Martensuta Temperada

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Tempered Martensite

Tempered martensite embrittlement is associated with the formation of coarse


cementite which reduces toughness in such a high strength matrix.

From: Phase Transformations in Steels: Fundamentals and Diffusion-Controlled


Transformations, 2012

Related terms:

Bainite, Ferrite, Heat Treatment, Martensite, Pearlite, Microstructure, Hardness,


Cementite

View all Topics

Stress corrosion cracking (SCC) in aero-


space vehicles
R.J.H. Wanhill, ... C.L. Smith, in Stress Corrosion Cracking, 2011

Background
High strength low alloy steels have a tempered martensite microstructure. The
degree of tempering determines the strength range. These steels are used mainly
in mechanical systems in aircraft, notably for landing gear and gearbox components
and high strength bolts and fittings. The alloys include the AISI grades 4330, 4330M
and 4340, and 300M, D6ac and H11. All are susceptible to SCC, and also hydrogen
embrittlement (Wanhill et al., 2008), at yield strengths above 1200 MPa; and they are
extremely susceptible at yield strengths above 1400 MPa. This is why the guidelines
in Table 16.3 advise restricting the ultimate tensile strength (UTS) to less than
1400 MPa. However, exceptions are made for landing gear, among other items, as
is also noted in Table 16.3.

It is important also to note that SCC in high strength steels involves hydrogen
embrittlement due to hydrogen generated at the crack tips, and that this can occur
in moist air as well as aqueous environments. The SCC fracture characteristics are
also similar – if not identical – to those of internal hydrogen embrittlement (IHE),
which is due to the presence of solute hydrogen in the steel.
> Read full chapter

Martensitic steels for cast components


in ultra-supercritical power plants
S. Roberts, in Materials for Ultra-Supercritical and Advanced Ultra-Supercritical
Power Plants, 2017

5.4.2 Heat treatment and microstructural control


The final microstructure of the 9%Cr steels is tempered martensite with controlled
delta ferrite. Delta ferrite is detrimental to creep life and is avoided by chemistry
balance and heat treatment parameters. Like supercritical grades, 9%Cr steels re-
quire an initial austenitization followed by rapid cooling, usually using forced air
blasting. The hardened material is then tempered (Fig. 5.7) to achieve a microstruc-
ture of tempered martensite, resulting in a material with an excellent balance of
strength while maintaining acceptable levels of room-temperature toughness. Table
5.2 shows the typical room mechanical properties that are achieved with 9%Cr steel
castings.

Figure 5.7. Schematic quality heat treatment for modified 9%Cr steels.

9%Cr steel castings often receive a double temper treatment. The martensite finish
(MF) temperature is relatively low compared to Cr-Mo and Cr-Mo-V steels, and this
needs to be taken into account before tempering and post-weld heat treatment are
performed.

> Read full chapter

FAILURE ANALYSIS OF CARRIER


CHAIN PINS
G.A. SLABBERT, ... R. PATON, in Failure Analysis Case Studies II, 2001
2.3 Metallographic analysis
The metallurgical structure of all the samples, except for pin C, was tempered
martensite and pin C contained ferrite stringers in a tempered martensite matrix,
as shown in Figs 2 and 3. Since all the pins failed in a similar manner, the presence
of ferrite stringers in pin C did not influence its failure.

All of the fractures, of which the example in Fig. 4 is typical, are intergranular.
They exhibit features, including yawning grain boundaries, micropores and no grain
boundary corrosion, which indicate that they were caused by hydrogen embrittle-
ment (HE) (Fig. 5) [2]. A small section of a fracture surface can be seen at the bottom
of both Figs 2 and 4. In Fig. 2, the intergranular nature of the fracture, as well as the
crack running parallel to the main fracture surface, confirms that the ferrite stringers
did not contribute to the propagation of the crack, and thus the failure of the pins.

Fig. 4. The tempered martensite matrix as found in all the pins, except pin C, whose
structure is shown in Figs 2 and 3. Note the intergranular nature of the fracture
surface on the bottom as well as the machining marks on the right.
Fig. 5. An SEM picture of the fracture surface of a pin showing the intergranular na-
ture of the fracture. Micropores, yawning grains, and the absence of grain boundary
corrosion, characteristic features of HE, are shown.

The surface on the right-hand side of the pin in Fig. 4 shows the original machining
marks on the outside of the pin. This confirms that little corrosion had taken place
on the pin.

> Read full chapter

Modelling high temperature creep of


academic and industrial materials us-
ing the composite model
M. Meier, W. Blum, in Fundamental Aspects of Dislocation Interactions, 1993

4 Modelling strain softening in X 20 CrMoV 12 1


It has been shown [4, 5] that during the long-term creep of the tempered martensite
steel X 20 CrMoV 12 1 (German notation) the carbide particles coarsen and the
subgrains grow while the corresponding curves exhibit a rather dramatic degree
of softening. The full curve in Fig. 5 has been obtained by eliminating the effect of
damage by cavities on on the basis of comparison between tensile and compressive
tests [5]. Consequently, the increase in in the tertiary range of creep after the
minimum of must be caused by changes in the structures of dislocations and
carbides.

Fig. 5. Modelling of strain softening during the long-term creep of the ferritic steel
X 20 CrMoV 12 1 at 823 K due to subgrain growth only (chain curve) and to a
combination of subgrain growth and particle coarsening (dotted curve). The solid
curve represents the experimental results corrected for the damage due to cavities
[5].

Since both X 20 CrMoV 12 1 and NiCr22Col2Mo are hardened by solutes and


carbides, the kinetic laws for the steel were chosen similar to the nickel-based
alloy. In contrast to the nickel-based alloy, the volume fraction fsub was equal to
1 in the initial state of the ferritic steel (owing to martensitic transformation plus
tempering). This eliminates the first term in eqn. (9). The kinetic law for the soft
region was defined by eqns. (5a) and (10) with = 0.075 MPa−1,B=6.14 × 10−16 ms−1
MPa−1 and G, =0.25 . Since carbide particles were preferentially associated with
subgrain boundaries, particle hardening in the subgrain boundary regions could not
be neglected and had to be assumed to be stronger than in the soft region. Thus h
in eqns.(6) was replaced by h– p, h. The estimates of h– p, hand h– p, s were made
on the basis of an Orowan stress using the measured volume fraction and size of the
particles(from ref. 5). Consequently, the particle stress components decrease with
time as the particles coarsen(from p, h,0 = 200 MPa at = 0 to p, h= 147 MPa at
= 0.08 in the test of Fig. 5; the time dependence of p, s= 25 MPa was neglected).
The measured values of subgrain size [5] were used and the activation area Δah in
the subgrain boundary was estimated to be 57b2.. The other parameters took their
standard values.

Figure 5 shows the results of the modelling for two cases. In the first case the
decrease in the particle hardening was neglected by setting p, h= p, h,0; in the second
case the decrease in p, h was taken into account. In both cases the initial decrease in
was modelled by introducing an initial build-up of G, It is seen that the growth of
subgrains causes a small increase in with (chain curve in Fig. 5). Particle coarsening
enhances this increase (dotted curve in Fig.5). Yet the combination of subgrain
and particle effects chosen in the modelling does not lead to agreement between
the modelled and measured curves. Agreement might be achieved by changing
the model, e.g. with regard to the parameters for particle hardening. However,
there are experimental results [6] indicating that an additional process of dynamic
(strain-dependent) softening comes into play, e.g. growth of the dislocation spacing
s in subgrain boundaries (decrease in subgrain misorientation) may be responsible
for the balance of softening.

These results have significance in understanding the technically important mini-


mum of the creep rate. Note that the proposed explanation works without consider-
ing the effects of damage by cavities, which become important only in the late stages
of tertiary creep during the last 15% of the lifetime [5].

> Read full chapter


Failure analysis cases of components of
automotive and locomotive engines
Zhiwei Yu, Xiaolei Xu, in Handbook of Materials Failure Analysis with Case Studies
from the Aerospace and Automotive Industries, 2016

3.2.4 Microstructure Examination


The longitudinal section of the plunger-sleeve was prepared for metallurgical ex-
amination. The core microstructure is composed of tempered martensite and fine
elongated MnS inclusions were seen along the axial direction of the plunger-sleeve
(Figure 17.15a). The carburized layer is composed of martensite and retained austen-
ite (Figure 17.15b and c), which is normal microstructure of the grade of the steel. It
was noted that the spear-like openings at the edge of oil-hole were always aligned
to MnS stringers (Figure 17.15b and c). In other words, the spear-like openings
propagated along MnS-inclusion stringers.

Figure 17.15. Microstructure of failed plinger-sleeve: (a) core, (b, c) surface layer
region of oil-hole.

> Read full chapter

Mechanical behavior of structural mate-


rials for Generation IV reactors
M. Sauzay, in Structural Materials for Generation IV Nuclear Reactors, 2017

6.8 Conclusions and recommended further work


The design of thermal power plants and new-generation nuclear reactors has been
the reason for carrying out many studies on the behavior of tempered martensitic
steels and austenitic stainless steels subjected to fatigue and/or creep at high tem-
perature (450–650°C). This chapter reviews firstly the numerous recent experimental
and simulation works concerning tempered martensite-ferritic steels. Then, creep
and fatigue properties of the two steel families are compared on both micro- and
macroscales. Finally, recommended further works are mentioned.

Concerning tempered martensite-ferritic steels, the instability of their microstruc-


ture under certain loading conditions (fatigue, creep) has been demonstrated in
the literature since the 1980s. This instability leads to softening that needs to be
understood and predicted. More recently, the effect of loadings corresponding to
in-service conditions applicable to components of thermal or new-generation nu-
clear power plants has been studied. They correspond to low cyclic strain amplitudes,
long creep holding times and low-stress creep. Obviously, experimental data are
relatively few. Creep times of up to 200 kh at 500–600°C were explored (about
20 years).

Significant cyclic softening is demonstrated even for cyclic amplitudes correspond-


ing to significantly lower stresses than the conventional yield stress. No saturation
is observed, even if the softening rate slows down during cyclic loading. A specimen
previously tested to several million cycles at very low amplitude then deforms in
creep 100 times faster than the as-received material. Cyclic softening is particularly
pronounced as amplitude or hold time are high. Low strain rates trigger softening.
Observations in transmission electron microscopy demonstrate the variation in the
initial tempered martensitic microstructure characterized by a small subgrain size
( 0.5 μm) and a high dislocation density ( 2 × 1014 m−2) in the as-received condition.
Growth in the subgrain size (factor two to three) and a reduction in dislocation
density (factor two to four) are observed. The effects of loading parameters discussed
above are also demonstrated at the microstructure scale.

The Manson–Coffin curve, viscoplastic strain amplitude–number of cycles to frac-


ture in pure fatigue, largely unifies experimental points for different martensitic
steels and temperatures. Oxidation plays an important role, particularly if a hold
time in compression is applied (repeated fracture of oxide layers).

Long-term creep tests (10–20 year lifetimes at 500–600°C) are characterized by a


fairly limited secondary stage followed by a long tertiary stage due to the strain
rate acceleration induced by the progressive softening of the material. Microscopy
observations demonstrate changes similar to those described in cyclic loading. An
even more pronounced drop in the dislocation density and precipitation is observed
for long creep lifetimes.

The Monkman–Grant plot (minimum strain–time to fracture) provides a mean


of combining results obtained for many martensitic steels within the 500–700°C
temperature range and for lifetimes varying from 1 min to 10–20 years on a single
curve, representing seven orders of magnitude in lifetime. Precycling or preaging
with a sufficiently long duration induce much higher creep rates than for the
as-received material due to material softening. Nevertheless, the fracture points
are always positioned satisfactorily on the Monkman–Grant master curve plotted
for the as-received material. Fracture takes place by necking that is only visible
macroscopically during the last few percents of lifetime. Necking is simulated by a
mechanical instability modeling in the viscoplasticity framework that takes account
of softening of materials under load. Predictions agree well with test results and
explain the existence of this master curve. Intergranular creep cavitation is observed
for longer lifetimes (100 kh at 600°C) but has little effect on the predictive nature of
the Monkman–Grant law that relates the minimum strain rate and lifetime, at least
for the longest known tests.

Polycrystalline homogenization is proposed based firstly on annihilations between


mobile dislocations and secondly on annihilations between mobile dislocations and
dislocations of sub-boundaries. Predicted variations in microstructure and macro-
scopic mechanical behavior during loading are in qualitative agreement with many
measurements and observations.

Because their as-received condition microstructure differs strongly from that of


tempered martensite-ferritic steels, the stress-strain behavior of austenitic stainless
steels differs strongly from that of martensitic steels. During creep and cyclic de-
formation with and without hold time, dislocation production and microstructure
are observed, which lead to hardening instead of softening. As creep strain rates
in martensitic steels are usually higher than in austenitic stainless steels, necking is
triggered in martensitic steels, which leads to a considerable delay in the observation
of the detrimental effects of grain boundary cavitation. For instance, a change in
the creep lifetime–stress plot is observed from 1 year in the 316L(N) steel at 600°C
whereas the effect is negligible after 10 years in the grade 91 steel. Similarly, the
detrimental effect of hold time is essentially due to tension hold and intergranular
creep cavitation in austenitic stainless steels, in contradiction with the observations
carried out in martensitic steels.

Nevertheless, the review of long-term creep test results and observations shows
that intergranular creep damage leads to considerable overestimations of lifetimes
based on the collection of short-term data. The modeling of short-term necking
and long-term intergranular damage allows predictions in fair agreement with
the existing data, regardless of the material, temperature, and load. In both steel
families, a transition in deformation mechanisms is observed too, as shown by the
change in slope in the Norton plot of the minimum strain rate depending on stress.
Accounting for the low-stress regime of strain rates allows lifetime predictions in
fair agreement with experimental data up to 25 years, in all considered steels and
for temperatures from 500 up to 750°C. Only the cavity nucleation rate should be
experimentally evaluated thanks to cavity counting using FEG-SEM.

Interestingly, even if the microstructural evolutions are different, advanced


austenitic stainless steels and the Incolloy 800 alloy, which is close to nickel-based
alloys, are subjected to the same creep damage mechanisms as martensitic steels
and conventional austenitic stainless steels. The same modelings may be applied
and lead to creep lifetime predictions in agreement with experimental data up to
the longest experimental lifetimes published in the literature.

Although these results provided by numerous and wide research programs provide
a better understanding of deformation and damage mechanisms of tempered
martensite-ferritic steels and austenitic stainless steels at high temperature, several
mechanisms need to be better understood and modeled, having in mind in-ser-
vice conditions and the design of components subjected to long-term creep and
fatigue-relaxation:

• How can the low-stress creep strain rate regime be explained? Additional tests
and microscopic observations are required, alongside a predictive model. The
dislocation density evolution will undoubtedly have to be taken into account
in such simulations;
• For predicting creep lifetimes up to 60 years, physically based modeling of
creep cavitation is required in addition to long-term test results. One of the
main open problems is the understanding and prediction of cavity nucleation.
As shown in this review, the measurement of the cavity nucleation coefficient
used in the Dyson law allows fair predictions of lifetimes up to 25 years.
Further work is then required concerning long-term precipitation and cavity
nucleation at grain boundary second-phase particles, along with observations
carried out on long-term specimens required for validating the modeling
predictions not only at the macroscopic cale but also at the microscopic cale;
• Fatigue in-service conditions correspond to small strain amplitudes and long
hold times which leads to the requirement of modeling efforts too. The
stress-strain behavior depends on many parameters such as strain amplitude,
temperature, hold time, etc., the hardening and recovery mechanisms should
be carefully characterized to allow their modeling and finally predictions;
• Such predictions may provide inputs for fatigue-relaxation damage modeling,
which should be based on the synergy between oxidation and oxide layer
fracture in tempered martensite-ferritic steels but creep cavitation in austenitic
stainless steels.

Long-term creep and fatigue-relaxation tests are essential for providing macroscopic
data and allowing the observation of the long-term dislocation microstructure and
damage evolutions.

> Read full chapter

Strengthening Mechanisms in Steels


G. Krauss, in Encyclopedia of Materials: Science and Technology, 2001

4 Martensite Strengthening in Medium- and High-carbon


Steels
Martensite is a crystal structure and morphology formed by rapidly cooling or
quenching austenite. The rapid cooling prevents carbon atoms from diffusing, and
therefore the crystal structure of martensite is body-centered-tetragonal because
carbon atoms are trapped between iron atoms and elongate the body-centered
unit cell in one direction. In contrast, as noted above, in steels slow-cooled from
the austenite, carbon atoms diffuse away from ferrite, and form microstructures
consisting of combinations of b.c.c. ferrite and cementite. The latter structures
are lower energy and more stable than martensite. Since diffusion is suppressed,
martensite crystals are formed by shear mechanisms of transformation. Not only
is the composition of the martensite the same as the parent austenite, as it must
be in view of the absence of diffusion, but also a very high density of dislocations
and/or fine twins are introduced into the martensite crystals to accommodate the
shear mechanisms of transformation and volume changes between austenite and
martensite.

As-quenched martensite tends to have low toughness, and is therefore heated, a


process referred to as tempering, to produce more fracture-resistant microstruc-
tures (Krauss 1990b). There are many microstructural changes that develop with
increasing tempering temperature, but the most important are those associated with
the precipitation and growth of carbides to relieve the supersaturation of carbon in
the martensite. Thus, carbon diffuses from interstitial sites in the martensite, and
the crystal structure reverts to b.c.c. ferrite and dispersions of carbide particles. With
increasing tempering temperature, more and more stable and coarser mixtures of
ferrite and carbides are formed within as-quenched martensitic microstructures.
4.1 Martensitic Strengthening Mechanisms
Figure 1 shows that the highest hardness of any steel microstructure is found in
as-quenched and in low-temperature-tempered martensite, and that the hardness
increases significantly with increasing carbon content. The very high strength of
martensite is a result of two interrelated strengthening mechanisms. The one mech-
anism is the very high rate of strain hardening that develops dynamically from the
high-dislocation-density substructure of martensite crystals. The other is dispersion
strengthening contributed by the carbides which precipitate in martensite during
tempering.

Figure 14 shows the mechanical properties of martensitic low-alloy, carbon steels


tempered at 150 °C as a function of carbon content (Krauss 1995). The separation
between the yield strength (YS) and ultimate tensile strength (UTS) show that
high rates of strain hardening, increasing with carbon content, develop in these
microstructures. The higher rates of strain hardening in the higher-carbon steels
are in part related to increasing densities of dislocations and carbide particles
precipitated in higher-carbon-content martensite. Strain hardening is also increased
by deformation-induced transformation of austenite that has not transformed dur-
ing quenching. Retained austenite increases in martensitic microstructures with
increasing carbon content.

Figure 14. Mechanical properties of low-alloy carbon steels quenched to martensite


and tempered at 150 °C; (a) strength properties, (b) ductility properties (Krauss 1995).

The various measures of ductility, except for uniform elongation, decrease with
increasing carbon content because the higher ultimate strengths are close to the
stresses that initiate ductile fracture. In lower-carbon steels, where lower ultimate
tensile strengths are generated by strain hardening, more postuniform elongation
is necessary in order to develop the stresses for ductile fracture. Uniform elongations
increase with increasing carbon content as a result of increased strain hardening in
the higher-carbon martensites, as shown in Sect. 2.2 of this article.

Hardness and strengths of martensitic microstructures decrease with heating at


increasing tempering temperatures as a result of decreasing dislocation densities
and the diffusion-assisted coarsening of the carbide dispersions. As the carbide dis-
persions coarsen, the interparticle spacings decrease, and lower stresses are required
for dislocation motion. At the highest tempering temperatures, around 700 °C,
the tempered microstructure has evolved into a dispersion of coarse, spheroidized
carbides in a matrix of ferrite, and in this condition, as shown in Fig. 1, strength is
the lowest of any microstructure in carbon steels.

> Read full chapter

Boiler steels, damage mechanisms, in-


spection and life assessment
A. Shibli, in Power Plant Life Management and Performance Improvement, 2011

Mo and W
The microstructure of the 9%Cr steels, with Mo, W, Nb and V as important alloying
elements, consists of tempered martensite in which islands of ferrite form during
cooling. The martensite has a typical lath structure within the prior austenite grain
boundaries. The long term stress rupture behaviour of these steels is mainly affected
by:

• solid solution hardening by C, Cr, Mo and W;

• the type, size and distribution of fine precipitates;

• dislocation density within the sub-structure;

• the austenite grain size, although this has a somewhat lesser effect.

Upon cooling from normalising/austenitising and especially during tempering,


carbide formation takes place. The more important carbides that form and their
transformations are listed below:

• M3C → M7C3 → M23C6;

• MC → M2C → M6C;
• various carbonitrides.

The formation of various precipitates at different temperatures is shown in Table 7.3.

Table 7.3. Basic characteristics of minor phases in 9–12%Cr steels [6]

Precipitate Formed during Preferred precipi- Characterised by Main contribution


tation to
M23C6 Tempering Prior austenite Medium growth Precipitation-
grainand M lath rateduring creep strengthening (PS)
boundaries
Nb(C, N) Solidification Austenite grain Restriction of
boundaries grain growth
M2X(Cr2N) Tempering below Inside sub-grain Dissolution dur- Precipitation
700 °C and M laths ing creep at high strengthening
temperatures
Secondary MX(VN) Tempering above Inside sub-grains, High dimension- Precipitation
700 °C creep ex- M laths and/or al stability during strengthening
posure dislocations creep
Cr(V, Nb)N Long term creep M23C6, Nb(C, N) Dissolution of Lowering PS
Z-phase exposure Nb(C,N), M2X and
VN
Laves phase Fe2- Creep exposure Grain and Dissolution above Decrease in SS
(Mo, W) sub-grain bound- 650 °C, fast coars- lowering PS, poor
aries, on M23C6 ening at 600 °C plasticity
M6X Tempering (> Grain and High coarsening Decrease in SS,
1.6% Mo) long sub-grain bound- rate.dissolution of lowering PS, poor
term creep expo- aries on M23C6 M2X, Nb(C, N); plasticity
sure decrease of Mo,
Win SS
AIN Tempering Grain and High coarsening Poor plasticity
sub-grain bound- rate, reduction of
aries N in SS

It is now well known that Mo and W both belong to the same group as Cr and thus
have a similar effect in high Cr steels; i.e. their principal contribution to enhancing
the creep rupture strength is due to their solid solution strengthening. In addition,
they contribute to rupture strength through precipitate hardening by the formation
of carbides and carboni- trides. However, as the atomic weight of W is nearly twice
the atomic weight of Mo, to achieve the same effect of strengthening twice as
much W is needed. It is also known that the low diffusion rate of W leads to a
stabilising effect on the carbides and carbonitrides. Tests in Japan and elsewhere
in the 1960s and 1970s found that the maximum creep strength was obtained for
an Mo equivalent (%Mo + 0.5 × %W) of 1.5. However, because of the increase in
ferrite formation, a decrease in long term rupture strength was observed. The
interesting thing, however, was that high strength was only obtained when Mo and
W were present together. Both in Japan and in European COST programmes 1%Mo
and 1%W were found to be optimal. However, as Mo is a stronger ferrite former
(which weakens the rupture strength and impact strength), the Mo content of 0.5 %
and W of about 2% are more common, as in the case of P92 and HCM12A/P122.
It is also known that with increasing test/creep rupture duration, W and Mo are
also increasingly incorporated to a certain extent into the M23C6 phase. The matrix
therefore becomes depleted in these elements, which are not very effective for
strengthening when present in precipitates compared with when they are dissolved
in the matrix.

In the typical 12%Cr steel X20CrMoV11-1 is the Cr rich M23C6 carbide. The addition
of V, Nb and N in P91 causes additional precipitation of MX particles, where M
consists of V or Nb and X consists of N or C. This precipitate is most stable over the
long term component exposure to creep temperature.

The precipitation reactions during heat treatment have a significant effect on the
long term creep strength of the material. The carbide and nitride particles precipitate
on:

• prior austenite grain boundaries;

• ferrite subgrain boundaries and inside the sub-grains.

> Read full chapter

Laser surface modification of steel


for slurry erosion resistance in power
plants
R.C. Shivamurthy, ... G. Padmanabham, in Laser Surface Modification of Alloys for
Corrosion and Erosion Resistance, 2012

Optical microstructures
The optical microstructure of 13Cr–4Ni steel taken at two different magnifications
is shown in Fig. 7.7. The microstructure consisted of acicular tempered martensite,
which is mainly of fine needle shape martensite.
7.7. Optical microstructure of 13Cr–4Ni steel at (a) low and (b) high magnifications.

During the LSA process, the pre-placed powder paste was melted together with
subsurface of substrate 13Cr–4Ni steel and formed an alloyed (coating) region.
Figure 7.8 shows the optical microstructure of near surface regions of Stellite 6
coatings. It consists of primary dendrites (white phases) and interdendritic eutectic
regions (dark regions). The eutectic regions in Fig. 7.8 consisted of secondary phases
and Co-rich solid solution. Quantitative analysis showed that volume fraction of
Co-phase is about 65% and finer carbides were about 35%. It was very difficult
to measure the sizes of secondary phases in the case of Stellite 6 coatings as the
secondary phases were very fine. However, from previous studies (Tiziani et al., 1987;
Xu and Kutsuna, 2006) it can be concluded that secondary phases of (< 1 μm) are
obtained during LSA process.

7.8. Optical microstructure of Stellite 6 coatings at (a) low and (b) high magnification.

Figure 7.9 shows optical microstructures of Colmonoy 88 coatings obtained at two


different magnifications. It is clearly seen from Fig. 7.9 that there is uniform distri-
bution of characteristic secondary phases throughout the coatings. The quantitative
analysis showed that about 55% of white phase regions and 45% of secondary
phases. Also, secondary phases of 2–10 μm were obtained from LSA.
7.9. Optical microstructure of Colmonoy 88 coatings at (a) low and (b) high magni-
fications.

> Read full chapter

WELDABILITY OF DUPLEX STRUC-


TURE 12Cr-(Mo,W) STEELS
E.J. Vineberg, ... C.C. Clark, in Welding in Energy-Related Projects, 1984

Microstructure
The microstructure of the 12Cr-1.5Mo-1W duplex structure steel in the normal-
ized and tempered condition, shown in Fig 1, consists of ferrite and tempered
martensite. To understand how this microstructure develops, it is useful to consider
a schematic equilibrium phase diagram for the alloy system, as shown in Fig. 2.
The nickel and low carbon contents of these 12Cr-(Mo,W) steels allow ferrite to be
stable at all temperatures up to the liquidus. Thus, when the steels are heated during
processing or in welding, they are never fully austenitic, but rather they contain
varying amounts of austenite. As seen in the phase diagram, at the temperatures
reached in a weld heat affected zone the microstructure is predominantly ferritic
with a relatively small fraction of austenite, whereas at the normalizing temperature,
1050 C (1920 F), the ferrite fraction is 15-25%.
Fig. 1. Microstructure of the normalized and tempered 12Cr-1.5Mo-1W steel (Heat
6455), etched in 1% picric acid and 5% HC1 in methanol.

Fig. 2. Cr-Mo-W steel phase diagram.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like