You are on page 1of 10

Materials Science and Engineering A 480 (2008) 289298

Phase evolution in P92 and E911 weld metals during ageing


A. Vyrostkova a, , V. Homolova a , J. Pecha b , M. Svoboda c
a

Institute of Materials Research, Slovak Academy of Sciences, Watsonova 47, 040 01 Kosice, Slovakia
b Slovak Energy Machinery, Tov
arenska 210, 93 528 Tlmace, Slovakia
c Institute of Physics of Materials, Academy of Sciences of Czech Republic, Zi
zkova 22, 616 62 Brno, Czech Republic
Received 2 October 2006; received in revised form 3 July 2007; accepted 16 July 2007

Abstract
Phase evolution in the weld metals of P92 and E911 steels weld joints were studied during ageing at 625 C for up to 9000 h. The phases:
ferrite + M23 C6 + MX + Laves found by means of analytical TEM in the annealed states agree with the results of the thermodynamic calculation of
equilibrium phases. The cross-weld hardness values, HV10, after 1000, 3000, and 9000 h ageing overlap each other and are approximately 15 units
below that of the post-weld heat-treated (PWHT) state. Charpy impact energy with the notch at the centerline of the weld metal was measured. Its
values decrease after ageing from approximately 6080 to 12 J compared to the PWHT state. In 912%Cr steel with W this phenomenon can be
explained by a priori heterogeneity in the weld metal, its large former austenite grain size, the precipitation and growth of M23 C6 and Laves phase
particles on grain and packet boundaries.
2007 Elsevier B.V. All rights reserved.
Keywords: Weld metal; P92; E911; Phase evolution; Thermodynamic calculations; Impact toughness

1. Introduction
Environmental and commercial demands on energy production have led to the development of steels operable at
supercritical conditions. An addition of tungsten and molybdenum (1.82W and 0.5Mo, respectively) and optimization of
other alloying elements in 912Cr advanced steels offer almost
30% higher creep rupture strength at 600 C after 105 h than
classical P91 steel [1,2]. Steels such as E911, P122, and P92
(ASME code) based on the steels developed in Japan are among
the most promising. In addition to the more severe operating
conditions in ultra supercritical plants, the newly developed
steels enable weight reduction of the thick section components
of boilers and turbines. Compared with P91 steel, W-containing
steels have worse ductility. The embrittlement has been ascribed
to higher amount of precipitates of intermetallic Laves phase
because of its inherent brittleness and unfavorable size [3,4].
After normalization and tempering, the advanced steels possess tempered martensite microstructure composed of ferrite and
precipitates. The precipitates are mostly of two types, M23 C6
precipitating mainly on grain and subgrain boundaries, and MX

Corresponding author. Tel.: +421 55 7922444; fax: +421 55 7922408.


E-mail address: avyrostkova@imr.saske.sk (A. Vyrostkova).

0921-5093/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2007.07.036

also inside the subgrains [5,6]. During the phase evolution at


ageing the carbide particles coarsen and in relation to their
chemical composition and the temperature and time of ageing
the Laves phase precipitation starts, usually associated with the
M23 C6 particles [7,8]. Tungsten and molybdenum strengthen
effectively the solid solution, resulting in the improvement of
creep resistance. At the same time they also influence precipitation hardening, mainly by affecting the kinetics of secondary
phase precipitation, which may have either positive or negative
influence on creep resistance. The combined effect of tungsten
and molybdenum can be described by an equivalent content of
molybdenum Moeq. [9]:
Moeq. = wt.%Mo + 0.5 wt.%W

(1)

There are different thoughts on the value of molybdenum


equivalent to achieve optimal creep rupture strength. According
to Foldyna and Koukal [10], Moeq. should range between 1.2
and 1.5%, on the other hand Oehmigen et al. [11] recommend
values less than 1. The precipitation of W, Mo-rich Laves phase
can improve creep strength if its coarsening does not proceed
too fast. Hattestrand and Andren [12] confirmed this positive
behavior of the Laves phase in NF616 (identical with P92) steel.
However, the presence of large M23 C6 and Laves phase particles above approximately 0.5 m are generally considered to be
deleterious.

290

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

Table 1
Chemical composition of filler metals
Filler metal

Chemical composition (wt.%)


C

Si

Mn

Cr

Mo

Ni

Nb

P92

W ZCrMoWVNb9 0,5 1,5 wire


E ZCrMoWVNb9 0,5 2 B 4 H5, electrode

0.11
0.11

0.37
0.2

0.45
0.6

8.8
8.8

0.4
0.5

0.6
0.7

1.6
1.6

0.2
0.2

0.06
0.05

0.04
0.05

E911

W ZCrMoWVNb9 1 1 wire
E ZCrMoWVNb9 1 1 B 42 H5, electrode

0.11
0.1

0.25
0.38

0.6
0.45

8.8
9.0

1.0
1.0

0.7
0.7

1.0
1.0

0.2
0.2

0.05
0.06

0.05
0.07

Another important property from the aspect of technology


is weldability. Welding the above-mentioned steels is relatively
difficult concerning the thermal regime of welding and post-weld
heat treatment (PWHT). Weld joints are usually the weakest
parts of a construction, and in the case of ferritic steel weld joints,
the type IV cracking is frequently observed. This is characterized
by a premature failure in the inter-critical and fine-grained heat
affected zones (HAZ) [13,14]. Abe and Tabuchi [15] described
the microstructure evolution during the creep of simulated HAZ,
focusing on the localities of type IV cracking. However, there
is not much information about the properties of the weld metals
in weld joints of the 912CrWMo modified steels. Post-weld
heat treatment of such weld joints is necessary to stabilize the

microstructure, improve its toughness and also to relieve residual


stresses after welding.
The objective of the present work is to describe the phase evolution in weld metals (WM) during isothermal ageing of P92 and
E911 weld joints at 625 C for up to 9000 h. Experimental results
of phase analyses are compared with the results of equilibrium
calculations using the software package Thermo-Calc.
2. Experiment
Weld joints of commercial thick wall tubes D282 56 mm
and D354 46 mm of P92 and E911 steels, respectively, have
been used as the experimental material. For welding the com-

Fig. 1. LM micrographs of P92 (left column) and E911 (right column) weld metals: (a and b) after PWHT and (c and d) after ageing at 625 C for 9000 h.

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

bination of TIG and MMA methods with weaving had been


used, followed by PWHT at 750760 C for 4 h with heating
and cooling rates of 100 C/h. All details concerning the welds
preparation are described in [16]. The chemical composition of
filler materials used is given in Table 1.
Samples of both welds were isothermally aged at 625 C for
1000, 3000, and 9000 h. After the ageing, Charpy impact test
was carried out on the standard 10 mm 10 mm 55 mm samples with V-notch located in the centerline of the weld metal.
For all experimental states the cross-weld hardness HV10 was
measured.
Light microscopy and analytical TEM were used to describe
the microstructure and phase composition of individual states.
Precipitated particles were extracted in carbon replicas and
examined in STEM/EDX (120 kV, Philips). For the particle type
determination the chemical composition evaluated from X-ray
spectra (EDAX software for thin samples without corrections for
absorption or fluorescence) and selected area diffraction (SAD)
were employed. The experimental results of phase analyses have
been compared with the phase equilibrium thermodynamic calculations.
The calculations of phase equilibria were performed by
Thermo-Calc software [17,18]. The database of thermodynamic
parameters formulated by Kroupa et al. [19,20] was used for
the calculations for a FeCrMoVMnWNiSiNbCN
system. In the database, thermodynamic parameters were modified for some phases existing in CrMoV alloy steels. Solid
solution phases (bcc and fcc), carbides, and carbonitrides
(M23 C6 , M7 C3 , M3 C, M6 C, M2 X, and MX), and intermetal-

291

lic phases (Laves, sigma) were taken into account in the calculation.
3. Results
3.1. Microstructure
The microstructure of P92 and E911 steels in the PWHT
state and after ageing at 625 C for 9000 h is shown in Fig. 1.
The tempered microstructure of both weld metals is macroscopically heterogeneous. The areas of former columnar grains with
typical martensite structure comprising large packets of laths
alternate with inter-bead heat affected areas with smaller polyhedral grains of finer martensite. No delta ferrite was observed.
Because the welding procedure used weaving, there is a smooth
transition between columnar areas and inter-bead heat affected
zones.
On the microscopic level, an inhomogeneity of particle distribution is characteristic, the localities with more and less
intensive precipitation in the matrix can be found. After 9000 h
ageing the former austenite, packet, and some of the lath boundaries are almost continuously decorated with large particles
compared to the PWHT conditions, as demonstrated in Fig. 2.
Analysis of the observed and calculated phases of the E911
and P92 weld metals is given in Table 2. In both cases the Cr-rich
M23 C6 and (V, Nb)-rich MX particles have been documented for
all investigated conditions. Laves phase was found only after
ageing at 625 C. Figs. 3 and 4 illustrate a typical distribution
and morphology of the above mentioned secondary phases.

Fig. 2. SEM metallography of P92 (left column) and E911 (right column) weld metals: (a and b) after PWHT and (c and d) after ageing at 625 C for 9000 h.

292

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

Fig. 3. Example of Laves phase in P92 weld metal after ageing at 625 C for 1000 h: (a) TEM micrograph of extraction replica, (b) electron diffraction pattern, and
(c) its solution for B = [2 0 1].

In both weld metals the Laves phase after 1000 h ageing


reached the size up to 600 nm and was situated mainly at boundaries, almost always tied up to the M23 C6 particles. After 3000 h
at 625 C the particles of Laves phase often achieved the size of
1.5 m, and similar results were found after 9000 h.
Chemical composition evolution of the precipitated phases is
schematically documented in Fig. 5. Laves phase in the experimental steels is of (Fe, Cr)2 (W, Mo) type.
Results of thermodynamic calculations are presented in
Figs. 6 and 7. The isothermal sections of phase diagrams for
the FeCrMoVMnWNiSiNbCN system at 625 C
are plotted in Fig. 6. Investigated weld metals are marked in

Fig. 6 by solid circles (weld metal P92 in Fig. 6a, weld metal
E911 in Fig. 6b). Calculated volume amounts of the equilibrium
phases in dependence on temperature are shown in Fig. 7 for
temperature range 527927 C.
3.2. Mechanical properties
Hardness HV10 of the weld metals after ageing decreased
by approximately 15 units compared to the PWHT state and the
values of all aged states overlap, as illustrated in Fig. 8.
A significant drop in impact energy was seen following
ageing (Table 3). The values of impact energy at ambient tem-

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

293

Fig. 4. TEM micrograph of (a) secondary phases in P92 weld metal after ageing at 625 C for 3000 h and (b) example of M23 C6 particle and its electron diffraction
pattern for B = [1 1 1] as inlay.

perature after PWHT reached 70 and 80 J for E911 and P92


welds, respectively. After 1000 h ageing the impact energies
dropped to 19 J for E911 and 21 J for P92 weld. Ageing for
3000 h led to further decrease, to approximately 12 J in both
welds.
Typical fracture surfaces after Charpy impact test of the
states aged for 1000 and 3000 h are shown in Fig. 9. In macroscopic view the fractures are flat, without signs of intensive

plastic deformation. Fig. 9 also indicates that the fractures are


brittle, with quasi-cleavage as a dominant fracture mode. However, there are no typical high ridges created by dimple tearing
between facets. Cleavage facets are inter-connected by narrow areas consisting of shallow dimples with large particles on
the bottom. The dimple inter-crystalline fracture can also be
assumed in many places (Fig. 9bd) where the grain boundaries
are revealed as a result of inter-crystalline de-cohesion.

Fig. 5. Evolution of chemical composition of phases in P92 and E911 weld metals with ageing time.

294

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

Table 2
Phase composition of experimental steels
Phase composition

M23 C6

MX

Laves

P92

PWHT
625 C/10 h
625 C/20 h
625 C/1000 h
625 C/3000 h
625 C/9000 h
Calculated

x
x
x
x
x
x
x

x
x
x
x
x
x
x

x
x
x
x

PWHT
625 C/1000 h
625 C/3000 h
625 C/9000 h
Calculated

x
x
x
x
x

x
x
x
x
x

x
x
x
x

E911

Table 3
Impact energy values at room temperature
State

P92 (J)

E911 (J)

PWHT
625 C/1000 h
625 C/3000 h
625 C/9000 h

87, 62, 66
18, 24, 23
12, 12, 10
13, 20, 20

72, 45, 80, 95, 57, 73


19, 21, 17
10, 13, 12
14, 11, 10

4. Discussion
4.1. Microstructure and phase analysis
The microstructure of the weld metals after PWHT comprised
of tempered martensite with clearly visible acicular structure
initially. This feature slowly disappears with the time of ageing,
resulting in the ferrite-carbide mixture after 9000 h.
According to the thermodynamic calculations, ferrite,
M23 C6 , MX, and Laves phase are the equilibrium phases
for the experimental materials and temperature conditions
(Figs. 6 and 7). These results are in accordance with experimentally found phases except for the states after PWHT, in
which the equilibrium conditions have not been achieved yet.
The volume fraction of M23 C6 phase is about three-times
higher than that of intermetallic Laves phase for the investigated
temperature 625 C, their amounts decrease with increasing
temperature. The amount of MX phase is stable in the calculated temperature range 527927 C for both weld metals
(Fig. 7).
EDX analysis of particles shows that the fine particles of up
to 80 nm in size (after PWHT only up to 20 nm) contain V and
Nb. The wt.% ratio V:Nb is approximately 4.8:1.3 and 4.3:1 for
E911 and P92 WMs, respectively. According to the thermodynamic calculation the particles are of MX type, with a similar
ratio of the above-mentioned elements, and contain nitrogen and
a small amount of carbon, in ratios 19:0.3 and 18:1. The measured particles can be defined as (V, Nb)-rich carbonitrides with
minor contents of Cr and Mo. In high Cr steels particles like these
can be found on all kinds of boundaries as well as in the matrix
[6,8]. After normalizing + tempering, MX carbonitrides precipitate coherently in lath interior. The increased Cr content reduces

Fig. 6. Calculated isothermal section of phase diagram at 625 C: (a) for


the Fe8.8CrMo0.2V0.6MnW0.7Ni0.2Si0.05Nb0.11C0.05N system, position of weld metal P92 is marked by solid circle and (b) for
the Fe9CrMo0.2V0.45MnW0.7Ni0.38Si0.06Nb0.1C0.07N system, position of weld metal E911 is marked by solid circle.

the MX lattice parameter resulting in smaller misfit between the


coherent precipitate and matrix [5,21].
In literature, the existence of unstable M2 X particles in
similar steels is presented as well. Vijayalakshmi et al. [22],
using the Phase evolution diagram concept, showed the coexistence of metastable M2 X particles with ferrite after ageing
of 9Cr1Mo weld metal resulting from the change in carbon
content. Supersaturation of carbon in -ferrite has been considered to be the main thermodynamic driving force for the
evolution of various metastable phases. M2 X phase was formed
after an ageing time as short as 2 h at 750 C, after only 75 h
at 650 C, while at 550 C it did not form at all. Bhadeshia
[23] and others [7,24] mention the M2 X phase existence in

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

Fig. 7. Calculated volume amounts of equilibrium phases MX, M23 C6 and Laves
phase as a function of temperature: (a) for P92 weld metal and (b) for E911 weld
metal.

295

calculated time temperature diagrams for steel similar to our


P92.
Laves phase predominantly nucleates on boundaries of the
former austenite grains and martensite laths in the vicinity of
M23 C6 particles [2527]. Its precipitation starts after a few thousand hours of ageing at 500600 C in classical 9%Cr steels.
In 1%W steel, Dimmler et al. [27] found Laves phase particles
after only 1 h ageing at 650 C. The small particles (equivalent diameter was about 80 nm) began intensively growing after
1000 h exposition. Then up to 30,000 h the area fraction of Laves
phase particles increased while their number was reduced. In
P92 steel with higher tungsten content (1.84%) similar behavior
was observed, but the coarsening stage started earlier and was
much faster. These findings led us to an additional experiment,
ageing our P92 welds for 10 and 20 h at 625 C. EDX spectra were measured for almost 100 particles with different size
(50300 nm), morphology, and location, however, no particles
of Laves phase were found (Table 2). To make sure about the
presence of the phase, more sophisticated methods should be
used.
The agreement between experimental and calculated chemical composition of Laves phase is satisfactory for both materials
examined. The Laves phase in P92 contains a higher amount of
tungsten and less molybdenum (53W:7Mo) than Laves phase
in E911 (41W:16Mo), which corresponds well with the nominal content of these elements in P92 (1.6W:0.4Mo) and E911
(1.0W:1.0Mo) weld materials. The chemical composition of
Laves phase does not change substantially during ageing, and is
close to the equilibrium composition practically from the very
beginning of its existence, as can be seen in Fig. 5.
Chromium-rich carbide M23 C6 was found in all states investigated. M23C6 particles precipitate at all interfaces in the
microstructure (Fig. 4). The existence of M23 C6 particles from
approximately 60 to 700 nm in size after 3000 h, and particularly after 9000 h ageing indicates that coarsening also occurs,
i.e., some particles are growing at the expense of the dissolving
ones. The calculated equilibrium content of the metallic elements such as CrFeWMo in M23 C6 carbide is almost the
same for both weld metals. The content of Fe decreases and the
content of Cr increases with ageing time, approaching the equilibrium values. Contrariwise, the content of tungsten remains

Fig. 8. Cross-weld hardness HV10 after ageing at 625 C: (a) P92 and (b) E911.

296

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

Fig. 9. SEM micrographs of fracture surfaces of P92 (left column) and E911 (right column) weld metals after (a and b) 1000 h and (c and d) 3000 h ageing.

stable during ageing, and in P92 it is almost three times higher


than the calculated value.
4.2. Mechanical properties
Vickers method HV10 was used for hardness testing with
10 kg load. The highest values of approximately 215 were
achieved after PWHT in both weld metals. The scatter of the
hardness values reflects the heterogeneity of the microstructure.
Average hardness after ageing for 1000 and 3000 h is lower by
approximately 15 units. The hardness values of P92 weld metal
after 9000 h ageing clearly overlap those after 3000 h ageing;
however, in the E911 weld metal the scatter of the hardness
values is far greater.

On the other side, the results of Charpy impact tests after


ageing are less promising. The values after PWHT are relatively high, which can be ascribed to the notch location in
the cap layer of the weld, and the tempering conditions used
(750760 C/4 h). Impact energy dropped from 7080 J after
the PWHT to 1712 J after 9000 h ageing (Table 3). Heuser
[28] gives impact energy values 48, 59, and 66 J for a weld
of similar size, composition, and PWHT with the notch in the
centerline of the weld, which compares with our results. Nath et
al. [4] compared the fracture appearance transition temperature
(FATT) of P91 steel with some other steels containing tungsten.
After 10,000 h of ageing at 600650 C, the FATT of the latter
reached as high as +80 C. The brittleness of the steels with W is
ascribed to the Laves phase precipitation and its coarsening, rep-

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

297

Fig. 10. Fracture surface and SEM metallography of weld metals after 9000 h ageing: (a and b) P92 and (c and d) E911.

resenting generally one of the most important processes leading


to the degradation of these steels. The Laves phase particles at
the boundaries after 9000 h ageing are large and dense, as can be
seen in Fig. 2c and d, which supports this idea. Attendant phenomena are the matrix depletion of solid solution strengthening
elements (W, Mo) resulting in the loss of creep strength. This
can be improved by alloying with the stabilizing elements such
as B, Cu, N, Ta that create fine stable particles in the matrix (Cu,
borides, TaC) and decelerates recrystallization of the matrix [29].
Or they segregate on the inter-phase boundaries and/or enrich
some of the precipitated phases (B in M23 C6 ) suppressing particle growth [13,6]. The reduction in toughness is generally related
to the austenite grain size and segregation of surface-active elements on the austenite boundaries. Above 550 C the effect of
segregation diminishes, however, the absorbed energy sharply
decreases because of large particle growth on the boundaries
[29]. Senior [30] discusses the loss of carbide-matrix interphase
strength leading to the formation of voids as a major cause of
the reduction in toughness.
Comparison of fracture surfaces and microstructure of both
weld metals examined after 9000 h ageing is given in Fig. 10 (the
depicted localities have been selected for illustration). The figure
documents that cleavage facets size matches former packet size,
or other smaller microstructure sub-units, and that for both weld
metals the original lath structure can be implied.
The results support the idea of low impact toughness of 9%Cr
steels with tungsten. This phenomenon can be explained by

a large prior austenite grain size of weld metals, and by the


presence of large particles (Laves phase, M23 C6 ) on all kinds
of boundaries resulting in low fracture energies. To avoid this
problem, it seems necessary to modify the thermal regime of
welding by adjusting the welding procedure, e.g., the lower heat
input to avoid grain growth, weaving from one side to another
to use the temper-bead effect and to avoid the probability of
a notch location into the area of two adjoining beads, and then
using suitable PWHT conditions. The abovementioned procedural modifications should result in optimum initial microstructure
for long-term high-temperature operation.
5. Conclusion
Microstructural analysis of the P92 and E911weld metals was
performed on weld joints after PWHT and ageing at 625 C for
up to 9000 h. The results can be summarized as follows:
After PWHT ferrite, M23 C6 , and MX phases were found.
Delta-ferrite was not detected.
Laves phase was found in all aged states.
Thermodynamic calculations predict ferrite, M23 C6 , MX, and
Laves phase as the equilibrium phases in both weld metals
investigated, which is in agreement with obtained experimental results.
Chemical composition of Laves phase is close to the equilibrium from the beginning of its existence. Laves phase in

298

A. Vyrostkova et al. / Materials Science and Engineering A 480 (2008) 289298

P92 weld metal contains more tungsten and less molybdenum (53W:7Mo) than that precipitating in E911 weld metal
(41W:16Mo).
Major metallic elements in M23 C6 carbide are Cr, Fe, W, and
Mo. The content of these elements is similar for both weld
metals, and during ageing the amount of Cr increases at the
expense of Fe.
The chemical composition and volume fraction of MX particles is nearly constant for the calculated temperature range
527927 C in both weld metals.
Hardness of the weld metals does not change substantially
after ageing for 10009000 h and is only approximately
15 units lower than the hardness after PWHT.
Ageing led to a decisive reduction of impact energy at
room temperature. This has been ascribed to an unfavorable microstructure containing large austenite grain size after
welding and very coarse Laves phase and M23 C6 particles
precipitated on interfaces.

Acknowledgements
The present work was supported by Slovak Grant Agency
(VEGA) under Grant No. 2/7197/27, COST Action 536, and
by Slovak Research and development agency under the contract
No. COST-0022-06. Our thanks belong also to Dr. Graham Holloway of Metrode Products Limited, Chertsey, for reading the
manuscript and valuable comments.
References
[1] The T91/P91 Book, Vallourec & Mannesmann tubes, 2002.
[2] The T92/P92 Book, Vallourec & Mannesmann tubes, 2000.
[3] S. Kunimitsu, Y. You, N. Kasuya, Y. Sasaki, Y. Hosoi, J. Nucl. Mater.
179181 (1991) 689692.
[4] B. Nath, E. Metcalfe, J. Hald, in: A. Strang, D.I. Gooch (Eds.), Microstructural Development and Stability in High Chromium Ferritic Power Plant
Steels, London, 1997, pp. 123143.
[5] K. Sawada, K. Kubo, F. Abe, Mater. Sci. Eng. A 319321 (2001) 784
787.

[6] F. Abe, T. Horiuchi, M. Taneike, K. Sawada, Mater. Sci. Eng. A 378 (2004)
299303.
[7] J. Hald, Z. Kubon, in: A. Strang, D.I. Gooch (Eds.), Microstructural
Development and Stability in High Chromium Ferritic Power Plant Steels,
London, 1997, pp. 159177.
[8] J.P. Ennis, A. Zielinska-Lipiec, O. Wachter, A. Czyrska-Filemonowicz,
Acta Mater. 45 (1997) 49014907.
[9] T. Fujita, in: R.D. Conroy, et al. (Eds.), Proceedings of the Materials Engineering in Turbines and Compressors, IOM, London, 1995, pp. 493499.
[10] V. Foldyna, J. Koukal, Zvaranie-Svarovan 12 (2003) 67.
[11] H.G. Oehmigen, P. Lenk, A. Shulze, D. Proft, Bretfeld VGB Kraftwerkstechnik 2 (1999) 80.
[12] M. Hattestrand, H.-O. Andren, Micron 32 (2001) 789797.
[13] S.K. Albert, M. Kondo, M. Tabuchi, F.X. Yin, K. Sawada, F. Abe, Metall.
Trans. 36A (2005) 333338.
[14] J.A. Francis, ISIJ Int. 44 (2004) 19661968.
[15] F. Abe, M. Tabuchi, http://www.msm.cam.ac.uk/phase-trans/2002/papers.
[16] J. Pecha, A. Vyrostkova, P. Brziak, Proceedings of the 15th Conference
Boilers and Energy Devices, March, Brno, 2006.
[17] http://www.thermocalc.se/.
[18] J.O. Andersson, T. Helander, L. Hoglund, P. Shi, B. Sudman, Calphad 26
(2002) 273312.
[19] Internal report of IPM Brno 2006.
[20] A. Kroupa, J. Havrankova, M. Coufalova, M. Svoboda, J. Vrestal, J. Phase
Equilib. 22 (2001) 312323.
[21] Hatterstrand, Schwind, Andren, Mater. Sci. Eng. A. 250 (1998) 2736.
[22] M. Vijayalakshmi, S. Saroja, V.S. Raghunanthan, Scripta Mater. 41 (1999)
149152.
[23] H.K.D.H. Bhadeshia, Proceedings of the Super-High Strength Steels,
Assoc. Italiana di Metallurgica, Rome, 2005, pp. 110.
[24] J.D. Robson, H.K.D.H. Bhadesia, Mater. Sci. Technol. 13 (1997) 640644.
[25] J. Orr, L.W. Buchanan, H. Everson, Proceedings of the International conference Advanced Heat Resistant Steels for Power Generation, San Sebastian,
April, 1998, IOM Communications.
[26] Kenji Hayashi, Toshifumi Kojima, Yusuke Minami, Proceedings of the
International Conference Advanced Heat Resistant Steels for Power Generation, San Sebastian, April, 1998, IOM Communications.
[27] G. Dimmler, P. Weinert, E. Kozeschnik, H. Cerjak, Mater. Charact. 51
(2003) 341352.
[28] H. Heuser, C. Jochum, F.W. Meyer, Entwicklung von Schweizusatzwerkstoffen fur moderne Kraftwerkstahle. Tagungsband der 6. Werkstoffseminar TU Graz, 1999, 114.
[29] M. Tamura, K. Shinozuka, H. Esaka, S. Sugimoto, K. Ishizawa, K.
Masamura, J. Nucl. Mater. 283287 (2000) 667671.
[30] D.A. Senior, Acta Metall. 36 (1988) 18551861.

You might also like