep regime, Norton [21] and Nadai [22] for the

secondary regime and Rabotnov [23], McHenry [24] or Sandström-Kondyr [25] for

tertiary regime. For completely fitting experimental creep curves of P91, the modified

Graham-Walles model [26], the Bolton Characteristic Strain Model [27] and the MHG

model [28] are recommended by Holdsworth et al. [29]. Application of such models is

restricted to cases, where the simulated material state closely resembles the

experimental material state. Unfortunately, life-times for testing materials such as P92

were shown to scatter between factor 3 and 10 [30]. This may be related to small

variations within the testing conditions, but also to varying initial microstructures.

Hence, when focusing on a deeper microstructural level, (semi-)physical models are

required to fully understand the processes governing creep and to explain the

experimental scatter.

(Semi-)physical models describe a material’s microstructural evolution and its link to

macroscopic properties, such as strength or strain rate ̇. Thus, (semi-)physical models

are ideal for studying variations of initial material states, which are included as starting

condition of the simulation. In addition, the models have the potential to estimate the

remaining life time of an individual sample by considering the current microstructure,

and thus avoiding the above-mentioned scatter of predictions. Due to this higher

flexibility and reliability, we focus on optimizing a (semi-)physical model in this work.

1.3 Applied creep model

Our semi-physical model is based on previous works of the hybrid model by Yadav et

al. [31], Basirat [32] and foremost Ghoniem [33]. In addition to these sources, further

3
contributions are implemented from Kreyca [34], Caillard and Martin [35] [36], the

seminal work of Hirth/ Lothe [37], Lothe [38] and Winning et al. [39]. In accordance

with Ghoniem [33], our model considers the following microstructural elements: mobile

dislocations (density 𝜌m), static dislocations (density 𝜌s), boundary dislocations (density

𝜌b) and subgrains (radius Rsgb). All these microstructural elements evolve over time,

including production, annihilation and transformation terms. Precipitate information is

represented by i classes, indicating radius rp,i and number density NV,p,i of particle type p

(e.g. M23C6 and MX). This notation ensures compatibility to corresponding

experimental data or data stemming from thermodynamic simulation tools such as

MatCalc, TC-PRISMA or PanPrecipitation. For the following discussion of the

individual equations, please note that the original source of each of the equations is

indicated in table 1, any alterations are discussed in the following text.

In our model, the strain rate ̇ is derived from the mobile dislocation density 𝜌m and the

velocity of dislocations, see eq. (1). In addition to original Orowan’s law [40], Yadav

[31] included damage from particle coarsening Dppt, see eq. (12), and subsequently

cavitation damage Dcav, see eq. (11), as treated by Basirat [32]. Please note that in

contrast to the treatment in the cited papers, the glide velocity vg in Orowan’s law has

been extended to an effective velocity veff (also in the follow-up equations eq. (2) & (3)).

veff is calculated as weighted combination of glide and climb processes, considering the

interaction of moving dislocations with precipitates in the matrix, see eq. (7). Rate

equations for 𝜌m, 𝜌s, 𝜌b and Rsgb are depicted in eq. (2) – (5) and are mostly adopted

from [33].

The rate equation for mobile dislocations 𝜌̇ m, eq. (2), contains positive contributions of

a) Frank-Read sources and b) the emission of statics from subgrain boundaries. Note

4
that a mistake from source [33] with respect to this effect was corrected: the static

dislocation density 𝜌s was missing in Ghoniem’s original work in the final eq. (47), the

correct original version can be found as eq. (4) in [33]. The mobile dislocation density

rate is reduced by c) immobilization at subgrains, d1) climb recovery into dipole

configuration and e1) dynamic recovery (spontaneous annihilation) of mobiles and

statics.

The rate equation for static dislocations 𝜌̇ s, eq. (3), includes c) the immobilization of

mobiles at subgrain boundaries, d2) climb recovery of statics in dipole configuration

and e2) dynamic recovery (spontaneous annihilation) of statics.

The rate equation of boundary dislocations 𝜌̇ b, eq. (4), includes f) boundaries produced

from static dislocations (minus a share of statics that annihilate during transformation to

boundaries) and g1) subgrain growth minus the Zener-pinning effect of precipitates at

subgrain boundaries, see also eq. (5).

The growth of the subgrain radius ̇ sgb, eq. (5), is driven by g2) subgrain mobility times

the subgrain pressure, whereas precipitates via Zener pinning effect act against and slow

down the growth process. Nucleation of new subgrains is reflected by term h) (adopted

from [33] and modified to achieve correct units). Term h) is inspired by the modulation

process for energy minimization during cell formation [41]. The subgrain mobility-

extended by the impact of misorientation [39]- fulfils case 1 in our simulation eq.

(5.4.1) and is linked to pipe and lattice diffusion coefficients. The subgrain growth

pressure eq. (5.1) is connected to boundary dislocation density and subgrain interface

energy eq. (5.2).

The mean dislocation spacing within subgrain boundaries hb, as used in eq. (2)-(4), is

defined involving boundary dislocations, statics and the subgrain radius, see eq. (6).

5
The glide velocity vg eq. (8) is adopted from works of hot-deformation of aluminium

[34], in order to avoid the mathematical singularity of the original formulation of Cadek

[42] which occurs in situations where the applied stress is lower than the internal stress

stemming from other dislocations. As a bonus, in the new formulation, the impact of

internal stress and applied stress is decoupled. vg is now associated to internal stress- see

eq. (10), activation energy and activation volume for glide determined from eq. (8.1).

The stress-dependent climb velocity vc eq. (9) plays an essential role for dislocations

surmounting precipitates located in the subgrains and also for the annihilation of

mobiles or static dislocations in dipole configuration. vc is split into contributions

stemming from lattice diffusion vcl and pipe diffusion vcp [33]. As for vg, both

contributions vcl and vcp are reformulated adopting the strategy of Kreyca [34], again

avoiding the mathematical singularities from the original treatment of Ghoniem [33]

and decoupling internal stress and applied stress. vcl in eq. (9.1) is governed by the

lattice diffusion coefficient Ds, the transfer coefficient of defects into jogs ηv, the total

dislocation density ρt (sum of mobiles, statics and boundaries) and Lα (elastic interaction

between dislocation and defects) [33]. vcp in eq.(9.2) is strongly dependent on the pipe

diffusion coefficient Dvp and the activation energy difference between lattice diffusion

Elattice and pipe diffusion Epipe, hence ΔW , see eq. (9.2.1.1) [33] [36] [37] [38]. Epipe is

derived from the self-diffusion energy Esd and the jog-pair energy Ejp, see eq. (9.2.1.1.2)

and eq. (9.2.1.1.2.1). Depending on the question, whether length effect plays a role in

the double-jog motion (linked to the frequency of double-jog nucleation and their effect

on the dislocation segment), the contribution of Ejp to Epipe has to be divided by a factor

for the length effect fL=1 (yes) or fL=2 (no) [36].

6
In summary, following improvements have been included in the model in addition to

the original works of Ghoniem [33], Basirat [32] and Yadav [31]:

(i) Corrected term for emission of statics from subgrain boundaries (faulty in [33]

and neglected in [31]), see eq. (2) term b, containing the density of sources,

(ii) Nucleation of subgrains (faulty in [33] and neglected in [31]); eq. (5) term h

(iii) Addition of misorientation angle to subgrain mobility [39]; eq. 5.4.1 and 5.4.2

(iv) Consideration of climb over precipitates into the effective velocity, see eq. (7)

(v) New model for dislocation glide velocity, eq. (8), with estimate of activation

volume from relaxation tests, see eq. (8.1)

(vi) New model for dislocation climb velocity, eq. (9.1) & (9.2), with a link to jog

energies for the pipe diffusion share, see eq. (9.2.1.1.2), (from [36] [37] [38])

(vii) Replacement of mean precipitate data by representation via size classes, see eq.

(4) term g1 and eq. (5) term g2

(viii) Thermodynamic simulation of the precipitate evolution

(ix) Improvement of precipitate damage parameter compared to [31] by use of a

multiplicative description/ sum-up, see eq. (12)

(x) Explicit indication of original sources of all equations and parameters

Table 1: Listing of the applied formulas in the model. For alterations from the original sources, see previous text.
Creep model equations and numbering of subterms Eq. Ref.
Creep strain rate (modified Orowan equation)
𝜌 [31]
(1)
[40]
( )( )
Mobile dislocation density rate
𝜌 ⁄ 𝜌 ⁄ [33];
⏟ 𝜌 𝜌 ⏟ 𝜌 ⏟ (𝜌 𝜌) 𝜌
⏟ ⏟ (2) term b)
correct.
Static dislocation density rate
𝜌
𝜌 𝜌 ⏟ 𝜌 𝜌
⏟ ⏟ (3) [33]

Boundary dislocation density rate

7
𝜌 ⏞ 𝜌
( ) 𝜌 [ (∑ ) ]
⏟ (4) [33]

Subgrain growth rate


⏞ [33];
[ (∑ ) ] (√𝜌 𝜌 )
⏟ (5) term h)
⏟ correct.

Subgrain growth pressure


𝜌 (5.1) [33]
Interface energy of subgrain boundaries
𝜌 (5.2) [33]
Calculation of shear modulus
( ) (5.3) [43]
Subgrain boundary mobility- case 1, if (∑ )
[33]
(5.4.1)
[39]
Subgrain boundary mobility- case 2, if (∑ )
[33]
(5.4.2)
∑ [39]
Mean dislocation spacing within the subgrain boundary
(6) [33]
(𝜌 𝜌)
Effective velocity

∑ ⏞ (7) new

Dislocation glide velocity


{ } { } { } (8) [34]
Stress change during relaxation test
(8.1) [35]
( ) { }
Corrected applied stress
( ) (8.2) See text
Dislocation climb velocity
(9) [33]
Dislocation climb velocity- lattice diffusion share
{ } { } (9.1) new
[ ( √𝜌 )]
Transfer parameter of defects into jogs
( )
(9.1.1) [33]
( )
Dislocation climb velocity- pipe diffusion share
{ } { } (9.2) new
Distance of core-vacancy diffusion before evaporation into lattice
[37]
√ { } (9.2.1)
p. 572
Activation energy difference between lattice and pipe diffusion
[37]
(9.2.1.1)
p. 572
Activation energy for lattice diffusion
(9.2.1.1.1) [38]

8
Activation energy for pipe diffusion
(9.2.1.1.2) [36]
Jog pair energy

[37]
(9.2.1.1.2.1)
( ) ⏟ ( ) p. 263

Internal stress [32]


√𝜌 𝜌 (10) [44]
Damage parameter for cavities
̇ ̇ (11) [32]
Overall damage caused by precipitate coarsening

∏( ) (12) new

Damage rate caused by precipitate coarsening


̇ ( ) (12.1) [31]
Normalized Ostwald ripening parameter
( )
[ ] (12.1.1) [32]

Ostwald ripening parameter


(12.1.1.1) [45]
Calculation of boundary dislocation density from EBSD data
𝜌 (13) [46]

2 MODEL SETUP
2.0 General strategy
In a nutshell, our model aims at predicting creep curves and microstructural evolution,

depending on the starting microstructure of the material. Side result of the simulation is

extrapolating the creep behaviour from short-term experiments to virtual long-term tests

by using the same starting microstructure and model parameters, and just altering the

system stress. A number of the required input parameters must first be adapted to match

one short-term creep curve, since they are difficult to measure otherwise (4 adaptions +

2 parameters, where a rough estimate was available). The general strategy on applying

our model thus consists of following individual steps, which are now discussed in

details:

Step 1: Selecting a material (here: P91)

Step 2: Literature study on the model input parameters for the material group

9
Step 3: Select a specific material badge with sufficient literature data or actual samples.

Step 4: Carry out one instrumented creep test of 103 to 104 hours duration.

Step 5: Investigate as-received microstructure

Step 6: Simulation, measurement or literature study on precipitate evolution during

creep exposure (in this work: thermodynamic simulations using MatCalc software)

Step 7: Simulation of one master creep curve. Adapting the remaining 4+2 input

parameters to match results of the instrumented creep test

Step 8: All the parameters are now fixed. Simulate multiple creep curves for varying

stresses. Verify calculated microstructural evolution. Construct time-to-rupture diagram.

2.1 Selecting a material

We chose P91 due to the availability of substantial quantity of literature data. As

indicated in the next subsection, much of this data can also be used in general for 9-12%

Cr steels. Thus, P91 appears to be the best choice for evaluating the capability of the

model.

2.2 Literature study on the model input parameters for the material group

A number of input parameters were available from literature sources, partly for Mo-

containing martensitic steels, partly for 9-12% Cr steels or more general for Fe-bcc or as

fundamental physical constant. Table 2 gives an overview on the acquired data,

indicating the symbol, the value, the range of applicability and the literature source.

Table 2: Values of the variables according to the literature study


Input Value Material (Group) Source
ag 2.866·10-10 m Fe - bcc [47]
0.02 9Cr-1Mo steel at 650°C [32]
b 2.48·10-10 m Fe - bcc [48]
cs 0.3 9Cr-1Mo steel at 650°C [32]
5b Edge dislocations [32] [49] [50]
Ds 2·10-19 m2/s Fe in Fe-1.5Mo at 650°C [51]
3.94-5.04·10-19 J 9-12 % Cr steels at 650°C [36] [37] [38]
2·10-4 Ferritic steels [33]
E 163 GPa (650°C) 9-12 % Cr steels at 650°C [52]
lM23C6 4 M23C6 in 9-12 % Cr steels [32]

10
lMX 6 MX in 9-12 % Cr steels [31]
kB 1.38065·10-23 J/K Fundamental constant [53]
M 3 Crystals (approximation) [32]
Q 4.01·10-19 J HT-9 at 550°C [33]
3°=0.0524 rad 9-12 % Cr steels [54] [55] [56]
Vr 43 P91 at 625°C [57] [35]
Vr 38-44 P92 at 650°C [58] [35]
0.317 (650°C) 9-12 % Cr steels [52]
0.034 9-12 % Cr steels [59]
2.3 Select a specific material badge or actual samples for investigations

We tested the 9% Cr steel P91, produced with the chemical composition according to

[60] (see table 3) and standard heat treatment data (see table 4).

2.4 Instrumented creep test

The instrumented creep test was carried out at 650°C and a stress of 70 MPa. Two rods

were fixed to the two collars of the creep sample. The rods transferred the position

changes to two extensometers located outside the creep furnace. A Magnescale LT20A

device then evaluated the signals of the two collar positions relative to the initial

location. The data were stored continuously on a connected PC system. The recorded

deformation (rate) was documented up to the fracture time of 8740 h.

2.5 As-received microstructure

Information on the as-received (=starting) microstructure were taken from three types of

sources: TEM data from literature on mobile and static dislocation density,

experimental EBSD data on subgrain sizes- which are used in addition for calculating

boundary dislocation densities- and thermodynamic simulation for precipitate data (see

section 2.6).

The starting value for mobile dislocation density of ρm,0=4.5·1014 m-2 was taken from

the TEM analysis of Panait et al. [7]. According to Basirat [32], the static dislocation

density ρs,0 is assumed 10 % of the mobile density, hence 4.5·1013 m-2.

11
EBSD samples were cut, ground, polished and vibration polished with a ATM Saphier

Vibro. EBSD investigations were carried out with a Tescan Mira 3-XM field-emission

SEM with AMATEK Octane Super EBSD Detector. The field of view was 70 x 70 µm

with a step size of 50 nm. Analysis was performed using EDAX TSL OIM Analysis 8

software. Subgrain boundaries were defined as having a misorientation angle of 1.5°-

15°. Prior austenite grain size (PAGS) was found by light optical microscopy (LOM)

and mean linear intercept method applied to as-received samples which were edged by a

Lichtenegger/ Bloech reagent; see table 4.

2.6 Precipitate evolution


In order to gather the most comprehensive precipitate data in as-received state as well as

during creep, we decided to conduct a thermodynamic precipitate kinetic simulation

with software package MatCalc 6.02 (based on the SFFK model [61]), using

thermodynamic database “mc_fe_v2.060.tdb” and mobility database

“mc_fe_v2.012.ddb”.

In addition to chemical composition (see table 3) and heat treatment information (see

table 4), MatCalc simulations also demand data on prior austenite grain sizes (PAGS),

subgrain sizes and dislocation densities (numbers and sources see table 4) to calculate

effective diffusion coefficients and number of nucleation sites for precipitates.

Table 3: Chemical composition of the used P91


Element C Si Mn Cr Mo Ni Al V N Nb Ref
%wt 0.10 0.35 0.45 8.75 0.95 0.35 0.02 0.22 0.05 0.07 [60]
Table 4: Input parameters for the MatCalc precipitate kinetic simulation
Input MatCalc Value Source
Normalizing 2h @ 1020°C Test Case
Heat
Tempering 2h @ 760°C Test Case
Treatm.
Service 105 h @ 650°C Test Case
PAGS 25 µm LOM
Subgrain Radius 0.4 µm EBSD
𝜌 Martensite 4.5·1014 m-2 [7]
𝜌 Austenite 1·1011 m-2 [62]
1st loop: eq. (14)
Recovery Rate of 𝜌 [63], then coupling
2nd loop: eq. (2)

12
Martensite Start Temp. 400°C [64]
Reaustenitization Temp. 833°C [64]
The first approximation of the time-evolution of mobile dislocation density during

service is taken from a non-linear recovery function adopted from Riedlsperger [65] and

El-Azim et al. [63] as indicated by eq. (14).

( ) (14)

where a=2.4·1016 m-2, k=7000 h, b=0.57 and the time t is inserted in the unit h. The

result of the corresponding precipitate evolution is used for a preliminary simulation of

the master creep curve, which in return, produces an improved 𝜌m(t) stemming from our

creep model. This improved 𝜌m(t) is then used for a 2nd loop of precipitate kinetic

simulation and also the subsequent final master creep curve.

The precipitate types of AlN, Cr2N, Fe3C, Laves phase, M7C3, M23C6, NbC, VN and

mod. Z-phase were considered in the simulation. The nucleation sites are indicated in

table 5.

Table 5: Nucleation sites for precipitates in P91


AlN Cr2N Fe3C Laves M7C3 M23C6 NbC VN Z
aust d d - - - - g, d g VNi
mart d d s s s g, s g, d, s g, d, s VNi
“mart”=Martensitic Matrix; “aust”=Austenitic Matrix; “g”=Grain Boundaries,
“d”=Dislocations, “s”=Subgrain Boundaries; “VNi”= Precipitates as Nucleation Sites
2.7 Simulation of one master creep curve
The stress and temperature of the simulated master creep curve were set according to

the instrumented creep test at 650°C and 70 MPa. Input parameters on the creep

simulation are stated in the previous sections. Some of the input parameters, however,

are not available from either the literature or experiment; these parameters have been set

in order to achieve the best agreement between experimental and simulated creep curve.

Most of these parameters have a physical interpretation (and can, thus, in principle be

measured), but no literature data were available for: the stress-independent pre-factor a1

for the velocity of undisturbed dislocation glide, the density of sources , the Holt

13
constant Kc and the pipe diffusion coefficient Dvp. Kc has been depicted for pure metals

[66]. It remains unclear, however, to what extent the result is also applicable to complex

alloys. Literature data on pipe diffusion coefficient Dvp show wide scatter in the range of

5-1000 depending on temperature level and experimental technique [67] [68]. In

addition, Dvp appears to be strongly dependent on the matrix composition [69]. For both

constants, Kc and Dvp, no data are available for P91. For the activation energy of climb

by pipe diffusion Epipe, only a range was available, allowing an adjustment of

approximately 20-25 %.

The apparent activation volume for thermally activated glide Vr [70] was determined

from literature relaxation data on P91 and P92. More precisely, a method from [35] was

applied to determine Vr in eq. (8.1) by fitting a relaxation curve of P91 at 625°C [57].

The obtained Vr=5·10-28 m3 equals 43 times the atomic volume Ω. Since the literature

sources [35] [70] indicated the temperature dependence of this parameter, the given

number for 625°C was taken as starting point for further optimization at 650°C.

Applying the same method to two relaxation curves of P92 at 650°C [58], yielded 38Ω

and 44Ω, confirming the above.

The damage parameter A is of phenomenological nature and thus cannot be measured

by an independent experiment; A was adjusted to the tertiary creep regime of the creep

curve.

2.8 Multiple creep simulations plus TTR diagram

It is essential to state at this point that the complete set of parameters, which were

adjusted to the 650°C/70 MPa creep curve, is adopted and applied to every other creep

simulation within the investigated stress range, without any additional alterations. This

statement is important in order to point out the ability of the model to extrapolate from

14
one stress state to another. The only parameter which was varied in order to produce a

time-to-rupture (TTR) diagram is the nominal stress, which was set from 50-110 MPa in

steps of 10 MPa, producing 7 individual creep curves. In addition, each of the individual

creep curves is accompanied by simulation of the according microstructure.

Since the creep model relies on uniform stress within the material, and necking in P91

creep samples within the investigated stresses and temperatures is expected at a nominal

strain of 4-8% [71], we take a deformation of 6% as end of the simulation. This

criterion neglects less than 10% of the remaining sample lifetime [71] [72]. The second

criterion for fracture is reaching 100 % of either precipitate damage Dppt or cavitation

damage Dcav.

3 RESULTS
3.1 EBSD investigations

Figure 1: EBSD data of (a) as received status; (b) 8740 h crept (650°C/70 MPa); IPF-map

EBSD measurements of as-received state revealed a mean subgrain radius of Rsgb,0=0.4

µm (see figure 1a). Together with a subgrain elongation factor of 4 [46] and a mean

misorientation angle for low-angle boundaries in tempered martensitic steels of app. 3°

[54] [55] [56], a boundary dislocation density of ρb,0=5.9·1014 m-2 was calculated, see

eq. (13). In this calculation, twice the subgrain radius was taken as subgrain width

(w=2·Rsgb,0) [73] [74], which is needed when assuming a cuboid-shaped subgrain

15
geometry [46]. This result is supported by literature data of ρb,0=6.2·1014 m-2 [54]. Both

Rsgb,0 and ρb,0 served as input parameters for the creep simulation. EBSD results of the

crept sample (8740h @ 70MPa/ 650°C; see figure 1b) were Rsgb=0.7 µm and

ρb=3.4·1014 m-2. This helped to verify our simulation (see amid a wider frame in section

4.3).

3.2 Precipitate kinetic simulation

This section presents the final result of the precipitate kinetic simulations after

implementing the 2 nd loop of calculating 𝜌m(t) (see model setup). Following precipitate

types turn out throughout creep exposure: M23C6, VN, NbC, AlN, Laves and mod. Z-

phase. Figure 2 comprises the evolution of the particles during creep, with respect to

mean diameters (a) and phase fractions (b). Labelling of the particles includes the

particle types and their nucleation sites (some were negligible with <0.01 % phase

fraction and are not shown for simplification, although stated in table 5 and

implemented in the calculation).

Figure 2: Results of the MatCalc precipitate kinetic simulation of P91; (a) mean diameter, (b) phase fraction

Detailed numbers on particle diameters, phase fractions and number densities (related to

the former and latter) in as-received state and after 103, 104 and 105 h of creep are noted

in tables A2 and A3 of the appendix. The complete precipitate evolution is now used as

one of the input data for the creep simulation, see next section.

16
3.3 Master creep curve

The master creep curve was calculated according to the available instrumented creep

test at 70 MPa/ 650°C. The input parameters were material data from literature (table 2),

results of MatCalc (section 3.2) and EBSD experiments on the as-received sample

(section 3.1). The remaining parameters to be optimized for a best match between

measured and simulated master creep curve as well as microstructural data are a1, β, A,

Dvp, Kc and Vr. a1 was mostly linked to the slope of the creep curve in the secondary

creep regime. The choice of β mainly impacted the shape of the creep curve within the

primary creep regime, whereas the damage parameter A controlled the last third of the

creep curve.

The obtained interaction volumes from relaxation tests in literature (Vr=43·Ω resp.

Vr=38·Ω to 44·Ω) were taken as starting point for further optimization. The Holt

constant Kc was found by restricting subgrain coarsening after end of creep life at 70

MPa to Rsgb=1.1 µm (from originally Rsgb=1.4 µm without use of a subgrain nucleation

term).

Table 6 comprises the final set of remaining parameters for the master creep curve.
Table 6: Final set of remaining simulation parameters
Parameter Number Unit Parameter Number Unit
a1 3.90 [m/s] Dvp 4.75·10-19 [m2/s]
β 0.0375 [-] Kc 2.1 [-]
A 560 [-] Vr 35·Ω [m³]
Figure 3a shows a comparison between simulated and measured creep curve at 70 MPa,

showing good agreement in all three creep regimes plus the final lifetime of the

material. Figure 3b compares modelled and measured creep strain rate. Figure 4a

depicts the evolution of all dislocation types and subgrains for the modelled creep

curve.

Boundary dislocation density is observed to shortly increase in primary creep regime

17
from 5.9 to 6.4·1014 m-2 due to intake from mobiles and statics, followed by a

continuous decrease down to 2.5·1014 m-2 due to subgrain coarsening. The subgrains

coarsen from 0.4 µm in as-received condition to Rsgb=1.1 µm after creep. Mobile

dislocations annihilate to a minimum of 8.2·1012 m-2 at 900 h. After this minimum, ρm

moderately increases to 1.4·1013 m-2 until the end of creep life. Recovery of statics takes

place rapidly in the first 500 h of creep, decreasing from 4.5·1013 m-2 to 2.3·1010 m-2.

Reduction of ρs then continues more slowly down to 1.4·1010 m-2 after end of creep life

at 8749 h.

Figure 3: (a) Result of the simulated creep behaviour of P91 for 70 MPa/ 650°C compared to the
experiment; (b) comparison between the measured and the simulated creep strain rate

Figure 4: (a) Simulated evolution of subgrain radius and dislocation densities at 70 MPa/ 650°C;
(b) simulated evolution of climb and glide velocity at 70 MPa/ 650°C

Figure 4b indicates the evolution of glide (vg) and climb velocity (vc) of mobile

dislocations. vg=7.8·10-13 m/s and vc=1.6·10-13 m/s have a ratio of 5:1, making vg clearly

dominate effective velocity veff=7.7·10-13 m/s. vc can be split into a lattice (vcl) and a

18
pipe diffusion part (vcp). vcl contributes 81 % and vcp only 19 % to the overall climb

velocity.

3.4 Extrapolation and time-to-rupture-diagram

The TTR diagram (figure 5b) was constructed using 7 individual simulated creep curves

between 50-110 MPa (figure 5a), obeying the 6 % criterion (section 2.8). All creep

curves share the same input parameters as the master creep curve: material parameters

(table 2), results of the MatCalc simulation (section 3.2), EBSD data on as-received

sample (section 3.1) and the parameter set for the master creep curve (section 3.3, table

6).

The accompanying microstructural results for each applied stress can be seen in table 7.

Figure 5: (a) All simulated creep curves for 50-110 MPa; (b) simulated time-to-rupture diagram vs. literature
Table 7: Selected rupture times tR and microstructural results for σapp=50-110 MPa @ 650°C
Unit Start 50 MPa 60 MPa 70 MPa 80 MPa 90 MPa 100 MPa 110 MPa
tR h - 49641 22037 8749 3121 1079 415 187
ρm m-2 4.5·1014 2.0·1012 5.0·1012 1.4·1013 4.1·1013 1.1·1014 2.4·1014 4.8·1014
ρs m-2 4.5·1013 7.8·108 3.1·109 1.4·1010 5.9·1010 2.1·1011 5.6·1011 1.2·1012
ρb m-2 5.9·1014 9.7·1013 1.5·1014 2.5·1014 4.3·1014 6.8·1014 9.1·1014 1.1·1015
Rsgb µm 0.40 2.40 1.66 1.10 0.74 0.53 0.43 0.40

Notably, for smaller σapp,0 and longer creep life, annihilation of mobile dislocations 𝜌

is more pronounced than for higher stresses. In the range of σapp,0 from 50 to 100 MPa,

annihilation of mobiles predominates over their production. Only at 110 MPa, more

mobiles are produced than annihilated, making ρm exceed the start value ρm,0. The higher

19
σapp,0, the more static dislocations ρs are predicted to exist, but reduction compared to

the starting value ρs,0 always takes place. Boundary dislocation density ρb after creep

turns out to exceed the starting value ρb,0 for σapp,0≥87.5 MPa. The smaller σapp,0, the

more pronounced is subgrain coarsening, e.g. culminating in Rsgb=2.4 µm for 50 MPa.

4 DISCUSSION
4.1 Creep curve
The agreement between simulated and measured creep curve at 70 MPa is very good:

The modelled primary creep regime exactly meets the measured creep data. The

simulated minimum creep rate of 2·10-6 h-1 agrees well with the measurement, although

it is reached earlier (after 900 h) than in reality (after 2000 h). In the secondary creep

regime up to 3000 h, simulated and measured creep curve have a nearly identical shape.

After 3000 h, the curvature in simulation exceeds the one measured. The error,

however, is never higher than 400 h in time or 0.3 % in strain. In the tertiary creep

regime, modelled and measured data start to closely approach each other once more

after 8000 h, leading to excellent agreement with respect to rupture times. It was

observed that the overall shape of creep curves and the ratios of primary, secondary and

tertiary creep become more realistic with higher applied stress. We assume that at lower

applied stress, damage in the model might be overrepresented and leads to exaggerated

curvature of the simulated creep curves.

4.2 TTR diagram

After variation of σapp,0 between 50 and 110 MPa, all modelled rupture times show

excellent agreement with data from ECCC [75] (except for 50 MPa), but lie below data

from NIMS [76]. ASME data [77] predict longer lives at high σapp,0 and shorter lives at

low σapp,0 than simulated. ECCC points are standardized average data obtained from an

ISO CRD procedure and a Manson-Haferd (MH) approach applied to 2195

20
experimental points (from many different heats) [75]. NIMS data are mean rupture

times of heat MgC [76]. ASME data are Orr-Sherby-Dorn (OSD) fits based on

experimental results of 14 heats. The deviation of modelled lifetime from ECCC data is

within the ±19 % dispersion criterion from EN ISO 204- except for 50 MPa. The range

of this zone is symbolized by a grey area in figure 3a. At low σapp,0 (long testing times),

it should be noted that ECCC considered few data points in the analysis. NIMS heat

MgC clearly depicts a drop of creep strength as observed in the simulation. The reason

for the TTR drop in our simulation is the big subgrain size caused by reduced Zener

pinning due to coarsened precipitates. The drop in reality might also be related to

diffusion creep [78].

4.3 Microstructural evolution

The simulated microstructural evolution can be partly confirmed by literature and our

measurements. STEM analysis of a 12 % Cr steel (exposed 3650h@ 80MPa/ 650°C)

revealed 𝜌 m-2 [79]. From another 12 % Cr sample (2875h@ 80MPa/

650°C), 𝜌 m-2 were obtained [79]. TEM measurements of a modified

9Cr1Mo steel (5556h @ 98 MPa/ 650°C) detected 𝜌 =3.5·1013 m-2. In a Ta-containing

12 % Cr steel (5793h @ 80 MPa/ 650°C), 𝜌 =2·1013 m-2 was measured by TEM [55].

The overall image of measurements confirms our simulated ρm=1.4·1013 m-2 at 70 MPa

(tR=8749 h) up to ρm=4.1·1013 m-2 at 80 MPa (tR=3121 h). At 50-60 MPa, our modelled

𝜌 appears to be very low. TEM measurements of long-term aged X20 (>5·104h @

550°C) have demonstrated, though, that the order of 1012 m-2 may be reached in extreme

conditions [80]. The phenomenon of hot-temperature work-hardening for high stresses

was documented both by Ghoniem [33] and by Orlova/ Cadek [81], supporting our

simulation result that 𝜌 increases over start value 𝜌 for σapp,0=110 MPa.

21
Subgrain sizes (diameters) for long-term creep of P91 at 600°C were found to lie

between 1 µm [7] and 1.5 µm [82] (both TEM). At 650°C, up to 2 µm (P91; 9000h@

60MPa) [83], 1.76 µm (10 % Cr steel; 100h@150MPa) [84] and 1.50 µm (12 % Cr steel

with Ta; 5793h@80MPa) [55] of subgrain size were detected by TEM. In [85] (P911;

4743h@120MPa/650°C), subgrain length (4.4 µm) is distinguished from width (1.3

µm).

In this work, EBSD evaluation of the ruptured sample’s gauge section (8740h @ 70

MPa/ 650°C) yielded a mean subgrain diameter of 1.4 µm (see figure 1b), confirming

most of the above stated findings from literature.

Thus, our simulated subgrain sizes at σapp,0≤60 MPa might be too big compared with

experiments (only subgrain length values in such range exist [85]), whereas for ≥70

MPa, our modelled 0.8 to 2.2 µm in diameter show good agreement with above stated

data.

[86] and [87] stated an experimentally calibrated equation to calculate a steady-state or

stationary subgrain size for each applied stress to which the initial subgrain size finally

converges. The following equation from [87] is used for a comparison to simulated

data:

(15)

dsgb,∞ ranges from 1.8 µm at 110 MPa to 3.9 µm at 50 MPa, showing excellent

agreement for simulated subgrain sizes at low σapp,0 and satisfactory agreement for high

σapp,0.

Simulated boundary dislocation density at the end of creep life also becomes higher

than the starting value for σapp,0≥87.5 MPa. This can be explained by a high number of

statics transformed into boundaries and by nucleation of new subgrains- see eq. (4) f)

22
and eq. (5) h). The simulated dynamic (strain-induced) recovery, going hand in hand

with subgrain formation and resulting in more boundary dislocations, is consistent with

microstructural analysis on short-term crept, highly stress-exposed 9% Cr-steels at

650°C [88] [89]. Parallels to hot working or CDRX are visible [90]. The phenomenon

of ρb increasing over ρb,0 was also observed in simulations employing a similar creep

model concept (P92; 90-110 MPa/ 650°C) [31]. For σapp,0=50-80 MPa, the decrease of

modelled ρb fits to the result of EBSD evaluation (3.4·1014 m-2) and is in a similar range

to [54], where ρb reduced from 6.1·1014 m-2 to 1·1014 m-2 (P92; 1271h @ 118 MPa/

650°C). In [55], ρb decreased from 5.2·1014 m-2 to 1.5·1014 m-2 (12 % Cr steel with Ta;

5793h @ 80 MPa/ 650°C).

4.4 Parameters

The pre-factor a1 of the glide velocity was selected so that glide and climb velocity were

balanced appropriately to meet the experimental creep curve. The simulation revealed a

vg:vc ratio of 5:1 for the master creep curve (@70 MPa/ 650°C), making vg contribute

significantly more to veff. This means that creep here is predominantly directed by glide.

The primary creep regime was improved by setting β (the density of sources) to a level

so that term b) became 30 to 60 % of term a) in eq. (2), which made additional mobile

dislocations nucleate from statics and partly compensated annihilation and

transformation of mobiles. This became only possible after correcting a mistake: ρs was

added to term b) in eq. (2), which was missing in eq. (47) of Ghoniem’s original work

[33].

Once primary and secondary creep regimes had met the experimental curve, tertiary

creep was approximated by adapting the damage parameter A in eq. (11) for cavitation

damage.

23
Precipitate damage in the model is only governed by MX and M23C6 precipitates of

which coarsening behaviour and number density evolution are imported step-wise from

MatCalc. Considering only MX and M23C6 seems reasonable since they have the

highest phase fractions. The use of different coarsening parameters lMX=6 [31] and

lM23C6=4 [32] in literature reflects different responsible diffusion mechanisms for the

coarsening process (volume diffusion for M23C6, but pipe diffusion for MX) [8].

The Holt constant Kc for nucleation of new subgrains (suggested to be around 10 in [33]

and dating back to dislocation cell formation in pure metals [41]) was changed to a

value of 2.1 in order to reduce subgrain coarsening, after term h) had been modified in

eq. (5).

For estimating the climb velocity driven by pipe diffusion vcp (in contrast to the climb

velocity due to lattice diffusion vcl), the activation energy difference between lattice and

pipe diffusion ΔW turned out to be essential. The activation energy for pipe diffusion

Epipe strongly varies, depending on the definition of the jog-pair energy Ejp. For Ejp, for

Epipe and for ΔW upper and lower limits were calculated, as can be seen in table 8.

Table 8: Calculation of activation energy differences for different approaches


Burgers
Interaction Case ΔW Epipe Elattice Ejp
Vector
fL=1 0.707 4.49
yes 1.65
fL=2 -0.119 3.66
b1/2<111> 3.78 ·10-19 J
fL=1 1.26 5.04
no 2.20
fL=2 0.156 3.94
It was decided to neglect the interaction term in eq. (9.2.1.1.2.1) and to use the upper

threshold Epipe=5.04·10-19 J which led to ΔW=1.26·10-19 J, since vcp in this case returned

the same order of magnitude as vcl (around 10-14 m/s), whereas for smaller values of

ΔW, vcp became unrealistically big compared to vcl.

Choice of the Burgers vector b not only had an impact on the mentioned ΔW (and

consequently on vcp), but also influenced interfacial energy (5.2) and mobility Msgb

24
(5.4.1) of subgrain boundaries. b is even contained in the internal stress σi (10) and

therefore, may modify vg (8). Based on a study of Cheng et al. [48] about (in)stability of

different Burgers vectors in bcc determined by molecular dynamics, b1/2<111> with a size

of 2.48 Å was preferred over b<100>=2.86 Å.

The equation for internal stress (10) was adopted from Basirat [32] and is strongly

linked to the dislocation interaction parameter α. α is usually thought to be a matrix of

interaction (rather than a single numerical value) which has been solved by discrete

dislocation simulations for fcc [91] and for bcc [92]. Unfortunately, issues still remain

concerning high temperature conditions and heterogeneous dislocation arrangements

that might occur during long-term deformation [44]. Therefore, we confine ourselves in

this work to take α as a single numerical value from Basirat [32] with a suggested value

of α=0.02.

The activation volume for dislocation glide, Vr, was correlated with stress relaxation

behaviour by applying logarithmic relaxation law [35] in eq. (8.1) to data from [57].

Evaluation from the relaxation tests reveal Vr=42Ω for P91 at 625°C [57]; the authors of

the work, however, note that the result is temperature dependent [35] [70]. Evaluation of

data from Khayatzadeh et al. suggests Vr=38Ω to 44Ω for P92 at 650°C, indicating

additional scatter stemming from the experimental evaluation [58]. These two findings

gave sufficient motivation for a further optimization of Vr, leading to a final result of

Vr=35Ω. Kiener et al. found Vr for W and Cr by micro-compression tests carried out

below the knee temperature Tc and claim the result to be valid for any bcc metal [93].

They verified their Vr result by a calculation based on kink/ bulge theory [94]. Apart

from the lack of high temperature data (and the failure of equations to calculate Vr for

T>Tc), it turned out that Vr in the investigated low-temperature range scattered by 20 %

25
in measurements [93]. Vr even scattered by a factor of two in calculation from the kink/

bulge theory [94], related to uncertainties with respect to active slip systems [95]. In this

light, 10-20 % adjustment of Vr to optimize our simulation seem to be more than

justified.

4.5 Overall performance


The simulated creep curve shapes are realistic and sufficiently predict primary,

secondary and tertiary creep zones. The modelled TTR diagram shows excellent

agreement with reference data. The microstructural evolution is satisfactory and

meaningful compared to TEM data from literature and our EBSD results. Introducing

faster algorithms drastically reduced the calculation time per creep curve. Firstly, this

enables us to include more extrapolation points to the TTR. Secondly, this opens up the

possibility for backward engineering and automatized parameter adjustments. With this

strategy, one might be able to find the ideal starting microstructure to optimize creep

life of P91 or other 9% Cr steels.

5 CONCLUSION
By applying an advanced mean-field dislocation creep model to P91, it was

demonstrated that from one experimental creep curve (here 70 MPa at 650°C), a

complete TTR diagram can be deduced with the same set of input parameters. For each

applied stress, a corresponding creep curve was modelled, being accompanied by the

microstructural evolution and serving as data point for TTR. Creep curve shape, rupture

times and microstructural changes were shown to agree well with measurements and

literature data. Creep curve and TTR calculation was successfully carried out from 50-

110 MPa.

6 OUTLOOK

Application of the shown creep model is possible for any temperature T where the same

26
creep mechanisms apply (e.g. 600°C). The following changes are required: a) data on

precipitate evolution have to be re-evaluated at T, b) Vr has to be modified according to

relaxation data at T and c) the diffusion coefficients have to be adapted to T.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

7 ACKNOWLEDGEMENTS

All authors thank Dr. Komai & Mr. Arisue from Mitsubishi Heavy Industries Japan for

discussions. B. Sonderegger, F. Riedlsperger & G. Zuderstorfer gratefully acknowledge

funding from Austrian Science Fund (FWF) within project “Software Development on

Dislocation Creep in Alloys” (P-31374). F. Riedlsperger appreciates Ricardo Buzolin

and Mateusz Skalon for discussions.

27
8 APPENDIX
8.1 List of variables
Table A1: Listing of the used variables
Var. Description Unit Source
a1 Parameter for glide velocity [m/s] Fit
ag Size of unit cell [m] [47]
aj Jog height- assumed as b [m] [37]
A Material specific constant [-] Fit
Asgb Subgrain surface [m2] [46]
Dislocation interaction factor [-] [32]
b Burgers vector ½ <111> for bcc [m] [48]
Parameter for density of sources [-] Fit
cs Weighting factor for statics in the internal stress [-] [32]
cr Time constant to determine apparent activation volume [s] [35]
Dislocation annihilation length [m] [32] [49] [96]
dsgb,∞ Stationary/ steady-state subgrain size (=diameter) [m] Eq. (15); [87]
Dcav Cavitation damage parameter [-] Eq. (11)
Dppt Precipitate damage parameter [-] Eq. (12)
Ds Lattice diffusion coefficient [m2/s] [51]
Dvp Pipe diffusion coefficient [m2/s] Fit
E Young’s modulus [Pa] [52]
Ef Vacancy formation energy [J] Eq. (9.2.1.1.1); [38]
Ejp Jog pair energy [J] Eq. (9.2.1.1.2.1); [37]
Elattice Activation energy for climb by lattice diffusion [J] Eq. (9.2.1.1.1); [38]
Em Vacancy migration energy [J] Eq. (9.2.1.1.1); [38]
Epipe Activation energy for climb by pipe diffusion [J] Eq. (9.2.1.1.2); [37]
Esd (Vacancy-) Self-diffusion energy [J] Eq. (9.2.1.1.1); [38]
Creep strain [-] Eq. (1)
̇ Creep strain rate [1/s] Eq. (1)
Factor for length effect in the double-jog motion; fL=1: length-
fL [-] [36]
effect regime; fL=2: jog-collision (no length effect)
Transfer coefficient for defects into jogs [-] [33]
G Shear modulus [Pa] Eq. (5.3)
Subgrain boundary (interface) energy [J/m2] Eq. (5.2); [33]
hb Mean dislocation spacing within the subgrain wall [m] Eq. (6); [33]
kB Boltzmann constant [J/K] [53]
kd Ostwald ripening parameter [m3/s] Eq. (12.1.1.1)
kp Ostwald ripening parameter normalized to initial particle radius [s-1] Eq. (12.1.1)
Kc Holt constant [-] [33]
l Material specific parameter for precipitate coarsening [-] [31] [32]
Parameter for elastic interactions between dislocation/defects [m] Eq. (9.1.1); [33]
Jog width- assumed as b [m] [37]
Diffusion path of core-vacancy before evaporation into lattice [m] Eq. (9.2.1); [33]
M Taylor factor [-] [32]
[m/Pa.s]= Eq. (5.4.1); [33]
Msgb Subgrain boundary mobility
[m3/Ns] Eq. (5.4.2); [33]
Nv,i,0 Precipitate number density starting value [m-3] MatCalc
Nv,i Precipitate number density [m-3] MatCalc
Poisson’s ratio [-] [52]
Psgb Subgrain growth pressure [Pa] Eq. (5.1); [33]
Q Activation energy for dislocation glide [J] [33]
rp,i,0 Precipitate mean radius starting value [m] MatCalc
rp,i Precipitate mean radius [m] MatCalc
Rsgb,0 Subgrain radius starting value [m] EBSD

28
Rsgb Subgrain radius [m] Eq. (5)
𝜌 Boundary dislocation density starting value [m-2] Eq. (13)
𝜌 Boundary dislocation density [m-2] Eq. (4)
𝜌 Mobile dislocation density starting value [m-2] [7]
𝜌 Mobile dislocation density [m-2] Eq. (2)
𝜌 Static dislocation density starting value [m-2] See text
𝜌 Static dislocation density [m-2] Eq. (3)
𝜌 Total dislocation density- sum of mobiles, statics, boundaries [m-2] [33]
Applied initial stress [Pa] User Input
Corrected applied stress [Pa] Eq. (8.2)
Internal stress [Pa] Eq. (10); [32]
( ) Stress change during relaxation test [Pa] Eq. (8.1); [35]
t Time [s] See text
tR Rupture time [s] See text
T Temperature [K] Input
Mean misorientation of subgrains [rad] [46]
vc Climb velocity [m/s] Eq. (9)
vcl Lattice diffusion share of vc [m/s] Eq. (9.1)
vcp Pipe diffusion share of vc [m/s] Eq. (9.2)
veff Effective velocity [m/s] Eq. (7)
vg Glide velocity [m/s] Eq. (8)
Vr Apparent activation volume for thermally activated glide [m3] Eq. (8.1); [35]
Vsgb Subgrain volume [m3] [46]
w Subgrain width [m] [73] [74]
Activation energy difference betw. pipe and lattice diffusion [J] Eq. (9.2.1.1); [37]
Atomic volume (unit cell size divided by 2 atoms in bcc) [m3] See text
Fraction of statics annihilating when transforming to boundaries [-] [33] [59]

8.2 Detailed results of precipitate kinetic simulation


Table A2: MatCalc results of as-rec. condition and after 1000 h ageing @ 650°C
MatCalc As-Received (Norm. + Temp.) Service 1000 h @ 650°C
Results P91 Phase Mean Number Phase Mean Number
(eie=0.393) Fraction Diameter Density Fraction Diameter Density
M23C6(mart,s) 1.879 % 149.4 nm 9.52·1018 m-3 1.898 % 200.6 nm 3.78·1018 m-3
Laves(mart,s) 0.529 % 6.6 nm 9.29·1021 m-3 0.034 % 286.1 nm 2.73·1016 m-3
VN(mart,d) 0.428 % 41.7 nm 1.09·1020 m-3 0.408 % 41.3 nm 1.03·1020 m-3
18 -3
NbC(aust,d) 0.091 % 75.5 nm 4.04·10 m 0.096 % 75.0 nm 4.04·1018 m-3
AlN(mart,d) 0.043 % 44.3 nm 8.85·1018 m-3 0.043 % 52.5 nm 5.17·1018 m-3
For the abbreviation convention of precipitate names see table 5; <0.01 % phase fraction not stated
Table A3: MatCalc results of 10 000 h aged @ 650°C and 100 000 h aged @ 650°C
MatCalc Service 10 000 h @ 650°C Service 100 000 h @ 650°C
Results P91 Phase Mean Number Phase Mean Number
(eie=0.393) Fraction Diameter Density Fraction Diameter Density
18 -3
M23C6(mart,s) 1.920 % 304.4 nm 1.07·10 m 1.931 % 590.6 nm 1.70·1017 m-3
VN(mart,d) 0.402 % 53.4 nm 4.45·1019 m-3 0.380 % 103.0 nm 5.46·1018 m-3
18 -3
NbC(aust,d) 0.088 % 74.5 nm 4.04·10 m 0.088 % 74.5 nm 4.04·1018 m-3
Laves(mart,s) 0.085 % 389.7 nm 2.73·1016 m-3 0.119 % 433.7 nm 2.73·1016 m-3
AlN(mart,d) 0.042 % 80.2 nm 1.45·1018 m-3 0.040 % 174.0 nm 1.38·1017 m-3
14 -3
Z(mart,VN_m_s) 0.001 % 330.8 nm 3.16·10 m 0.020 % 1126.6 nm 3.16·1014 m-3
For the abbreviation convention of precipitate names see table 5; <0.01 % phase fraction not stated

29
9 LITERATURE
[1] G. Zeiler, Martensitic Steels for rotors in ultra-supercritical power plants, in: A. Di Gianfrancesco (Ed.), Ultra-Supercritical and Advance
Ultra-Supercritical Power Plants, Woodhead Publishing Series in Energy: no. 104, 2017, pp. 143-147
[2] B. Raj, B. Choudhary, A perspective on creep and fatigue issues in sodium cooled fast reactors, T. Indian I. Metals 63 (2010) 75-84
[3] C. Kalck, B. Fournier, F. Barcelo, L. Forest, F. Dalle, P.-F. Giroux, I. Tournié, A.-F. Gourgues-Lorenzon, High temperature creep
properties and microstructural examinations of P92 welds, 9th Liege conference: Materials for advanced power engineering, Liege, Belgium,
2010, pp. 424-434
[4] R. Viswanathan, D. Gandy, K. Coleman, Advances in Materials Technology for Fossil Power Plants, Proceedings from the 4th
International Conference (2004) pp. 37-40, Hilton Head Island, South Carolina, USA
[5] J. Hald, Microstructure and long-term creep properties of 9-12% Cr steels, Int. J. Pres. Ves. Pip. 85 (2008) 30-37
[6] F. Abe, M. Taneike, K. Sawada, Alloy design of creep resistant 9Cr steel using a dispersion of nano-sized carbonitrides, Int. J. Pres. Ves.
Pip. 84 (2007) 3-12
[7] C. Panait, A. Zielińska-Lipiec, T. Koziel, A. Czyrska-Filemonowicz, A.-F. Gourgues-Lorenzon, W. Bendick, Evolution of dislocation
density, size of subgrains and MX-type precipitates in a P91 steel during creep and during thermal ageing at 600 °C for more than 100,000h,
Mater. Sci. Eng. A 527 (2010) 4062-4069
[8] K. Maruyama, K. Sawada, J. Koike, Strengthening Mechanisms of Creep Resistant Tempered Martensitic Steel, ISIJ Int. 41 (2001) 641–
653
[9] C. Pandey, M. Mahapatra, P. Kumar, R. Vidyrathy, A. Srivastava, Microstructure-based assessment of creep rupture behaviour of cast-
forged P91 steel, Mater. Sci. Eng. A 695 (2017) 291-301
[10] E. Cerri, E. Evangelista, S. Spigarelli, P. Bianchi, Evolution of microstructure in a modified 9Cr-1Mo steel during short term creep,
Mater. Sci. Eng. A 245 (1998) 285-292
[11] A. Kostka, K.-G. Tak, R. Hellmig, Y. Estrin, G. Eggeler, On the contribution of carbides and micrograin boundaries to the creep
strength of tempered martensite ferritic steels, Acta Mater. 55 (2007) 539-550
[12] C. Panait, W. Bendick, A. Fuchsmann, A.-F. Gourgues-Lorenzon, J. Besson, Study of the microstructure of the grade 91 steel after more
than 100.000h of creep exposure at 600°C, Int. J. Pres. Ves. Pip. (2010) 1-13
[13] H. K. Danielsen, Review of Z phase precipitation in 9-12 wt-%Cr steels, Mater. Sci. Tech. 32 (2016) 126-137
[14] P. Ennis, A. Zielinska-Lipiec, O. Wachter, A. Czyrska-Filemonowicz, Microstructural Stability and Creep Rupture Strength of the
Martensitic Steel P92 for Advanced Power Plant, Acta Mater. 45 (1997) 4901-4907
[15] N. Saini, R. Mulik, M. Mahapatra, Study on the effect of ageing on Laves phase evolution and their effect on mechanical prope rties of
P92 steel, Mater. Sci. Eng. A 716 (2018) 179-188
[16] J. Lee, H. Armaki, K. Maruyama, T. Muraki, H. Asahi, Causes of breakdown of creep strength in 9Cr-1.8W-0.5Mo-VNb steel, Mater.
Sci. Eng. A 428 (2006) 270-275
[17] F. Larson, J. Miller, A Time-Temperature Relationship for Rupture and Creep Stresses, ASTM Proc. 74 (1952) 765-775
[18] F. Monkman, N. Grant, An Empirical Relationship Between Rupture Life and Minimum Creep Rate in Creep Rupture Tests, ASTM
Proc. 56 (1956) 593-620
[19] P. McVetty, Factors affecting the choice of working stresses for high temperature service, Trans. ASME 55 (1933) 99
[20] J. Conway, M. Mullikin, An evaluation of various first stage creep equations, Proc. of AIME Conf., Detroit, Michigan, 1962
[21] F. Norton, The creep of steel at high temperatures, McGraw-Hill, New York, 1929
[22] A. Nadai, The Influence of Time upon Creep, The Hyperbolic Sine Creep Law, S. Timoshenko 60th Anniversary Vol., Macmillan Co.,
New York, 1938, pp. 155-170
[23] Y. Rabotnov, Creep Problems in Structural Members, North Holland, Amsterdam, 1969
[24] D. McHenry, A new aspect of creep in concrete and its application to design, ASTM Proc. 43 (1943) 1069–1086
[25] R. Sandstroem, A. Kondyr, Model for tertiary-creep in Mo and CrMo-steels, Proceedings-Computer Networking Symposium (1980)
275-284
[26] A. Graham, K. Walles, Relations between long and short time properties of commercial alloys, JISI 193 (1955) 105-120
[27] J. Bolton, A "characteristic-strain" model for creep, Mater. High Temp. 25 (2008) 197-204
[28] S. Holmström, P. Auerkari, Prediction of creep strain and creep strength of ferritic steels for power plant applications, Proc. of Baltica
Conference on Life Management and Maintenance for Power Plants, Helsinki, Finland, 2004, pp. 513-521
[29] S. Holdsworth, M. Askins, A. Baker, E. Gariboldi, S. Holmström, Klenk A., M. Ringel, G. Merckling, R. Sandstrom, M. Schwienheer,
S. Spigarelli, Factors influencing creep model equation selection, Int. J. Pres. Ves. Pip. 85 (2008) 80-88
[30] K. Kimura, Y. Takahashi, Evaluation of Long-Term Creep Strength of ASME Grades 91, 92, and 122 Type Steels, Proc. of the ASME
2012 Pressure Vessels and Piping Conference, vol. 6: Materials and Fabrication, Parts A and B. Toronto, Ontario, Canada, 2012, pp. 309 -316
[31] S. D. Yadav, B. Sonderegger, M. Stracey, C. Poletti, Modelling the creep behaviour of tempered martensitic steel based o n a hybrid
approach, Mater. Sci. Eng. A 662 (2016) 330-341
[32] M. Basirat, T. Shrestha, G. Potirniche, I. Charit, K. Rink, A study of the creep behavior of modified 9Cr-1Mo steel using continuum-
damage modeling, Int. J. Plasticity 37 (2012) 95-107
[33] N. Ghoniem, J. Matthews, R. Amodeo, A Dislocation Model for Creep in Engineering Materials, Res Mech. 29 (1990) 197-219
[34] J. F. Kreyca, State parameter based modelling of stress-strain curves in aluminium alloys, PhD Thesis, Institute of Materials Science and
Technology, TU Wien, Austria, 2017, pp. 25-27
[35] D. Caillard, J. Martin, Experimental Characterization of Dislocation Mechanisms, in: Thermally activated mechanisms in crystal
plasticity, Pergamon Materials Series, Elsevier Science, Cambridge, UK, 2003, pp. 15-18
[36] D. Caillard, J. Martin, Dislocation Climb, in: Thermally Activated Mechanisms in Crystal Plasticity, Pergamon Materials Series,
Elsevier Science, Cambridge, UK, 2003, pp. 281-319
[37] J. P. Hirth, J. Lothe, Theory of Dislocations, 2nd Edition, John Wiley, New York, 1982
[38] J. Lothe, Theory of Dislocation Climb in Metals, J. Appl. Phys. 31 (1960) 1077- 1081
[39] M. Winning, A. Rollett, G. Gottstein, L. Shvindlerman, J. Srolovitz, A. Lim, Mobility of Low Angle Grain Boundaries in Pure Metals,
Philos. Mag. 90 (2010) 3107-3128
[40] E. Orowan, Problems of plastic gliding, Proc. of Phys. Soc. 52 (1940) 8-22
[41] J. D. Holt, Dislocation Cell Formation in Metals, J. Appl. Phys. 41 (1970) 3197-3201
[42] J. Čadek, The back stress concept in power law creep of metals: A review, Mater. Sci. Eng. 94 (1987) 79-92
[43] S. H. Crandall, N. C. Dahl, T. J. Lardner, An Introduction to the Mechanics of Solids, McGraw-Hill, New York, 1959
[44] H. Mughrabi, The alpha-factor in the Taylor flow-stress law in monotonic, cyclic and quasi-stationary deformations: Dependence on slip
mode, dislocation arrangement and density, Curr. Opin. Solid St. M. 20 (2016) 411-420
[45] I. Lifshitz, V. Slyozov, The kinetics of precipitation from supersaturated solid solutions, J. Phys. Chem. Solids 19 (1961) 35-50
[46] J. Pesicka, A. Dronhofer, G. Eggeler, Free dislocations and boundary dislocations in tempered martensitic ferritic steel s, Mater. Sci.
Eng. A 387-389 (2004) 176-180

30
[47] D. R. Askeland, P. P. Fulay, D. K. Bhattacharya, Atomic and Ionic Arrangements, in: Essentials of Materials Science and Engineering,
Second Edition, Cengage Learning, Stamford, USA, 2009, p. 62
[48] G. M. Cheng, W. Z. Xu, W. W. Jian, H. Yuan, M. H. Tsai, Y. T. Zhu, Dislocations with edge components in nanocrystalline bcc Mo, J.
Mater. Res. 28 (2013) 1820-1826
[49] W. Blum, P. Eisenlohr, Dislocation mechanics of creep, Mater. Sci. Tech. A 510-511 (2009) 7-13
[50] U. Essmann, H. Mughrabi, Annihilation of dislocations during tensile and cyclic deformation and limits of dislocation densities, Philos.
Mag. A 40 (1979) 731-756
[51] H. Nitta, K. Miura, Y. Iijima, Self-diffusion in iron-based Fe-Mo alloys, Acta Mater. 54 (2006) 2833-2847
[52] Y. Hasegawa, Grade 92 creep-strength-enhanced ferritic steel, in: A. Shibli (Ed.), Coal Power Plant Materials and Life Assessment-
Development and Applications, Woodhead Publishing Series in Energy no. 62, Sawston, UK, 2014, p. 62
[53] CODATA, The NIST Reference on Constants, Units, and Uncertainty, 2014. Online: https://physics.nist.gov/cgi-bin/cuu/Value?k
(Accessed 24 Jan 2020 from Graz, Austria)
[54] V. Dudko, A. Belyakov, R. Kaibyshev, Evolution of Lath Substructure and Internal Stresses in a 9% Cr Steel during Creep, ISIJ Int. 57
(2017) 540-579
[55] S. Yadav, S. Kalácska, M. Dománkova, D. Yubero, R. Resel, I. Groma, C. Beal, B. Sonderegger, C. Sommitsch, C. Poletti, Evolution of
the substructure of a novel 12 % Cr steel under creep conditions, Mater. Charact. 115 (2016) 23-31
[56] A. Pineau, S. Antolovich, High temperature fatigue: Behaviour of three typical classes of structural materials, Mater. High Temp. 32
(2015) 298-317
[57] K. Guguloth, J. Swaminathan, N. Roy, R. Ghosh, Uniaxial creep and stress relaxation behavior of modified 9Cr-1Mo steel, Mater. Sci.
Eng. A 684 (2017) 683-686
[58] S. Khayatzadeh, D. Tanner, C. Truman, P. Flewitt, D. Smith, Creep deformation and stress relaxation of a martensitic P92 steel at 650
°C, Eng. Fract. Mech. 175 (2017) 57-71
[59] J. Schmid, Modelling the microstructure of a creep resistant steel, Master Thesis at IMAT Institute, Graz University of Technology,
Austria, 2018
[60] B. Hahn, W. Bendick, Pipe steels for modern high-output power plants; Part 1: Metallurgical principles- Long-term properties-
Recommendations for use, 3 R International (2008) 398-407
[61] J. Svoboda, F. Fischer, P. Fratzl, E. Kozeschnik, Modelling of kinetics in multi-component multi-phase systems with spherical
precipitates: I: Theory, Mater. Sci. Eng. A 385 (2004) 166-174
[62] S. Vujic, MatCalc 5 Precipitation Kinetic Script, IWS Institute (now IMAT), Graz University of Technology, Austria, 2015
[63] A. M. El-Azim, O. Ibrahim, O. El-Desoky, Long term creep behaviour of welded joints of P91 steel at 650 °C, Mat. Sci. Eng. A 560
(2013) 678-684
[64] Mitsubishi Heavy Industries Japan, Data from e-mail communication, 2015
[65] F. Riedlsperger, Thermodynamic Precipitation Kinetic Simulation in Martensitic Cr-Steels, Master Thesis, IWS Institute (now IMAT),
Graz University of Technology, Austria, 2016
[66] D. Holt, Dislocation Cell Formation in Metals, J. Appl. Phys. 41 (1970) 3197-3201
[67] G. Stechauner, E. Kozeschnik, Assessment of substitutional self-diffusion along short-circuit paths in Al, Fe and Ni, Calphad 47 (2014)
92-99
[68] Y. Shima, Y. Ishikawa, H. Nitta, Y. Yamazaki, K. Mimura, M. Isshiki, Y. Iijima, Self-Diffusion along Dislocations in Ultra High Purity
Iron, Mater. Trans. 43 (2002) 173-177
[69] H. Oikawa, Y. Iijima, Diffusion behavior of creep-resistant steels, in: F. Abe, T.-U. Kern, R. Viswanathan (Eds.), Creep-Resistant
Steels, Woodhead Publishing, Cambridge, UK, 2008, pp. 241- 264
[70] W. Blum, Discussion: Activation volumes of plastic deformation of crystals, Scripta Mater. 146 (2018) 27-30
[71] B. Krenmayr, Statistische und systematische Abweichungen von Kriech-Simulation und -Experiment in martensitischen Stählen,
Unpublished Manuscript of PhD Thesis, IMAT Institute, Graz University of Technology, Austria, 2020
[72] R. Lim, M. Sauzay, F. Dalle, I. Tournie, P. Bonnaillie, A.-F. Gourgoues-Lorenzon, Modelling and experimental study of the tertiary
creep stage of Grade 91 steel, Int. J. Fract. 169 (2011) 213-228, 2011
[73] A. Aghajani Bazazi, Evolution of Microstructure during Long-Term Creep of a Tempered Martensite Ferritic Steel, PhD Thesis, Ruhr
Universität Bochum, Germany, 2009, pp. 41-44 and pp. 66-69
[74] W. Blum, G. Götz, Evolution of dislocation structure in martensitic steels: the subgrain size as a sensor for creep stra in and residual
creep life, Steel Res. 70 (1999) 274-278
[75] W. Bendick, L. Cipolla, J. Gabrel, J. Hald, New ECCC assessment of creep rupture strength for steel grade X10CrMoVNb9-1
(Grade91), Int. J. Pres. Ves. Pip. 87 (2010) 304-309
[76] K. Kimura, K. Sawada, H. Kushima, Y. Toda, Influence of Chemical Composition and Heat Treatment on Long-term Creep Strength of
Grade 91 Steel, Procedia Eng. 55 (2013) 2-9
[77] R.W. Swindeman, M.J. Swindeman, B.W. Roberts, B E. Thurgood, D.L. Marriott, Verification of allowable stresses in ASME section
III, subsection for NH for for grade 91 steel, Technical Report, 2007, Online: https://www.osti.gov/servlets/purl/974278 (Accessed 8 May
2020 from Graz, Austria)
[78] R. Kaybyshev, V. Skorobogatykh und I. Shchenkova, New Martensitic Steels for Fossile Power Plants: Creep Resista nce, The Physics
of Metals and Metallography 109 (2010) 186-200.
[79] D. Jara, 9-12% Cr heat resistant steels: alloy design, TEM characterisation of microstructure evolution and creep response at 650°C,
PhD Thesis, Ruhr Universität Bochum, Germany, 2011, pp. 58-64
[80] J. Pesicka, A. Aghajani, C. Somsen, A. Hartmaier, G. Eggeler, How dislocation substructures evolve during long-term creep of a 12%
Cr tempered martensitic ferritic steel, Scripta Mater. 62 (2010) 353-356
[81] A. Orlova, J. Cadek, Dislocation Structure in the High Temperature Creep of Metals and Solid Solution Alloys: a Review, Mater. Sci.
Eng. 77 (1988) 1-18
[82] P. Ennis, A. Czyrska-Filemonowicz, Recent advances in creep-resistant steels for power plant applications, Sadhana 28, parts 3 & 4
(2003) 709- 730
[83] S. Yadav, T. Scherer, G. Prasad Reddy, G. Sasikala, S. Albert, C. Poletti, Creep modelling of P91 steel employing a microst ructural
based hybrid concept, Eng. Fract. Mech. 200 (2018) 104-114
[84] S. H. Kim, B. Song, W. Ryu, J. Hong, Creep rupture properties of nitrogen added 10Cr ferritic/martensitic steels, J. Nucl. Mater. 329-
333, Part A (2004) 299-303
[85] A. Kipelova, R. Kaibyshev, A. Belyakov, D. Molodov, Migration of dislocation boundaries in a modified P911 + 3%Co heat resistant
steel during tempering, ageing and creep, Mater. Sci. Forum 715-716 (2012) 953-958
[86] A. Orlova, J. Bursik, K. Kucharova, V. Sklenicka, Microstructural development during high temperature creep of 9% Cr steel, Mater.
Sci. Eng. A245 (1998) 39-48
[87] P. Polcik, T. Sailer, W. Blum, S. Straub, J. Bursik, A. Orlova, On the microstructural development of the tempered martensitic C r-steel
P 91 during long-term creep—a comparison of data, Mater. Sci. Eng. A260 (1999) 252-259

31
[88] L. Maddi, D. Barbadikar, M. Sahare, A. Ballal, D. Peshwe, R. Paretkar, K. Laha, M. Mathew, Microstructure Evolution During Short
Term Creep of 9Cr-0.5Mo-1.8W Steel, Trans. Indian Inst. Met. 68 (2015) 259- 266
[89] A. Benaarbia, X. Xu, W. Sun, A. Becker, M. Jepson, Investigation of short-term creep deformation mechanisms in MarBN steel at
elevated temperatures, Mater. Sci. Eng. A 734 (2018) 491-505
[90] V. Skorobogathykh, I. Schenkova, V. Dudko, A. Belyakov, R. Kaibyshev, Microstructure Evolution in a 9%Cr Heat Resistant Ste el
during Creep Tests, Mater. Sci. Forum 638-642 (2010) 2315-2320
[91] R. Madec, B. Devincre, L. Kubin, T. Hoc, D. Rodney, The Role of Collinear Interaction in Dislocation-Induced Hardening, Science 301
(2003) 1879-1882
[92] R. Madec, L. P. Kubin, Second-order junctions and strain hardening in bcc and fcc crystals, Acta Mater. 58 (2008) 767-770
[93] D. Kiener, R. Fritz, M. Alfreider, A. Leitner, R. Pippan, V. Maier-Kiener, Rate limiting deformation mechanisms of bcc metals in
confined volumes, Acta Mater. 166 (2019) 687-701
[94] A. Seeger, Dislocations in: P. Veyssiere, L. Kubin, J. Castaing (Eds.), C.N.R.S. Paris, France, 1984, pp. 141 -178
[95] C. Weinberger, B. Boyce, C. Battaile, Slip planes in bcc transition metals, Int. Mater. Rev. 58 (2013) 296-314
[96] W. Blum, F. Roters, Spontaneous Dislocation Annihilation Explains the Breakdown of the Power Law of Steady State Deformation,
Phys. Status Solidi A 184 (2001) 257-261

32
Graphical abstract

33

You might also like