Você está na página 1de 28

Creep resistant steels for power plant OMMI (Vol 1, No.

1) April 2002

Recent Advances in Creep Resistant Steels for Power Plant Applications

P J Ennis1 and A Czyrska-Filemonowicz2


1
Research Centre Jülich, Institute for Materials and Processes in Energy Systems, IWV-2
D-52425 Jülich, Germany
2
University of Mining and Metallurgy, Faculty of Metallurgy and Materials Science
Al.Mickiewicza 30, PL-30059 Kraków, Poland

Mr Phil Ennis is at the Research Centre Jülich, Institute for Materials and
Processes in Energy Systems, Julich, Germany. He is concerned withh
materials development for advanced fossil-fired power plant, with emphasis
on the creep rupture properties and microstructure of high chromium steels
and nickel-based alloys.

Abstract

The higher steam temperatures and pressures required to achieve an increase in the thermal
efficiency of fossil fuel fired power generation plant necessitate the application of steels with
improved creep rupture strength. The 9% chromium steels developed during the last three
decades are of great interest for components of advanced, high efficiency power plant. In this
Paper, the development of the steels P91, P92 and E911 are described. It is shown that the
martensitic transformation in all three steels produces a high dislocation density that confers a
high degree of transient hardening. However, the dislocation density decreases during
exposure at service temperatures due to recovery effects and for the long-term creep strength
the sub-grain structure that is produced by the martensitic transformation and the precipitation
of carbides, nitrides and carbonitrides are the decisive microstructural features. The changes in
the microstructure mean that great care is needed in the extrapolation of experimental data to
obtain design values. Only data from tests with rupture times above 3000 h provide reasonable
extrapolated values. It is further shown that for the 9% chromium steels, the oxidation
resistance in steam is not sufficiently high for their use as thin-walled components at
temperatures of 600°C and above. The potential for the development of steels of higher
chromium content (above 11%) to give an improvement in steam oxidation resistance whilst
maintaining creep resistance to the 9% chromium steels is discussed.
Key words: chromium steels, microstructure, creep, steam oxidation

1. Introduction

The constraints that are currently placed on power generation plant in terms of environmental
impact and economics have focussed attention on the development of high efficiency, low
emission systems. If the thermal efficiency of generating plant can be increased, fuel can be
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 2

conserved (less fuel is required for a given power output) and emissions reduced (lower fuel
consumption means lower emissions of environmentally damaging gases). An increase in the
thermal efficiency of power plant can be most effectively achieved by increasing the
temperature and, to a lesser extent, the pressure of the steam entering the turbine. The most
modern steam power stations now in operation reach efficiencies of around 42% with steam
temperatures of 600°C and pressures of 250-300 bar. The next generation of steam power
plant should be capable of operating with steam at 625-650°C, to enable thermal efficiencies
of around 45% to be achieved. A further enhancement of the thermal efficiency may be
obtained by combining an advanced steam cycle plant with a gas turbine; in this way,
efficiencies of over 50% are possible. Of course, the increasing operating temperatures and
pressures impose increasingly stringent requirements on the materials of construction. In the
present paper the authors will consider the developments that have taken place in the high
chromium ferritic/martensitic steels for advanced steam power plants.

In Figure 1, the stress rupture strengths of the currently used and the new power station steels
are compared on the basis of the maximum service temperature for a 100 000 h stress rupture
strength of 100 MPa. It may be seen that the maximum service temperature increases with
increasing complexity of the steel composition and the more highly alloyed steels have
sufficient stress rupture strength to be considered for application at temperatures in excess of
600°C. Indeed the new high chromium steels have similar stress rupture strengths to the
austenitic stainless steels. There are several reasons for the reluctance to use the austenitic
steels; obviously the increased cost of the steel with the high chromium and nickel contents is
a disadvantage but there are technical problems because the thermal expansion coefficient of
austenitic materials is at least 50% higher than that of ferritic steels. This means that care has
to be taken during cooling and heating to avoid excessive thermal stresses that can lead to
fatigue failures.

1Cr0.5Mo

2.25Cr1Mo

9Cr1Mo

12Cr1MoV

P91

E911

P92

AISI 316

AISI 321

AISI 347

500 525 550 575 600 625


maximum operating temperature in °C, based on a
100 000 h average stress rupture strength of 100
MPa

Figure 1: Stress rupture strengths of the currently used and the newly developed power station
steels
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 3

2. Development of 9% chromium steels

In the 1970s, there was considerable interest in the 9% chromium steels for components of the
fast breeder nuclear reactor. On the basis of the familiar Fe9Cr1Mo steel used since the 1950s
in petrochemical plant, an improved steel was developed by the Oak Ridge National
Laboratory (Sikka et al 1981) and subsequently incorporated into the ASTM specifications
under the designation P91 (ASTM 1986). A remarkable increase in the stress rupture strength
was achieved by the addition of 0.2% V, 0.06 Nb and 0.05 N. In Japan, a steel development
programme of Nippon Steel led to the steel NF616 (Nippon Steel 1991), which is now
designated P92 in the ASTM specification. With P92 a further increase in stress rupture
strength was obtained by an addition of 1.8% W and a reduction of the Mo content from 1 to
0.5%. In the European COST (Co-operation in Science and Technology) Action 501, a similar
9% chromium steel was developed; this steel is designated E911, contains 1% Mo and 1% W,
and offers similar stress rupture strength to P92 (Staubli et al 1998).

Chemical compositions and production details for the high chromium steels that were
investigated are given in Table 1. Although much of the work described was carried out on the
steel P92, the findings are in principle also applicable to P91 and E911. The essential
difference between the steels is the tungsten content and with increasing tungsten the
propensity for Laves phase Fe2(Mo,W) formation increases.

Table 1. Details of high chromium steels, chemical compositions in wt%

Element P9 P91 P92 E911


C max. 0.15 0.10 0.124 0.105
Si 0.20-0.65 0.38 0.02 0.20
Mn 0.80-1.30 0.46 0.47 0.35
P max. 0.030 0.020 0.011 0.007
S max. 0.030 0.002 0.006 0.003
Cr 8.5-10.5 8.10 9.07 9.16
Mo 1.70-2.30 0.92 0.46 1.01
W - - 1.78 1.00
V 0.20-0.40 0.18 0.19 0.23
Nb 0.30-0.45 0.073 0.063 0.068
B - - 0.003 -
N - 0.049 0.043 0.072
Ni max. 0.30 0.33 0.06 0.07
Al - 0.034 0.002 -
Form and pipe, Ø 159, pipe, Ø 300, flat bar,
dimensions, 20 wall thickness 40 wall thickness 100 x 16
mm
Heat treatment 1 h/1050°C + 2 h/1070°C + 1 h/1050°C +
1 h/750°C, 2 h/775°C, 1 h/750°C,
air cooled air cooled air cooled
100000h stress
rupture strength 35 94 115 110
at
600°C, MPa*

Values for P91 from Canonico 1991, for P92 from Wachter et al 1995, for E911 from Staubli et al 1998
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 4

3. Characteristics of high chromium steels

The Fe-Cr constitutional diagram is shown in Figure 2.

1600

1400

1200
temperature, °C

austenite
1000
ferrite
800

600

400
Ms
200
0 5 10 15 20
Cr content, wt %

Figure 2: Fe-Cr constitutional diagram

At compositions near to 9% Cr, there is an extensive austenitic region from 820 to 1200°C
and the two-phase region between austenite and ferrite has a very narrow temperature range.
This means that it is possible to austenitise the steel and on cooling to produce a practically
fully martenisitic structure, with minimal amounts, if any, of delta ferrite, which is generally
regarded as detrimental for the high temperature strength properties. The high creep rupture
strength of the P91 steel relies on the martensitic transformation hardening with additional
strengthening by precipitation of carbides, nitrides and carbonitrides of Nb and V. In P92 and
E911, the W additions contribute to additional solid solution strengthening of the martensitic
matrix. The development line is illustrated in Figure 3 which shows the 100 000 h stress
rupture strengths of P9 (data from ISO 1981), P91 (Canonico 1991), P92 (Wachter et al 1995)
and E911 (Staubli et al 1998) at 600 and 650°C.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 5

150

- 0.5 Mo + 0.8 W

105 h rupture stress, MPa


+1W
100

+ 0.04 N + 0.2 V
+ 0.08 Nb

50
600°C

650°C

0
P9 P91 E911 P92

Figure 3: The development of 9% chromium steels

3.1 Microstructure

The 9-12% chromium steels are heat treated to produce a martenisitic microstructure that is
subsequently tempered to improve the ductility and impact strength at low temperatures. The
heat treatment consists of austenitisation at temperatures around 1100°C followed by a
tempering at around 750°C.

3.1.1 Microstructure of P92 after austenitising

Because of the high chromium content, cooling in air is sufficient to initiate the martensitic
transformation after austenitising treatment. Figure 4 shows the microstructure of the steel
P92 after the austenitising at 970°C for 2 h. The steel exhibited a martensitic structure with a
high dislocation density and a small amount of retained austenite at the lath boundaries.
Within the large martensite laths, needle-like Fe-rich M3C particles formed a Widmanstätten
structure with the usual Bagaryatskii orientation to the ferritic matrix. During austenitisation
at 970°C, not all M23C6 particles were dissolved, whereas austenitisation at 1070°C and above
led to complete dissolution of this carbide type. The Nb(C,N) precipitates were observed in all
specimens after austenitisation and their presence in the structure may have inhibited the
austenite grain growth. The size of the particles suggested that they were not dissolved during
austenitisation.

The effect of variations in the austenitising temperature on the microstructure is shown in


Figure 5. The main difference in the microstructure was the increase in lath width (from 0.38
nm at 970°C, to 0.42 nm at 1070°C and 0.58 nm at 1145°C) and in the prior austenite grain
size, which increased from 10 µm at 970°C to 20 µm at 1070°C and 60 µm at 1145°C. The
microstructural changes caused by austenitising of P92 steel in different temperatures are
described by Ennis et al 1997 and Zielinska et al 1997.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 6

Figure 4: Microstructure of the steel P92 after austenitising at 970°C for 2h.
(a) optical micrograph showing martensite laths with a small amount of retained austenite
(b) transmission electron micrograph showing needle-like Fe- rich M3C particles with a
Widmanstätten structure within martensite laths
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 7

0.25 µm

0.25 µm

Figure 5: TEM micrographs of P92 austenitised for 2 h at (a) 970°C and (b) 1145°C, and
subsequently tempered 2 h / 775°C, showing the increased martensite lath width at the higher
austenitising temperature

3.1.2 Microstructure of P92 after tempering

Figure 6. shows the microstructures of P92 austenitised at 1070°C and tempered at 715, 775
and 835°C. During tempering, two main processes have taken place. First, recovery caused a
reduction in the high dislocation density after austenitization and the formation of sub-grains
and dislocation networks. These processes were accelerated at the higher tempering
temperatures, so that tempering at 715°C led to slightly higher dislocation density than the
standard tempering at 775°C. Tempering at 835°C caused a sharp reduction in the dislocation
density of about 75%.

Secondly, the precipitation of carbides, nitrides or carbonitrides occurred during tempering


(Figure 7). The M3C precipitated after austenitisation dissolved as the more stable carbides or
nitrides of chromium, molybdenum, niobium and vanadium formed. M23C6 was precipitated
on prior austenite grain boundaries, on subgrain boundaries and within the martensite laths.
The precipitates that are important for the mechanical properties of P92 are the fine M(C,N):
spheroidal Nb-rich carbonitrides and the plate-like V-rich nitrides. The larger, spheroidal
particles of Nb(C,N) appear to have remained undissolved during austenitizing and during
tempering act as nucleation sites for the plate-like V-rich nitrides, thus forming the V-wing
complexes (Zielinska et al 1997). The small number of such precipitates, however, make it
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 8

unlikely that they are of great significance for the mechanical properties of this steel.

0.25 µm

Figure 6: TEM micrographs of P92 austenitised at 1070°Cfor 2h and tempered for 2 h at (a)
715°C, (b) 775°C and (c) 835°C; the micrographs show increased recovery of martensite and
decreased dislocation density as the tempering temperature is raised
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 9

0.25 µm

Figure 7: TEM micrograph of P92 austenitised at 970°C for 2h and tempered at 775°C.
Large M23C6 on sub-grain boundaries and fine M.(C,N) within the sub-grains.

The results of the microstructural parameter measurements (dislocation density, sub-grain


width and precipitate dimensions) are summarised in Figure 8, with the quantitative
measurements normalised to the steel in the usual heat treatment condition (2 h/1070°C +
2h/775°C). The actual values are given in Table 2. Data for creep specimens tested at 600°C
are included for comparison. It can be seen that the dislocation density decreased by a factor
of around 3 after high temperature tempering treatment and after creep exposure for a few
thousand hours at 600°C. The sub-grain width was substantially increased by creep
deformation, being a factor of 3-4 higher in the creep specimens tested at 600°C for 10 000 h
or more than in the pre-test condition. The precipitate dimensions were not so strongly
affected, but significant increases in the M23C6 dimensions were seen after creep testing.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 10

2
970/775
1070/715 3.3 3.6
1070/775
1070/835
1145/775
1.5 creep 600/1500 h
creep 600/10000 h
creep 600/33000 h
relative value

0.5

0
1
dislocation 2
subgrain mean3diameter 4
mean diameter
density width of M23C6 of M(C, N)
microstructural feature

Figure 8: Results of quantitative measurements of P92 microstructural features after different


heat treatments: 2h/970°C +2h/1070°C, 2h/1070°C + 2h/715°C, 2h/1070°C + 775°C (as
received), 2h/1070°C + 835°C, 2h/1145°C+2h/775°C and after creep exposure at 600°C for
1500, 10000 and 33000 h

Table 2: Quantitative microstructural parameters for P92 in different heat treated


conditions and after creep testing

Taust Ttemp Creep Test Dislocation Sub-grain Mean Dia Mean Dia
°C °C Density, Width, µm M23C6, nm MX, nm
1014 m-2
970 775 - 8.7 ± 1.2 0.38 ± 0.1 - -

1070 715 - 9.0 ± 1.0 0.37 ± 0.1 72 ± 16 14 ± 1

1070 775 - 7.5 ± 0.9 0.42 ± 0.1 89 ± 13 16 ± 1

1070 835 - 2.3 ± 0.6 0.50 ± 0.1 82 ± 12 16 ± 1

1145 775 - - 0.58 ± 0.1 68 ± 18 16 ± 1

1070 775 600°C/1500 h 5.3 ± 0.6 0.70 ± 0.1 119 ± 8 18 ± 1

1070 775 600°C/10000 h 2.5 ± 0.5 1.4 ± 0.1 125 ± 10 21 ± 2

1070 775 600°C/33000 h 1.5 ± 0.4 1.5 ± 0.1 131 ± 12 30 ± 3

The other particle precipitated in 9-12% chromium steels is Laves phase, Fe2(W,Mo), see
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 11

Figure 9. This takes place after long term ageing or creep at 600 and 650°C. It was observed
that the growth of the Laves phase occurred during first 10 000h of ageing at 600°C; and
Hättesrand and Andrén 2001 determined the volume fraction to be 1%. The final size of Laves
phase is much larger in the specimens aged at 650°C in comparison to those aged at 600°C.

0.25 µm

Figure 9: Laves phase precipitated in P92 steel after creep at 650ºC for 6500 h.

3.1.3 Comparison of the microstructure of P91, P92 and E922 steels

After the standard heat treatment (given in Table 1), all three steels exhibited similar
microstructures, as shown in Figure 10. Austenitising produced a martensitic structure with a
high dislocation density within the martensite laths. During tempering recovery caused the
formation of sub-grains and dislocation networks. The creep strength of 9-12% Cr steels, is
correlated inversely with the martensite lath width and therefore with the sub-grain size.
Measurements of the average sub-grain width and of the dislocation density within the sub-
grains, performed by means of quantitative TEM, are presented in Table 3.

It can be seen that sub-grain size is fairly similar in all steels investigated. The small
differences are connected with different prior austenite grain size. Dislocation densities in P91
and P92 steels were similar, in both steels a little higher than in E911 steel. However, it must
be borne in mind that only one heat of each steel has been examined in detail and the small
differences observed may not be significant in the light of heat to heat variations.

Besides recovery processes, the precipitation of carbides, carbonitrides and nitrides occurred
during tempering. In all three steels examined, the M23C6 carbides containing Cr, Fe, Mo (W)
precipitated preferentially on the prior austenite grain boundaries and on the martensite lath
boundaries. These precipitates retard the sub-grain growth and therefore increase the strength
of the materials. In P91 steel mainly spheroidal Nb-rich carbonitrides are observed within the
martensite laths. In the P92 and E911 steels three types of MX; Nb(C,N), plate-like VN and
small complex Nb(C,N)-VN, were found (Ennis et al 1997, Zielinska et al 1997, Hättesrand
and Andrén 2001, Ennis et al 2000, Vanstone 1998, Hald et al 1998.

The microstructural evolution of the P91, P92 and E911steels during ageing at service
temperatures is discussed in detail by Hald 1998, and Ennis and Cyrska-Filemonowicz 2001.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 12

As an example, the microstructural evolution of P92 steel during very long term creep
deformation at 600 and 650oC up to about 57500 h will be discussed in detail in the following
section.

0.25 µ m

0.25 µ m

0.25 µ m

Figure 10: Microstructures of the as received steels (a) P91, (b) P92 and (c) E911
showing the elongated sub-grains of tempered martensite (TEM).
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 13

Table 3. Dislocation density & mean sub-grain size of as received P91, P92, E911 steels

Steel Dislocation Density x 10 –14 , m.-2 Mean Sub-Grain Size, µm

P91 7.5 ± 0.8 0.4 ± 0.06


P92 7.9 ± 0.8 0.4 ± 0.09
E911 6.5 ± 0.6 0.5 ± 0.05

3.2 Creep rupture properties

From the creep curves, the secondary creep rate was derived by taking the slope of the
straight-line portion of each curve. Figure 11 shows the creep strength, plotted as iso-stress
curves at 120 MPa for the secondary creep rate, of the P92 material in different heat treatment
conditions. Decreasing the austenitising temperature from the usual 1070°C to 970°C did not
significantly affect the creep strength. However, increasing the tempering temperature from
775 to 835°C, a temperature just above the Ac1 transformation temperature of 825°C, led to a
marked fall in the creep strength. The low initial dislocation density produced by this heat
treatment, with no other significant differences in the microstructural parameters, is the reason
for the low creep strength. This effect is particularly relevant for weldments, since in the heat
affected zone there will always be a region which has been exposed to temperatures around
850°C and therefore this region will be the most likely site for creep failure according to
Eggeler et al 1987.
1

2 h/1070°C + 2 h/835°C
0.1
secondary creep rate, %/h

initial ferritic structure


0.01 (2 h/1070°C, fc to 780°C,
8 h hold, fc to RT)

0.001 2 h/970°C + 2 h/775°C

0.0001

0.00001 as received
(2 h/1070°C + 2 h/775°C)

0.000001
575 600 625 650 675 700
test temperature, °C

Figure 11: Iso-stress plots at 120 MPa of secondary creep rate for P92 in as-received
condition and after different heat treatments: 2h/970ºC+2h/775ºC; 2h 1070ºC+2h/775ºC;
2h/1070ºC+2h/835ºC; 2h/1070ºC, furnace cool to 780ºC, 8h/780ºC, furnace cool to RT
(ferritic structure, no martensite)
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 14

The dislocation density will decrease not only as a result of an increased tempering
temperature but also during prolonged exposure at lower temperatures typical of service
conditions. This can be seen in the quantitative TEM investigations carried out on tested creep
specimens. After long testing times, the dislocation density was found to be around 2 x 10 14
m -2, a decrease of 75% compared with the as received material: The microstructural changes
that occur in the first 3000 h of exposure mean that there is therefore a danger of over-
estimating the long-term creep strength of P92 if there is a preponderance of short duration
data in the evaluation, in which the dislocation density remains at a high level during the
whole test. The extrapolation of the experimental data for estimation of the long-term creep
rupture strength will be discussed later in more detail.

Figure 12 shows the secondary creep rates of P91, P92 and E911 at 600 and 650°C plotted
against the applied stress (σ). Data taken from published literature for the same heat are
included where available. It can be seen that at high stresses the differences in the secondary
creep rates of the three steels are relatively small. In the low stress region, however, the
differences between the steels become more pronounced. The creep deformation
characteristics may be described by the Norton equation (minimum creep rate is proportional
to the applied stress to the power n) with two different values for the Norton stress exponent
n. At high stresses, the value of n is around 16 while at lower stresses, the data conform to an
n value of 6. The change in n is indicative of a change in the creep characteristics. Figure 12
also shows that at high stresses the differences in the secondary creep rates of the three steels
were relatively small and the high dislocation densities resulting from the martensite
transformation dominate the deformation. In the low stress region, however, the differences
between the steels became more pronounced. The evolution of the microstructure was
examined to investigate the reason behind the two deformation regions.

10
Secondary creep rate, %/h

1 650°C

0.1

0.01 P91
600°C
0.001
E911
0.0001
P92
0.00001 open points: data from
Nippon Steel for P92

0.000001 1.8

70 80 100 120 140 160 200 240


Stress, MPa

Figure 12: Secondary creep rates for P91, P92 and E911
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 15

3.3 Microstructural stability of P92 during long creep deformation

The degree of change in the microstructure is dependent of duration and the temperature of
creep deformation. Well developed sub-grains of low dislocation density in the interiors are
characteristic features of long-term exposed specimens. Figure 13 shows the microstructures
of P92 steel creep deformed at 600oC up to 32909 h and at 650oC up to 27433 h.

The results of quantitative measurements of dislocation density and sub-grain width with
increasing creep deformation time for P92 specimens exposed at 600 and 650°C are shown in
Figure 14. The quantitative evaluation of the microstructure showed that in the first 3000 h of
exposure there was a rapid reduction in the dislocation density and an increase of the sub-
grain width. After long testing times, the dislocation density was around 2 x 10 14 m -2, a
decrease of 75% compared with the as received material.

Similar effects were observed for the P91 and E911 steels. The microstructure of the P91
specimens creep specimens tested at 600oC for 9 200 h and steel E911 tested for 17 500 h
exhibited polygonal sub-grains as a result of the extensive deformation. The dislocation
densities in the steels P91, P92 and E911 after creep testing at 600°C are shown in Table 4. As
observed in P92 steel, the mean sub-grain widths of E911 and P91 increased and the
dislocation densities within the sub-grains decreased with increasing creep exposure.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 16

0.25 µm

0.25 µm

Figure 13: TEM micrographs of P92 steel after creep deformation


(a, top) at 600oC for 33000 h, and (b, bottom) at 650ºC for 27500 h
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 17

10
P92 creep specimens

dislocation density in 10 14 x m -2
subgrain width in µ m or
8 600°C
650°C

dislocation density
4

subgrain width
0
0 10000 20000 30000 40000 50000 60000
exposure time, h

Figure 14: Dislocation densities & sub-grain widths for P92 after creep testing at 600, 650°C

Table 4. Dislocation density in steels P91, P92 and E 911 creep tested at 600°C

ca 1000 h ca 10 000 h ca 17 000 0h 22 000h

P91 (4.8 ± 0.5) x 10 14 (1.1 ± 0.4) x 10 14 nd nd

P92 (5.4 ± 0.5) x 10 14 (2.5 ± 0.5) x 10 14 (2.3 ± 0.5) x 10 14 nd

E911 (5.0 ± 0.5) x 10 14 nd (2.2 ± 0.4) x 10 14 (2.1 ± 0.4) x 10 14

nd- not determined

The other important microstructural changes during creep deformation of the investigated
steel were the size, morphology and distribution of the carbide, nitride and carbonitride
precipitates as well as the chemical composition of the precipitates and matrix. The fine MX,
mainly plate-like VN and spheroidal Nb(C,N), precipitated intragranularly and act as obstacles
for moving dislocations, thus contributing to increased creep strength of P92 steel (Figure 15).
Some complex carbonitrides, consisting of a spheroidal Nb-rich particle to which the V-rich
particle was attached, were also present. Sub-grain boundaries and prior austenite grain
boundaries were pinned by M23C6. These precipitates retarded the sub-grain growth and
therefore increased the strength of the materials.

In the steel exposed for longer times, Laves phase Fe2(W,Mo) (Figure 16) and other
intermetallic phases were formed. The particles of Laves phase formed during creep at 650°C
were much larger than those precipitated at 600°C. The stacking faults formed within the
Laves phases could be observed as characteristic streaks on the selected area electron
diffraction pattern (Figure 16b). This effect allowed easy distinction between Laves phase and
the M23C6, as both particles were precipitated very frequently in close vicinity, preferentially
on sub-grain boundaries. Statistical measurements of the mean particle sizes revealed
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 18

significant coarsening of the M23C6 and intermetallic phases. The mean particle diameter of
M23C6 increased with increasing exposure time at 600 and 650°C, while the MX revealed
insignificant change in size.

0.25 µm

Figure 15: Dislocation interactions with fine carbide precipitates in a steel P92 specimen
after creep testing at 650°C for 6468 h

0.2 µm

Figure 16: (a) Laves phase formed during creep deformation of P92 steel at 650°C for 27500
h and (b ) the corresponding electron diffraction pattern

Precipitation processes in P92 are influenced not only by temperature but also by stress.
Figure 17 shows the results of statistical measurements of the M23C6 and MX particles formed
in P92 specimens (head and gauge lengths) after creep deformation at 600°C. It can be seen
that the coarsening of M23C6 carbides is accelerated by stress while the effect of stress-
induced coarsening of MX is insignificant. These findings concerning the coarsening of
carbides in chromium steels are in good agreement with other studies (Hättesrand and Andrén
2001, Eggeler et al 1987, Schaffernak et al 1998).
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 19

200

mean diameter of particles in nm


gauge length
150
M 23C6

specimen head
100
P92 creep specimens
600°C
50
MX

0
0 10000 20000 30000 40000 50000 60000
exposure time in h

Figure 17: Statistical measurements of the M23C6 and MX particles formed in P92 specimens
(head and gauge lengths) after creep deformation at 600°C

In addition to precipitate morphology and distribution, the chemical compositions of the


phases that precipitate with increasing exposure duration are important. STEM/EDS analyses
showed that the M23C6 are enriched in Cr, Fe, Mo and W, whereas Laves phase is enriched in
W and Mo (Czyrska-Filemonowicz et al 2001). An increase in the creep deformation
temperature from 600 to 650°C results in precipitation of much larger particles of Laves
phase. With increasing creep duration, the precipitation of the Laves phase removes Mo and
W from the matrix solid solution and the strengthening of the matrix is decreased. This will be
discussed in the next section.

Our results are in general agreement with other TEM investigations describing the
microstructure during creep deformation of 9% Cr steels (Hättesrand and Andrén 2001b,
Vanstone 1998, Haald et al 1998, Foldyna et al 1996, Nowakowski et al 1998). Strang and
Vodarek 1998, however, reported that there is another important precipitate that contributes to
the softening of high Cr steels, namely the formation of large particles of Cr(Nb,V)N nitride
(Z phase), replacing to some extent the fine VN precipitates. In the P92 steel the Z phase has
not yet been observed in specimens exposed for up to 57 000 h (Ennis et al 2000).

The question arises as to the relevance for long-term creep strength of the high initial
dislocation density, which results from the martensitic transformation and which during the
first few thousand hours of exposure at service temperatures decreases. To clarify this, creep
tests were carried out on P92 specimens heat treated so that the martensitic transformation
was suppressed. The microstructure consisted of ferrite with a very low dislocation density
(9.8 x 1013 m-2) with carbide (M23C6) and carbonitride precipitates. Some of the M23C6
particles were particularly coarse, up to around 1 µm diameter. The secondary creep rate at
600°C, 120 MPa for the specimen with the ferritic microstructure was found to be 10 000
times higher that that of P92 in the usual austenitised and tempered condition. Figure 18
compares the 600°C stress rupture strengths of martensitic and ferritic P92 specimens.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 20

2.38

20
P92
18

14 ferritic matrix P92


stress, MPa

10

P91
80

P9

50

1.48

30
100 1000 10000 100000
time to rupture, h

Figure 18: Stress rupture curves at 600°C for P92, P91 and P9 compared with the values for a
ferritic structure

It is clear that the martensite transformation makes a considerable contribution to the creep
rupture strength of the 9% Cr steels, even though the martensitic structure degenerates into
ferrite and the initially high dislocation density decreases with exposure time. The initial
martensitic structure does allow the formation of a stable subgrain structure and it is this
substructure together with the fine carbonitride precipitates which confers the high creep
rupture strength.

A ferritic structure containing similar fine carbonitride precipitates to the martensitic P92
cannot provide the creep rupture strength required. This means that steel developments to
strengthen the 9% Cr steels for applications at 600°C and above will be restricted by the
necessity for the steels to exhibit the martensitic transformation. The maximum operating
temperature must be sufficiently below the tempering temperature to avoid too rapid a
recovery of the martensite.

4. Assessment of long-term creep rupture strength

The prediction of the long-term strength, using any extrapolation procedure is made difficult
due to several factors. The scatter in the experimental data, especially between different heats
within the compositional and heat treatment specifications for a given steel; the temperature
and duration ranges of the available data; microstructural changes that occur in the materials
during testing and influence the deformation; and environmental effects, such as oxidation,
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 21

that reduce the effective load-bearing cross-section may all lead to erroneous predictions of
strength. One of the earliest attempts to carry out extrapolation of creep rupture data was made
by Norton 1929, although this book is better known for the first publication of the creep
power law equation, now widely used in creep deformation modelling. By relating the
secondary creep rate to the applied stress and then using for design a maximum tolerable creep
rate, a high temperature component could be dimensioned. Larson and Miller 1952 applied the
Arrhenius rate equation for thermally activated creep processes. By assuming that the
secondary creep rate was inversely proportional to the rupture time, tR, that is, the faster the
specimen creeps, the shorter the rupture time, a time-temperature parameter could then be
derived:

PLM = T (C + log tR )
where T is the temperature in K, tR the rupture time and C is a constant, related to the
activation energy for creep. This implies that long times at low temperatures are equivalent to
shorter times at higher temperatures. This enables all stress rupture data to be plotted on a
single master curve and for estimation of the stress rupture strength, no extrapolation of the
curve beyond the experimental data points is needed. The use of this parameter does,
however, dictate the form of the stress rupture data curves; if the log of the rupture time is
plotted as a function of the reciprocal test temperature in K for constant stress, straight lines
should be obtained which intersect at 1/T = 0, log tR = -C.

The microstructural studies and the analysis of the stress dependence of the secondary creep
rate (Figure 12) showed that there seemed to be a difference in creep behaviour at low stress
and high stress. Using the Larson-Miller time-temperature parameter, the experimental data
were evaluated and the 100 000 h stress rupture strengths at 600 and 650°C were calculated
firstly using all the data and secondly using only data for rupture times above 3000 h. and the
extrapolated 100 000 h stress rupture strengths are shown in Figure 19. After 3000 h at 600
and 650°C, the dislocation density of P92 had reached a more or less constant value. It can be
seen from Figure 19 that the extrapolation using only data from the longer duration tests gave
significantly lower 100 000 h rupture strengths at both temperatures.

The Larson-Miller assumption that the secondary creep rate ε&s was inversely proportional to
the rupture time t R was further refined by Monkman and Grant 1956 to give the Monkman-
Grant (MG) equation

ε&s m t R = K

1

where K1 and m are constants. Rearranging, we obtain


1
logε&s = − log t R + log K1
m

Plotting log rupture time as a function of log secondary creep rate therefore gives a straight
line from which K1 and m can be determined. The constant m is usually in the range 0.8 - 1.2.
The MG equation together with the Norton equation can be used for extrapolation. From the
MG plot the secondary creep rate for a given rupture life can be read off and from the Norton
equation plot, the stress which gives rise to this creep rate can be determined. The Monkman-
Grant plots for P91, P92 and E911 are shown in Figure 20 and indicate that a rupture life of
100 000 h at 600°C corresponds to a secondary creep rate of 1.5 x 10-5 %/h. From Figure 12,
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 22

this creep rate is obtained at a stress of 110 MPa for P92, 105 MPa for E911 and 85 MPa for
P91. The values obtained for P92 agree well with the Larson-Miller extrapolation made using
only data of above 3000 h duration.

130

Extrapolations using the Larson-Miller


time-temperature parameter for P92 steel
100 000 h rupture stress, MPa

120

110

100
23-32000 h 3000-32000 h 5000-32000 h
range of rupture times in evaluation

Figure 19: Extrapolated 100 000 h rupture strengths at 600°C for P92 calculated with data
from different rupture time ranges

10
E911
secondary creep rate, %/h

1
P92
0.1
P91
0.01

0.001

0.0001

0.00001
10 100 1000 10000 100000
rupture time, h

Figure 20: Monkman-Grant equation plots for P91, P92 and E911 steels
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 23

5. Steam oxidation resistance of high chromium steels

Almost without exception, high temperature materials rely on the selective oxidation of one or
more alloy constituents to form a protective oxide scale. Two conditions need to be fulfilled;
firstly, there must be a sufficiently high concentration of the selectively oxidising elements in
the matrix and secondly, the diffusion rate of these elements must be fast enough to ensure
that they replenish the matrix below the growing scale, thus ensuring long-term protection. In
the high Cr steels, clearly Cr is the most important constituent with regard to oxidation
resistance.

In the development of the modified 9Cr1Mo steels, the emphasis was on improvements in the
stress rupture strength. Long-term creep tests carried out at temperatures up to 650°C had
shown that the oxidation resistance in air was excellent due to the formation of tightly
adherent, protective oxide scales. The protective scales formed in air on 9% Cr steels were
identified as consisting of (Fe,Cr)2O3 and (Fe,Cr,Mn)3O4. However, in steam-containing
atmospheres, the scales formed at 600 and 650°C were thick and consisted of an external
Fe3O4 scale and an internal two-phase scale of Fe3O4 and (Fe,Cr,Mn)3O4 (Ehlers and
Quadakkers 2001). Below the oxide scale, internal oxidation of Cr to form (Fe,Cr,Mn)3O4
occurred at the martensite lath boundaries. Figure 21 shows typical oxide scales formed on the
9% Cr steels in steam-containing atmospheres at 650°C and Figure 22 summarises the mass
changes that occur.

Ni coating

Fe3O4

(Fe, Cr, Mn) spinel

FeO + Cr2O3 stringers


steel matrix

Figure 21: Scale microstructure of P91 exposed for 1 000 h at 650°C in Ar-50% H2O

There are several concerns about the high oxidation rates seen in steam. Firstly, the loss in
load-bearing cross-section due to the oxide scale formation and the internally oxidised zone
will lead to stress increases and therefore a reduction in service lives. The reduction in life is,
of course, dependent on the initial wall thickness of the components under consideration.
Calculations reported by Quadakkers and Ennis 1998 have shown that the life reduction at
600°C in steam will be significant for components with wall thicknesses below about 6 mm.
A second, and more difficult to quantify, effect is the spalling of the oxide. The presence of
spalled oxide particles in the steam entering the turbine can cause erosion problems and local
blockages. Furthermore thick oxide scales on heated tubes can lead to a decrease in the heat
transfer across the tube walls, resulting in overheating and subsequent creep failure of tubes.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 24

40
650°C
P92
open points - air
30 closed points - Ar + 50% H2O
mass change, mg.cm-2

20
P91

10
12Cr1MoV (low Cr)
12Cr1MoV (high Cr)
0
P91, air AISI 316LN
12Cr1MoV (high Cr), air
-10
0 500 1000 1500 2000 2500 3000 3500
exposure time, h

Figure 22: Mass change data for Cr steels exposed in Ar-50% H2O and in air at 650°C

In order to achieve acceptable oxidation resistance at 600 and 650°C, according to Ehlers et al
2001 a Cr content of at least 11% is required, to enable the formation of a protective spinel
scale. The oxidation resistance may be enhanced further by the addition of selectively
oxidisable elements, the main contenders being Si and Mn. Si is, however, a ferrite former,
and when added to the steel in amounts that are clearly beneficial for the steam oxidation
resistance can lead to toughness and fabrication difficulties.

There are some additional factors that need to be considered in assessing the steam oxidation
resistance of the high Cr steels. The effect of steam test parameters, such as flow rate and
pressure, are not sufficiently well established and there is a need for further investigations to
ensure that laboratory data reflects sufficiently well the expected behaviour in a power plant.
Such tests are being carried out in the new COST 522 Action, see Allen et al 1998.

The steam oxidation resistance of the high Cr steels is enhanced by:


• Cr content of at least 11%;
• the addition of oxygen active elements, such as Si and Mn;
• increasing the diffusion of Cr to the surface by suitable bulk alloying additions or by
surface deformation treatments

6. Potential for further development

The need to obtain the optimum microstructure for high creep rupture strength and the
requirement for improved steam oxidation resistance make contradictory demands on the steel
composition. For satisfactory creep and stress rupture strength, Cr contents of around 9-10%
allow the desired fully martensitic microstructure to be obtained. For adequate resistance to
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 25

steam oxidation, Cr contents above 11% are necessary. The aim of the current steel
developments is to raise the Cr content to 11-12%, and to add austenite stabilising elements to
produce the fully martensitic structure. In this way, it is hoped that similar stress rupture
strength levels to the 9% Cr steels can be reached at higher Cr contents in the range 11-12%.
Some success has been reported by Tsuda et al 1998, but long-term creep data are not yet
available and extrapolations are therefore uncertain.

A promising line of development being followed in the new COST 522 Action is the addition
of Co, which enlarges the austenite field in the composition direction with no decrease in the
α/γ transition temperature, to 11-12% Cr steels (Allen et al 1998). The disadvantage of Co is,
of course, the high cost of the raw material. Additions of V and Nb are necessary for
precipitation hardening of the matrix. Regarding the solid solution strengthening additions of
Mo and W, there are two approaches; in the first, additions of Mo alone are being
investigated, in order to prevent the formation of Laves phase. There is, however, sufficient
evidence to suggest that W additions do confer improved stress rupture strength, at least to the
test durations of around 50 000 h that have been achieved in creep testing.

There are other development routes towards a ferritic, high strength Cr steel with good
resistance to steam oxidation. The effectiveness of increased amounts of nitride precipitates in
conferring high stress rupture strength has been investigated by Goecmen et al 1998. The
steels under development contain increased amounts of V and N and are given an appropriate
heat treatment to produce finely dispersed nitrides. A similar development has been reported
by Pugh 1998, the strengthening particles being in this case TiN, introduced by a solid state
reaction between CrN and the base steel containing 12% or more Cr and 1-2% Ti. The Osprey
process and a powder metallurgical production route have been investigated. Another idea
under investigation is the addition of FeWTi carbide powder to steel melts with the aim of
producing a uniform and fine dispersion of Ti and W carbides in the steel matrix, as described
by Nutting 1999.

The application of surface treatments and coatings is a promising, innovative line of


development. The coating of large components and small diameter tubes does, however,
present a considerable technological and economic challenge.

The aim of the current developments is to combine in one steel the strength of the 9% Cr
steels and the steam oxidation resistance of the 12% Cr steels, for the temperature range up to
625°C. This target appears to be achievable. However, the extension of the operating
temperature range of the steels to 650°C will require an further increase in creep strength; it is
by no means clear whether such an increase can be achieved either by enhancing the already
established mechanisms of strengthening or by the new concepts being considered.

Acknowledgements

The support of the Polish State Committee for Scientific Research and German BMBF is
gratefully acknowledged. The authors also wish to thank their colleagues, H J Penkalla and W
J Quadakkers, for their valuable contributions to the work described in this paper.
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 26

References

Allen D, Oakey J and Scarlin B, The New COST Action 522 – Power Generation in the 21st
Century: Ultra Efficient, Low Emission Plant, in Materials for Advanced Power Engineering
1998, ed J.Lecomte-Beckers, F.Schubert, P.J.Ennis, ISBN 389336 228-2, Forschungszentrum
Julich, Energy Technology Series, Volume 5, 1998, Part III, 1825-1839

ASTM Standard A335-P91 Standard Specification for Seamless Ferrite Steel Pipe for High
Temperature Service, American Society for Testing of Materials, 1986

Canonico D: Proc. Second Int. Conf. Interaction of Steels with Hydrogen in Petroleum
Industry Pressure Vessel and Pipeline Service, 19-21 October 1994, Vienna; Vol 2, pp 607-
618

Czyrska-Filemonowicz A, Penkalla HJ, Zielinska-Lipiec A and Ennis PJ: Proc. 9th Int. Conf.
on Creep Resistant Metallic Materials, 8-10.04.2001, Prague, Czech Republik, J.Purmensky
(ed.), pp.204-212

Eggeler G, Nilsvang N and Ilschner B: Steel Research 2 (1987), pp. 97-103

Ehlers R J and Quadakkers W J, Oxidation von ferritischen 9-12% Cr-Staehlen in


wasserdampfhaltigen Atmosphaeren bei 550 bis 650°C, Doctor thesis published as report of
the ResearchCentre Juelich Juel-3883, ISSN 0944-2952, June 2001

Ehlers R J, Ennis P J, Singheiser L, Quadakkers W J and Link T, Significance of Scale


Spalling for the Life Time of Ferritic 9-10% Cr Steels During Oxidation in Water Vapour at
Temperatures Between 550 – 650°C, Workshop ‘Life Time Modelling of High Temperature
Corrosion Processes, EFC Event, 22-23 February 2001, Frankfurt

Ennis P J and Czyrska-Filemonowicz A: Proc. XVI Conf. on Advanced Materials and


Technologies, AMT’2001, Gdansk-Jurata, 16-20.09.2001, Inzynieria Materialowa, pp.311-
318

Ennis P J, Zielinska-Lipiec A and.Czyrska-Filemonowicz A: Materials Science and


Technology, October 2000, Vol 16, pp 1226-1232

Ennis P J, Zielinska-Lipiec A, Wachter O and Czyrska-Filemonowicz A: Acta Materialia, Vol


45, No 12, Nov 1997, 4901-4907

Ennis P J, Zieliñska-Lipiec A and Czyrska-Filemonowicz A: Proc. 5 Int. Charles Parsons


Turbine Conf. on Advanced Materials for 21st Century Turbines and Power Plants, 3-
7.07.2000, Cambridge, UK, Strang A et al (eds), IOM Communications Ltd, the Institute of
Materials, 2000, pp 498-507

Foldyna V, Kubon Z, Jakobová A and Vodárek V: in Proceedings of Ninth International


Symposium on Creep Resistant Metallic Materials, Hradec nad Moravicí, Czech Republic,
23-26 September 1996, pp 203-216

Göcmen A, Uggowitzer P J,Solenthaler C, Speidel M and Ernst P, Alloy Design for Creep
Resistant Martensitic 9-12% Chromium Steels, in Microstructural Stability of Creep Resistant
Alloys for High Temperature Applications, ed A Strang, J Cawley G W Greenwood, Institute
of Materials, ISBN 1 86125 045 2, 1998, 311-322
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 27

Hald J, Straub S and Foldyna V: in Materials for Advanced Power Engineering 1998, ed
J.Lecomte-Beckers, F.Schubert, P.J.Ennis, ISBN 389336 228-2, Forschungszentrum Julich,
Energy Technology Series, Volume 5, 1998, Part I, pp. 171-189

Hättestrand M and Andrén H O: Evaluation of particle size distribution of precipitated in a


9% Cr steel using EFTEM, accepted for publication in Micron

Monkman F C and Grant N J, ASTM Proceedings, vol 56, 1956, pp 593-620

Nippon Steel Technical Report No. 50, July 1991

Norton F H, ‘The creep of steel at high temperatures’, McGraw-Hill, New York, 1929.

Larson F R and .Miller J, A time-temperature relationship for rupture and creep stresses,
Trans ASME, 74, 1952, 765-775

Nowakowski P, Straube H, Spiradek K and Zeiler G: in Materials for Advanced Power


Engineering 1998, ed J.Lecomte-Beckers, F.Schubert, P.J.Ennis, ISBN 389336 228-2,
Forschungszentrum Julich, 1998, Energy Technology Series, Volume 5, Part I, pp.569-576

Nutting J, The Long Term Structural Stability of Power Generation Steels – Some Basic
Considerations, in Advanced Heat Resistant Steel for Power Generation, ed R Viswanathan, J
Nutting, Institute of Materials ISBN 1 86125 079 7, 1999, 12-30

Pugh J, A New Titanium Nitride Dispersion Strengthened Ferritic Steel for High Temperature
Applications, in Materials for Advanced Power Engineering 1998, ed J.Lecomte-Beckers,
F.Schubert, P.J.Ennis, ISBN 389336 228-2, Forschungszentrum Julich, Energy Technology
Series, Volume 5, 1998, Part I, 471-480

Quadakkers W J and Ennis P J, The oxidation behaviour of ferritic and austenitic steels in
simulated power plant service environments, in Materials for Advanced Power Engineering
1998, ed J.Lecomte-Beckers, F.Schubert, P.J.Ennis, ISBN 389336 228-2, Forschungszentrum
Julich, Energy Technology Series, Volume 5, 1998, Part I, 123-138

Schaffernak B, Hofer P and Cerjak H: in Materials for Advanced Power Engineering 1998, ed
J.Lecomte-Beckers, F.Schubert, P.J.Ennis, ISBN 389336 228-2, Forschungszentrum Julich,
Energy Technology Series, Volume 5, 1998, Part I, pp 521-530

Sikka VK, Ward CT and Thomas KC: Modified 9Cr-1Mo Steel – in Proceedings of
Conference 'Ferritic Steels for High Temperature Applications,' Warren, PA, USA, 6-18
October 1981; ASM 1983, pp 65-84

Staubli M, Bendick W, Orr J, Deshayes F and Henry Ch: in Materials for Advanced Power
Engineering 1998, ed J.Lecomte-Beckers, F.Schubert, P.J Ennis, ISBN 389336 228-2,
Forschungszentrum Julich, Energy Technology Series, Volume 5, 1998, Part I, pp.87-104

Strang A and Vodarek V: in Materials for Advanced Power Engineering 1998, ed J.Lecomte-
Beckers, F.Schubert, P.J Ennis, ISBN 389336 228-2, Forschungszentrum Julich, Energy
Technology Series, Volume 5, 1998, Part I, pp.603-614

Summary of Average Stress Rupture Strengths of Wrought Steels for Boilers and Pressure
Vessels, ISO Technical Report ISO/TR 7468-1981
Creep resistant steels for power plant OMMI (Vol 1, No.1) April 2002 28

Tsuda Y, Yamada M, Ishii R, Watanabe O and Miyazaki M, Newly Developed 12%


Chromium Heat Resistant Steels for Steam Turbines, in Materials for Advanced Power
Engineering 1998, ed J.Lecomte-Beckers, F.Schubert, P.J.Ennis, ISBN 389336 228-2,
Forschungszentrum Julich, Energy Technology Series, Volume 5, 1998, Part I, 341-350

Vanstone R W: in Materials for Advanced Power Engineering, ed J Lecomte-Beckers, F


Schubert, P J Ennis, Forschungszentrum Jülich, Energy Technology Series, Volume 5, Part I,
139-154, ISSN 1433-5522, ISBN 3-89336-228-2

W J Quadakkers and P J Ennis, The oxidation behaviour of ferritic and austenitic steels in
simulated power plant service environments, in Materials for Advanced Power Engineering,
ed J Lecomte-Beckers, F Schubert, P J Ennis, Forschungszentrum Jülich, Energy Technology
Series, Volume 5, Part I, pp 123-138, ISSN 1433-5522, ISBN 3-89336-228-2

Wachter O, Ennis P J, Czyrska-Filemonowicz A, Zielinska-Lipiec A and Nickel H: Report of


the Research Centre Jülich, Jül-3074, ISSN 0944-2952 (1995).

Zieliñska-Lipiec A, Czyrska-Filemonowicz A and Ennis P J: Journal of Materials and


Processing Technology, 64(1997)3997-405

Você também pode gostar