Presentation Chp4 2

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 63

4.

7 Heat Treatment of Ferrous Alloys


The various microstructures can be modified by heat-treatment techniques—that is, by controlled heating and
cooling of the alloys at various rates. These treatments induce phase transformations, which greatly influence such
mechanical properties as strength, hardness, ductility, toughness, and wear resistance.
The effects of thermal treatment depend on the particular alloy, its composition and microstructure, the degree of
prior cold work, and the rates of heating and cooling during heat treatment.
This section focuses on the microstructural changes in the iron–carbon system. Because of their technological
significance, the structures considered are pearlite, spheroidite, bainite, martensite, and tempered martensite. The
heat-treatment processes described are annealing, quenching, and tempering.
4.7 Heat Treatment of Ferrous Alloys
Pearlite. If the ferrite and cementite lamellae in the pearlite structure of the eutectoid steel, shown in Fig. 4.9, are
thin and closely packed, the microstructure is called fine pearlite; if they are thick and widely spaced, it is called
coarse pearlite. The difference between the two depends on the rate of cooling through the eutectoid temperature,
which is the site of a reaction in which austenite is transformed into pearlite. If the rate of cooling is relatively high
(as in air), fine pearlite is produced; if cooling is slow (as in a furnace), coarse pearlite is produced.

Pearlite is a two-phased, lamellar (or layered) structure composed of alternating


layers of ferrite (87.5 wt%) and cementite (12.5 wt%) that occurs in some
steels and cast irons. During slow cooling of an iron-carbon alloy, pearlite forms
by a eutectoid reaction as austenite cools below 723°C (1,333°F) (the eutectoid
temperature). Pearlite is a microstructure occurring in many common grades of
steels.
Fig. 4.9. Pearlite in 1080 steel,
formed from austenite of eutectoid
composition.
4.7 Heat Treatment of Ferrous Alloys
The eutectoid composition of austenite is approximately 0.8% carbon; steel with less carbon content
(hypoeutectoid) will contain a corresponding proportion of relatively pure ferrite crystallites that do not participate
in the eutectoid reaction and cannot transform into pearlite. Likewise steels with higher carbon content
(hypereutectoid) will form cementite before reaching the eutectoid point. The proportion of ferrite and cementite
forming above the eutectoid point can be calculated from the iron/iron—carbide equilibrium phase diagram using
the lever rule.
Steels with pearlitic or near-pearlitic microstructure can be drawn into thin wires. Such wires, often bundled into
ropes, are commercially used as piano wires, ropes for suspension bridges, and as steel cord for tire reinforcement.
Some hypereutectoid pearlitic steel wires, when cold wire drawn to true strains above 5, can even show a maximal
tensile strength above 6 GPa.
4.7 Heat Treatment of Ferrous Alloys
Spheroidite annealing: hypereutectoid steel with a microstructure of
pearlite and cementite network generally gives poor machinability.
Since cementite is hard and brittle, the cutting tool cannot cut through
these plates. Instead, the plates have to be broken. Therefore, the tool
is subjected to continual shock load by the cementite plates and results
in a ragged surface finish. A heat-treating process which will improve
the machinability is known as spheroidize annealing. This process will
produce a spheroidal or globular form of carbide in a ferritic matrix
as shown in the figure.
Fig. 4.12. Microstructure of eutectoid steel.
Spheroidite is formed by tempering the
steel at 700°C.
4.7 Heat Treatment of Ferrous Alloys
Spheroidite. When pearlite is heated to just below the eutectoid
temperature and then held at that temperature for a period of time,
such as at 700°C for a day, the cementite lamellae transform to
roughly spherical shapes. Unlike the lamellar shapes of cementite,
which act as stress raisers, spheroidites (spherical particles) have
smaller stress concentrations because of their rounded shapes.
Consequently, this structure has higher toughness and lower
hardness than the pearlite structure. It can be cold worked, because
the ductile ferrite has high toughness and the spheroidal carbide
Fig. 4.12. Microstructure of eutectoid steel.
particles prevent the initiation of cracks within the material.
Spheroidite is formed by tempering the
steel at 700°C.
4.7 Heat Treatment of Ferrous Alloys
For spheroidizing, the hypereutectoid steel is held for a long
time at a temperature just below the lower critical
temperature line or is heated and cooled alternatively
between temperatures that are just above and just below the
lower critical line. Prolonged time at the elevated
temperature will completely break up the pearlitic structure
and cementite network. The cementite will become spheres.
The cementite particles and the entire structure is called
spheroidite. This structure is desirable when minimum
hardness, maximum ductility and maximum machinability
are required.
Low carbon steels are seldom spheroidized for machining,
because they are excessively soft and gummy in the
spheoridized conditions. The cutting tool will tend to push
the material rather than cut it, causing excessive heat and
wear on the cutting tip.

If steel is kept too long at spheroidize-annealing temperature, the cementite particles will coalesce and become
elongated, thus reducing machinability.
4.7 Heat Treatment of Ferrous Alloys
Bainite. Visible only through electron microscopy, bainite is a
very fine microstructure, consisting of ferrite and cementite,
similar to pearlite, but having a different morphology. Bainite can
be produced in steels with alloying elements and at cooling rates
that are higher than those required for pearlite. This structure,
called bainitic steel (after the American metallurgist E.C. Bain,
1891–1971), is generally stronger and more ductile than pearlitic
steels at the same hardness level.
4.7 Heat Treatment of Ferrous Alloys
Bainite is a plate-like microstructure that forms in steels
at temperatures of 125–550°C (depending on alloy
content). It is one of the products that may form when
austenite is cooled past a temperature where it no longer
is thermodynamically stable with respect to ferrite,
cementite, or ferrite and cementite. Davenport and Bain
originally described the microstructure as being similar
in appearance to tempered martensite.

A fine non-lamellar structure, bainite commonly


consists of cementite and dislocation-rich ferrite. The
large density of dislocations in the ferrite present in
bainite, and the fine size of the bainite platelets, makes
this ferrite harder than it normally would be.

The hardness of bainite can be between that of pearlite and untempered martensite in the same steel. The fact that it
can be produced during both isothermal or continuous cooling is a big advantage, because this facilitates the
production of large components without excessive additions of alloying elements. Unlike martensitic steels, alloys
based on bainite often do not need further heat treatment after transformation in order to optimize strength and
toughness.
4.7 Heat Treatment of Ferrous Alloys
Martensite. When austenite is cooled at a high rate, such as by quenching in
water, its FCC structure is transformed into a body-centered tetragonal (bct)
structure, which can be described as a body-centered rectangular prism that is
slightly elongated along one of its principal axes.
Because martensite does not have as many slip systems as a bcc structure,
and the carbon is in interstitial positions, it is extremely hard and brittle.
Martensite transformation takes place almost instantaneously, because it
involves not the diffusion process but a slip mechanism and thus involves
plastic deformation. Martensite in AISI 4140 steel

Body centered tetragonal unit cell photomicrograph of 0.35% carbon steel, water-quenched
martensite structure from 870 °C
4.7 Heat Treatment of Ferrous Alloys
Martensite is a body-centered tetragonal form of iron in which some carbon is dissolved. Martensite forms during
quenching, when the face centered cubic lattice of austenite is distorted into the body centered tetragonal structure
without the loss of its contained carbon atoms into cementite and ferrite.
Instead, the carbon is retained in the iron crystal structure, which is
stretched slightly so that it is no longer cubic.
Martensite is more or less ferrite supersaturated with carbon. Compare
the grain size in the micrograph with tempered martensite.

Body centered tetragonal unit cell photomicrograph of


martensite structure
While the unit cell of austenite is a perfect cube, in the
transformation to martensite this cube is distorted so
that it's slightly longer than before in one dimension
and shorter in the other two.
4.7 Heat Treatment of Ferrous Alloys
Martensite:
The difference between austenite and martensite is quite small: The mathematical description of the two structures is
quite different, for reasons of symmetry, but the chemical bonding remains very similar. Unlike cementite, which has
bonding reminiscent of ceramic materials, the hardness of martensite is difficult to explain in chemical terms. The
explanation hinges on the crystal's subtle change in dimension, and the speed of the martensitic transformation.
Austenite is transformed to martensite on quenching at approximately the speed of sound - too fast for the carbon
atoms to come out of solution in the crystal lattice. The resulting distortion of the unit cell results in countless
lattice dislocations in each crystal, which consists of millions of unit cells. These dislocations make the crystal
structure extremely resistant to shear stress - which means, simply that it can't be easily dented and scratched.

Fig. 4.18. (a) Hardness of martensite, as a


function of carbon content.
(b) Micrograph of martensite containing
0.8% carbon. The gray platelike regions
are martensite; they have the same
composition as the original austenite
(white regions). Magnification: 1000X.
4.7 Heat Treatment of Ferrous Alloys
During transformation of austenite to martensite the carbon atoms do not have time to diffuse out of the crystal
structure in large enough quantities to form cementite (Fe3C). As a result of the quenching, the FCC austenite
transforms to a highly strained body-centered tetragonal form of martensite that is supersaturated with carbon. The
shear deformations that result produce a large number of dislocations, which is a primary strengthening mechanism
of steels. The highest hardness of a pearlitic steel is 400 HB, whereas martensite can achieve 700 HB.

The martensitic reaction begins during cooling when the austenite reaches the martensite start temperature (Ms), and
the parent austenite becomes mechanically unstable. As the sample is quenched, an increasingly large percentage of
the austenite transforms to martensite until the lower transformation temperature Mf is reached, at which time the
transformation is completed.
The growth of martensite phase requires very little thermal activation energy because the process is a diffusionless
transformation, which results in the subtle but rapid rearrangement of atomic positions.
Fig: Time-Temperature-Transformation (TTT) diagram for plane carbon eutectoid steel.
4.7 Heat Treatment of Ferrous Alloys
Retained Austenite. Austenite that does not transform to martensite upon quenching is called retained austenite
(RA). If the temperature to which the alloy is quenched is not sufficiently low, only a portion of the structure is
transformed to martensite. The rest is retained austenite, which is visible as white areas in the structure, along with
the dark, needlelike martensite. Retained austenite can cause dimensional instability and cracking, and lower the
hardness and strength of the alloy.

Fig. 4.18. (a) Hardness of martensite, as a


function of carbon content.
(b) Micrograph of martensite containing
0.8% carbon. The gray platelike regions
are martensite; they have the same
composition as the original austenite
(white regions). Magnification: 1000X.
4.7 Heat Treatment of Ferrous Alloys
Retained austenite occurs when steel is not quenched to the M f, or martensite finish, temperature; that is, low enough
to form 100% martensite. Because the Mf is below room temperature in alloys containing more than 0.30% carbon,
significant amounts of untransformed, or retained austenite, may be present, intermingled with martensite at room
temperature.
Retained austenite is a specific crystalline form of iron and steel. The dark-colored needles shown are tempered
martensite crystals and the light-colored areas are retained austenite crystals.

The amount of retained austenite is a function of the carbon content, alloy


content (especially nickel and manganese), quenchant temperature and
subsequent thermal and/or mechanical treatments.
Depending on the steel chemistry and specific heat treatment, the retained
austenite level in the case can vary from over 50% of the structure to nearly
zero. While large amounts of retained austenite (>15%) can be detected and
estimated by optical microscopy, specialized equipment and techniques,
such as x-ray diffraction methods, are required to accurately measure the
amount of retained austenite to as low as 0.5%. Fig. RA present in a case carburized
component. Etchant: 2% nital. 1,000×
4.7 Heat Treatment of Ferrous Alloys
Why is retained austenite problematic?
The very characteristics that give retained austenite many of its unique properties are those responsible for
significant problems in most applications. Austenite is the normal phase of steel at high temperatures, but not at
room temperature. Because retained austenite exists outside of its normal temperature range, it is metastable. This
means that when given the opportunity, it will change or transform from austenite into martensite. In addition, a
volume change (increase) accompanies this transformation and induces a great deal of internal stress in a
component, often manifesting itself as cracks.

How does RA behave?


Martensite is hard, strong and brittle while austenite is soft and tough. In some instances, when combined, the
mixture of austenite and martensite creates a composite material that has some of the benefits of each, while
compensating for the shortcomings of both. For any given application, mechanical properties are affected by a high
percentage of retained austenite content. For example, retained austenite affects the following properties of bearing
steels:
4.7 Heat Treatment of Ferrous Alloys
• Dimensional stability: Retained austenite will transform to martensite if the temperature drops significantly
below the lowest temperature to which it was quenched, or if it is subjected to high stress levels at room
temperature. Martensite, a bct crystal structure, has a larger volume than the FCC austenite that it replaces. Where
transformation occurs, there will be a localized 4-5% increase in the volume of the microstructure at room
temperature and a resulting dimensional change in the geometry of the component. If great enough, this
dimensional change could lead to growth and in severe instances, crack initiation.

• Fatigue: Low retained austenite content and fine austenitic grain sizes, which create a microstructure of finely
dispersed retained austenite and tempered martensite, prevent nucleation of fatigue cracks, or retard fatigue crack
initiation until very high stress levels are reached. In contrast, low-stress applications that fracture at low cycles are
related to high retained austenite levels and coarse austenite grain sizes.
For example, one type of fatigue strength of interest to many users is rolling contact fatigue. Two aspects of
retained austenite can improve rolling contact fatigue life. First, the inherent ductility of retained austenite helps to
delay crack growth by blunting the tips of cracks as they form. Second, as retained austenite transforms during
service, compressive residual stresses increase in the case. These compressive stresses delay crack growth. These
benefits are not present in a part with insufficient retained austenite.
4.7 Heat Treatment of Ferrous Alloys
• Impact: Impact strength is the measure of the ability of a steel to resist fracture when subjected to a sharp blow.
Austenite is not only very tough, but also it has higher impact strength than martensite. The steel's impact strength
increases with increasing austenite content. Higher impact strength can provide extra protection against cracking,
which, in turn, helps prevent such problems as spalling.
It is important to recognize that a balance must be created between the mechanical properties of a component and
the optimum percentage of retained austenite for a given application.
How some industries view retained austenite?
Retained austenite is highly undesirable in components for the tool and die industry. RA is recognized as a major
cause for premature failure. The low hardness of RA is also incompatible with most applications that require the
maximum attainable hardness to resist wear.
The bearing and gear industries have a more favorable view toward having some retained austenite (5 to 30%
determined by optical metallography). Gears are typically made of case-hardened steel, which has high impact
strength.
4.7 Heat Treatment of Ferrous Alloys
While most tools fail by wear or fracture, many gear failures are the result of spalling in the tooth area. Spalling is
progressive macropitting that occurs when pits coalesce and form irregular craters, which cover a significant area of
the tooth surface. Spalling occurs when the surface of a metal component is subjected to repeated cyclic loads.
A crack forms and grows until a small portion of the surface breaks loose, damaging the surface and adding debris to
the system. The gear industry balances the amount of retained austenite in a gear tooth to delay the onset of spalling
by suppressing crack growth.

Fig. Gear tooth failure due to spalling (macropitting).


4.7 Heat Treatment of Ferrous Alloys
Tempered Martensite. Martensite is tempered in order to improve its mechanical properties. Tempering is a heating
process by which hardness is reduced and toughness is increased. The bct martensite is heated to an intermediate
temperature, typically 150–650°C, where it decomposes to a two-phase microstructure, consisting of bcc alpha
ferrite and small particles of cementite.

With increasing tempering time and temperature, the hardness


of tempered martensite decreases (Fig. 4.14). The reason is
that the cementite particles coalesce and grow, and the
distance between the particles in the soft ferrite matrix
increases as the less stable and smaller carbide particles
dissolve.

Fig. 4.14. Hardness of tempered martensite as a function of


tempering time for 1080 steel quenched to 65 HRC. Hardness
decreases because the carbide particles coalesce and grow in size,
thereby increasing the interparticle distance of the softer ferrite.
4.7 Heat Treatment of Ferrous Alloys
Tempered Martensite. Tempering is a term historically associated with the heat treatment of martensite in steels. It
describes how the microstructure and mechanical properties change as the metastable sample is held isothermally at
a temperature where austenite cannot form. The changes during the tempering of martensite can be categorized
into stages.
• During the first stage, excess carbon in solid solution segregates to defects or forms clusters within the solid
solution. It then precipitates, either as cementite in low-carbon steels, or as transition iron-carbides in high-
carbon alloys. The carbon concentration that remains in solid solution may be quite large if the precipitate is a
transition carbide.
• Further annealing leads to stage 2, in which almost all of the excess carbon is precipitated, and the carbides all
convert into more stable cementite. Any retained austenite may decompose during this stage. Continued
tempering then leads to the coarsening of carbides, extensive recovery of the dislocation structure, and finally to
the recrystallisation of the ferrite plates into equiaxed grains.
4.7 Heat Treatment of Ferrous Alloys
The virgin microstructure obtained immediately after quenching from austenite consists of plates or laths of
martensite which is supersaturated with carbon. In the vast majority of steels, the martensite contains a substantial
density of dislocations which are generated during the imperfect accommodation of the shape change accompanying
the transformation. The plates may be separated by thin films of retained austenite, the amount of untransformed
austenite becoming larger as the martensite-start temperature MS is reduced.

Transmission electron micrograph of as-quenched Corresponding dark-field image showing the


martensite in a Fe-4Mo-0.2C wt% steel. distribution of retained austenite.
4.7 Heat Treatment of Ferrous Alloys
Although most textbooks will begin a discussion of tempering with this first stage of tempering, involving the
redistribution of carbon and precipitation of transition carbides, cementite can precipitate directly. This is particularly
the case when the defect density is large. Trapped carbon atoms will not precipitate as transition carbides but
cementite is more stable than trapped carbon.

Transmission electron micrograph of martensite in a Fe-4Mo- Corresponding dark-field image showing the
0.2C wt% steel after tempering at 190oC for 1 hour. The carbon distribution of retained austenite, which has not
has in this case precipitated as fine particles of cementite. been affected by the tempering.
4.7 Heat Treatment of Ferrous Alloys
Tempering at higher temperatures, in the range 200-300oC for 1h induces the retained austenite to decompose into a
mixture of cementite and ferrite. When the austenite is present as a film, the cementite also precipitates as a
continuous array of particles which have the appearance of a film.

Dark field transmission electron micrograph of martensite in a Fe-4Mo-0.2C wt%


steel after tempering at 295oC for 1 hour. Only the cementite is illuminated. The
film of cementite at the martensite plate boundaries is due to the decomposition of
retained austenite.
4.7 Heat Treatment of Ferrous Alloys
Tempering at even higher temperatures leads to a coarsening of the cementite particles, with those located at the
plate boundaries growing at the expense of the intra-plate particles. The dislocation structure tends to recover, the
extent depending on the chemical composition. The recovery is less marked in steels containing alloying elements
such as molybdenum and chromium.

Bright field transmission electron micrograph of martensite in a Fe-4Mo-0.2C


wt% steel after tempering at 420oC for 1 hour.
4.7 Heat Treatment of Ferrous Alloys
4.7.1 Time–temperature–transformation Diagrams
The percentage of austenite transformed into pearlite is a
function of temperature and time (Fig. 4.15a). This
transformation is best illustrated in diagrams called isothermal
transformation (IT) diagrams, or time–temperature–
transformation (TTT) diagrams, constructed from the data given
in Fig. 4.15a. The higher the temperature or the longer the
time, the more austenite is transformed into pearlite.

Fig. 4.15 (a) Austenite-to-pearlite transformation


of iron–carbon alloy as a functionof time and Fig. 4.15 (b) Isothermal transformation diagram obtained from
temperature, (a) for a transformation temperature of 675°C.
4.7 Heat Treatment of Ferrous Alloys
4.7.1 Time–temperature–transformation Diagrams
For each temperature, there is a minimum time for the transformation
to begin. This time period defines the critical cooling rate; with
longer times, austenite begins to transform into pearlite (Fig. 4.15c).
The TTT diagrams shown allow the design of heat-treatment
schedules to obtain desirable microstructures.
In Fig. 4.15c; the steel can be raised over the eutectic temperature to
start with a state of austenite. If the material is cooled rapidly, it can
follow the 140°C/s cooling rate trajectory shown, resulting in
complete martensite. On the other hand, it can be more slowly cooled
(in a molten salt bath) to develop pearlite- or bainite-containing
steels.
If tempered martensite is desired, the heat treat and quench stages is
followed by a tempering process.
Fig. 4.15 (c) Microstructures obtained for a
eutectoid iron–carbon alloy as a function of
cooling rate.
4.7 Heat Treatment of Ferrous Alloys
4.7.1 Time–temperature–transformation Diagrams
If tempered martensite is desired, the heat treat and quench stages is followed by a tempering process. The
differences in hardness and toughness of the various structures obtained are shown in Fig. 4.16. Fine pearlite is
harder and less ductile than coarse pearlite. The effects of various percentages of carbon, cementite, and pearlite on
other mechanical properties of steels are shown in Fig. 4.17.

Fig. 4.16 (a) and (b) Hardness and (c) toughness for annealed plain-carbon steels as a function of carbide shape. Carbides in the
pearlite are lamellar. Fine pearlite is obtained by increasing the cooling rate. The spheroidite structure has sphere-like carbide
particles.
4.7 Heat Treatment of Ferrous Alloys
4.7.1 Time–temperature–transformation Diagrams
The effects of various percentages of carbon, cementite, and pearlite on other mechanical properties of steels are
shown in Fig. 4.17.

Fig. 4.17. Mechanical properties of annealed steels as a function of composition and microstructure. Note in (a) the increase in
hardness and strength, and in (b), the decrease in ductility and toughness, with increasing amounts of pearlite and iron carbide.
4.8 Hardenability of Ferrous Alloys
Hardenability. The capability of an alloy to be hardened by heat treatment is called its hardenability, and is a
measure of the depth of hardness that can be obtained by heating and subsequent quenching. The term
“hardenability” should not be confused with “hardness,” which is the resistance of a material to indentation or
scratching. From the discussion thus far, it can be seen that hardenability of ferrous alloys depends on the (a) carbon
content, (b) grain size of the austenite, (c) alloying elements present in the material, and (d) cooling rate.

High hardenability refers to the ability of the alloy to produce a high martensite percentage throughout the body of
the material upon quenching. Hardened steels are created by rapidly quenching the material from a high temperature.
This involves a rapid transition from a state of 100% austenite to a high percentage of martensite. If the steel is
more than 0.15% carbon, the martensite becomes a highly strained body-centered cubic form and is supersaturated
with carbon. The carbon effectively shuts down most slip planes within the microstructure, creating a very hard and
brittle material.
If the quenching rate is not fast enough, carbon will diffuse out of the austenitic phase. The steel then becomes
pearlite, bainite, or if kept hot long enough, ferrite. None of the microstructures just stated have the same strength as
martensite after tempering and are generally seen as unfavorable for most applications.
4.8 Hardenability of Ferrous Alloys
Hardenability curves are dependent on carbon content. A greater percentage of carbon present in steel will increase
its hardness. Carbon is not the only alloying element that can have an effect on hardenability.
Table below shows a comparison of the alloying content in some steels. 1040 is a plain carbon steel and therefore
has the lowest hardenability as there are no other elements besides iron to block the carbon atoms from escaping
the matrix. The nickel added to 4340 allows for a slightly greater amount of martensite to form compared to 4140,
giving it the highest hardenability of these three alloys. Most metallic alloying elements slow down the formation
of pearlite, ferrite and bainite, therefore they increase a steel’s hardenability.

Table: Shows the alloying contents of 4340, 4140, and 1040 steel
Type of Steel: Nickel (wt %) Molybdenum (wt %) Chromium (wt %)
4340 1.85% 0.25% 0.80%
4140 0.00% 0.20% 1.00%
1040 0.00% 0.00% 0.00%
4.8 Hardenability of Ferrous Alloys
4.8.1 The End-quench Hardenability Test
In this commonly used Jominy test, a round test bar 100 mm long, made from a particular alloy, is austenitized—that
is, heated to the proper temperature to form 100% austenite. It is then quenched directly at one end (Fig. 4.18a) with
a stream of water at 24°C. The cooling rate thus varies throughout the length of the bar, the rate being highest at the
lower end, being in direct contact with the water.

Fig. 4.18. (a) End-quench test and cooling rate,


4.8 Hardenability of Ferrous Alloys
4.8.1 The End-quench Hardenability Test
The hardness along the length of the bar is then measured at various distances from the quenched end. As shown,
hardness decreases away from the quenched end of the bar (Fig. 4.18b). The greater the depth to which hardness
increases, the greater the hardenability of the alloy.

Each composition of an alloy has its particular


hardenability band. Note that the hardness at the
quenched end increases with increasing carbon content,
and that 1040, 4140, and 4340 steels have the same
carbon content (0.40%) and thus they have the same
hardness (57 HRC) at the quenched end.

Fig. 4.18. (b) Hardenability curves for five different steels, as


obtained from the end-quench test. Small variations in
composition can change the shape of these curves. Each curve is
actually a band, and its exact determination is important in the
heat treatment of metals, for better control of properties.
4.8 Hardenability of Ferrous Alloys
4.8.2 Quenching Media
The fluid used for quenching the heated specimen also has an effect on hardenability. Quenching may be carried out
in water, brine (salt water), oil, molten salt, or air; caustic solutions, and polymer solutions; gases may also be used.
Because of the differences in thermal conductivity, specific heat, and heat of vaporization of these media, the rate of
cooling of the specimen (severity of quench) is also different. In relative terms and in decreasing order, the cooling
capacities of several quenching media are: agitated brine, 5; still water, 1; still oil, 0.3; cold gas, 0.1; and still air,
0.02. Agitation also is a significant factor in the rate of cooling; the more vigorous the agitation, the higher is
the rate of cooling.
The cooling rate also depends on the surface area-to-thickness or surface-area-to-volume ratio of the part to be
quenched; the higher this ratio, the higher is the cooling rate. For example, a thick plate cools more slowly than a
thin plate with the same surface area. These considerations are also significant in the cooling of metals and of
plastics in casting and in molding processes.
4.8 Hardenability of Ferrous Alloys
4.8.2 Quenching Media
Water is a common medium for rapid cooling. However, the heated specimen may form a vapor blanket along its
surfaces due to the water-vapor bubbles that form when water boils at the metal–water interface. This blanket creates
a barrier to heat conduction, because of the lower thermal conductivity of the vapor. Agitating the fluid or the part
helps to reduce or eliminate the blanket; also, water may be sprayed onto the part under high pressure.
Brine is an effective quenching medium, because salt helps to nucleate bubbles at the interfaces, which improves
agitation; note, however, that brine can corrode the part.
Polymer quenchants can be used for ferrous as well as for nonferrous alloys; they have cooling characteristics that
generally are between those of water and petroleum oils. Typical polymer quenchants are polyvinyl alcohol,
polyalkaline oxide, polyvinyl pyrrolidone, and polyethyl oxazoline. These quenchants have such advantages as
better control of hardness, elimination of fumes and fire (as may occur when oils are used as a quenchant), and
reduction of corrosion (as may occur when water is used).
4.9 Heat Treatment of Nonferrous Alloys and Stainless Steels
Nonferrous alloys and some stainless steels cannot be heat treated by the techniques described for ferrous alloys. The reason is
that nonferrous alloys do not undergo phase transformations like those in steels; the hardening and strengthening mechanisms
for these alloys are fundamentally different.
Precipitation hardening: Heat-treatable aluminum alloys,
copper alloys, and martensitic and some other stainless
steels are hardened and strengthened by a process called
precipitation hardening. In this process, small particles of
a different phase, called precipitates, are uniformly
dispersed in the matrix of the original phase (Fig. 4.3a).
Precipitates form because the solid solubility of one
element (one component of the alloy) in the other is
exceeded.
Fig. Age hardening of a typical Al-Cu alloy.
1) Solid solution α at T1 ,
2) Quenched to form super saturated solid solution,
3) Annealing, θ phase is segregated at coarse grain
boundaries,
4) Age-hardening, submicron precipitation,
5) Averaging coarsen the precipitates.
4.9 Heat Treatment of Nonferrous Alloys and Stainless Steels
Three stages are involved in precipitation hardening, which can best be described by reference to the phase diagram
for the aluminum–copper system (Fig. 4.19a). For a composition of 95.5% Al–4.5% Cu, a single-phase (α)
substitutional solid solution of copper (solute) in aluminum (solvent) exists between 500°C and 570°C.
This alpha phase is aluminum rich, has an FCC structure, and is ductile.
Below the lower temperature (that is, below the
lower solubility curve) there are two phases:
alpha and theta (θ) which is a hard intermetallic
compound of CuAl2. This alloy can be heat
treated and its properties modified by two
different methods: solution treatment and
precipitation hardening.

Fig. 4.19 (a) Phase diagram for the


aluminum–copper alloy system. (b) Various
microstructures obtained during the age-
hardening process.
4.9 Heat Treatment of Nonferrous Alloys and Stainless Steels
• In solution treatment, the alloy is heated to within the solid-solution kappa phase (say, 540°C) and then cooled
rapidly, such as by quenching it in water. The structure obtained soon after quenching (A in Fig. 4.19b) consists
only of the single phase alpha. This alloy has moderate strength and considerable ductility.
• The structure obtained in A in Fig. 4.19b can be made stronger by precipitation hardening. In this process, the
alloy is first reheated to an intermediate temperature and then held there for a period of time, during which
precipitation takes place. The copper atoms diffuse to nucleation sites and combine with aluminum atoms. This
process produces the theta phase (θ), which forms as submicroscopic precipitates (shown in B by the small dots
within the grains of the kappa phase). The resulting structure is stronger than that in A, although it is less ductile.
The increase in strength is due to increased resistance to dislocation movement in the region of the precipitates.

Al – 4% Cu alloy; a) slow cooling,


b) quenching and precipitation hardening
4.9 Heat Treatment of Nonferrous Alloys and Stainless Steels
Aging. Because the precipitation process is one of time and temperature, it is also called aging, and the property
improvement is known as age hardening. If carried out above room temperature, the process is called artificial
aging. Several aluminum alloys harden and become stronger over a period of time at room temperature; this process
is then called natural aging. Such alloys are first quenched and then, are shaped by plastic deformation at room
temperature. Finally, they are allowed to develop strength and hardness by aging naturally. The rate of natural aging
can be slowed by refrigerating the quenched alloy (cryogenic treatment).

Over-aging. In the precipitation process, if the reheated alloy is held at the elevated temperature for an extended
period of time, the precipitates begin to coalesce and grow. They become larger but fewer in number, as is shown by
the larger dots in C in Fig. 4.19b. This process is called over-aging, and the resulting alloy is softer and weaker.
4.9 Heat Treatment of Nonferrous Alloys and Stainless Steels
Aging. There is an optimal time–temperature relationship in the aging process that must be followed in order to
obtain the desired properties (Fig. 4.20). It is apparent that an aged alloy can be used only up to a certain maximum
temperature in service, otherwise it will over-age and lose its strength and hardness. Although weaker, an over-aged
part has better dimensional stability.

Fig. 4.20. The effect of aging time and temperature on the yield
stress of 2014-T4 aluminum alloy; note that, for each
temperature, there is an optimal aging time for maximum
strength.
4.10 Case Hardening
The heat-treatment processes described thus far involve microstructural alterations and property changes in the bulk
of the component by means of through hardening. It is not always desirable to through harden parts, because a hard
part lacks the required toughness for some applications. For example, a small surface crack could propagate rapidly
through a part and cause sudden and total failure.
In many cases, modification of only the surface properties of a part (hence, the term surface or case hardening) is
desirable. This widely used method is particularly useful for improving resistance to surface indentation, fatigue, and
wear.
Typical applications for case hardening are gear teeth, cams, shafts, bearings, fasteners, pins, automotive clutch
plates, tools, and dies.
Several case-hardening processes are available (Table 4.1):
1. Carburizing (gas, liquid, and pack carburizing)
2. Carbonitriding
3. Cyaniding
4. Nitriding
5. Boronizing
6. Flame hardening
7. Induction hardening
4.10 Case Hardening

Table 4.1. Outline of Heat-treatment Processes for Surface Hardening


Element
Typical
Process Metals hardened added to Procedure General characteristics
applications
surface
Heat steel at 870°C–950°C in
A hard, high-carbon surface
an atmosphere of Gears, cams,
is produced. Hardness = 55–
Low-carbon steel carbonaceous gases shafts, bearings,
65 HRC. Case depth =
Carburizing (0.2% C), alloy steels C (gas carburizing) or carbon- piston pins,
<0.5–1.5 mm. Some
(0.08–0.2% C) containing solids (pack sprockets, clutch
distortion of part during heat
carburizing); then plates
treatment.
quench.
Heat steel at 700°C–800°C in Surface hardness = 55–62
an atmosphere of HRC. Case depth = 0.07–0.5
Carbonitriding Low-carbon steel C and N carbonaceous gas mm. Less Bolts, nuts, gears
and ammonia; then quench in distortion than in
oil. carburizing.
Heat steel at 760°–845°C in a
C and N Surface hardness up to 65
Low-carbon steel molten bath of solutions of Bolts, nuts,
(from HRC. Case depth = 0.025–
Cyaniding (0.2% C), alloy steels cyanide screws, small
cyanide 0.25 mm. Some
(0.08–0.2% C) (e.g., 30% sodium cyanide) gears
NaCN) distortion.
and other salts.
4.10 Case Hardening
Table 4.1. Outline of Heat-treatment Processes for Surface Hardening
Element
Process Metals hardened added to Procedure General characteristics Typical applications
surface
Nitriding Steels (1% Al, 1.5% N Heat steel at 500°–600°C in an Surface hardness up to Gears, shafts, sprockets,
Cr, 0.3% Mo), alloy atmosphere of ammonia gas 1100 HV (~68HRC). valves, cutters, boring
steels (Cr, Mo), (NH3) or mixtures of molten Case depth = 0.1–0.6 bars, fuel-injection pump
stainless steels, cyanide salts; no further mm and 0.02–0.07 mm parts
high-speed tool steels treatment. for high-speed steel.
Boronizing Steels B Part is heated using boron- Extremely hard and Tool and die steels
containing gas or solid in wear-resistant surface.
contact with part. Case depth = 0.025–
0.075 mm.
Flame Medium-carbon steels, None Surface is heated with an Surface hardness = 50– Gear and sprocket teeth,
hardening cast irons oxyacetylene torch, then 60 HRC. Case depth = axles, crankshafts, piston
quenched with water spray or 0.7–6 mm. Little rods, lathe beds and
other quenching methods. distortion. centers
Induction Medium-carbon steels, None Metal part is placed in copper Surface hardness = 50– Gear and sprocket teeth,
hardening cast irons induction coils and is heated by 60 HRC. Case depth = axles, crankshafts, piston
high frequency current; then 0.7–6 mm. Little rods, lathe beds and
quenched. distortion. centers
4.10 Case Hardening
• Basically, these are operations where the component is heated in an atmosphere containing elements (such as
carbon, nitrogen, or boron) that alter the composition, microstructure, and properties of surface layers.
• For steels with sufficiently high carbon content, surface hardening takes place without using any of these
additional elements; only the heat-treatment processes, usually by either flame hardening or induction
hardening, are needed to alter the microstructures.
• Because case hardening involves a localized layer, case-hardened parts have a hardness gradient. Typically, the
hardness is maximum at the surface and decreases inward,
the rate of decrease depending on the composition and
physical properties of the metal and processing
variables.

• Decarburization is the phenomenon in which alloys lose carbon from their surfaces as a result of heat treatment or of hot
working in a medium, usually oxygen, that reacts with the carbon. Decarburization is undesirable because it affects the
hardenability of the surfaces of a part, by lowering its carbon content; it also adversely affects the hardness, strength, and
fatigue life of steels, significantly lowering their endurance limit.
Boriding/ Boronizing Heat Treatment for Extreme Wear
Boriding, or Boronizing, is a thermo-chemical surface hardening process used to improve the life and performance
of metal components. The process is used to strengthen resistance to corrosion and abrasive wear, decrease
coefficient of friction, and, of course, greatly increase surface hardness.
The Boriding process involves the use of specially formulated boron-yielding material heated to temperatures
between 700 and 1000°C. Boron atoms diffuse into the substrate, forming very hard borides within the surface of the
material.
During the boriding process, complex intermetallic compounds are created between the elements iron, boron,
chromium, nickel, vanadium, etc. The resulting borides form a hard peripheral surface layer consisting of Fe 2B and
other compounds. Because of its crystalline structure, the boride layer is anchored exceptionally well to the base
material.
The boriding procedure is especially suitable for unalloyed and low alloyed steels. With increasing content of
alloying elements, the diffusion rate and thus the thickness of the achievable boride layer decreases. At the same
time, alloying elements such as nickel and chromium contribute to an increase in the hardness and wear
resistance.
Boriding / Boronizing Heat Treatment for Extreme Wear
• Types of material for boronizing. Materials that generally can be borided include low and medium carbon
steels, tool steels, stainless steels, Ni-base alloys, and tungsten carbide alloys. A restriction is constituted by
materials containing silicon and aluminum. Steels with correspondingly high levels of the elements mentioned
can affect the boriding result.
• Low friction properties. The increased resistance to abrasive wear is accompanied by an improved resistance to
cavitation. The reduced friction coefficient of boronized surfaces improves the lubricating properties and reduces
wear. In the event of a lubricating film break, the boride layer provides excellent emergency running properties,
even at high temperatures up to 1000 °C.
• Diffusion in boriding. The boron diffusion process consists of two separate reactions. The first reaction is a slow
process between the boron and the material that produces a very hard, thin boride layer at the surface.

The second reaction involves the diffusion of boron


further into the substrate at a much quicker rate.
Boriding / Boronizing Heat Treatment for Extreme Wear
The boride layer formed at the surface can be that of either a one or two-phase formation. In ferrous materials, the
one-phase formation consists of a single Fe2B layer, while the two-phase formation consists of an FeB layer in
addition to the Fe2B layer. (FeB is harder than Fe2B, but is more brittle and more easily fractured upon impact.)

The thickness of the boride layers varies with


respect to temperature, treating time, and
material. The FeB and Fe2B layers will
typically have hardness readings of 1700 - 1900
HV and 1400 - 1600 HV respectively with a
total depth 0.0127 - 0.508 mm.
The FeB layer, while harder, is more brittle and
more prone to fracture upon impact.
4.11. Annealing
In metallurgy and materials science, annealing is a heat treatment that alters the physical and sometimes chemical
properties of a material to increase its ductility and reduce its hardness, making it more workable. It involves heating
a material above its recrystallization temperature, maintaining a suitable temperature for an appropriate amount of
time and then cooling.
Annealing is a general term used to describe the restoration of a cold-worked or heat- treated alloy to its original
properties—for instance, to increase ductility (and hence formability) and to reduce hardness and strength, or to
modify the microstructure of the alloy.
The annealing process is also used to relieve residual
stresses in a part, as well as to improve machinability and
dimensional stability. The term “annealing” also applies to
the thermal treatment of glass and similar products, castings,
and weldments.

Fig. 4.21. Heat-treating temperature ranges for plain carbon


steels, as indicated on the iron-iron carbide phase diagram.
4.11. Annealing
The annealing process consists of the following steps:
1. Heating the workpiece to a specific temperature in a furnace,
2. Holding it at that temperature for a period of time (soaking),
3. Cooling the workpiece, in air or in a furnace.
In annealing, atoms migrate in the crystal lattice and the number of dislocations decreases, leading to a change in
ductility and hardness. As the material cools it recrystallizes. For many alloys, including carbon steel, the crystal
grain size and phase composition, which ultimately determine the material properties, are dependent on the heating
rate and cooling rate.
Hot working or cold working after the annealing process alters the metal structure , so further heat treatments may be
used to achieve the properties required. With knowledge of the composition and phase diagram, heat treatment can
be used to adjust from harder and more brittle to softer and more ductile.
The annealing process may be carried out in an inert or a controlled atmosphere, or it may be performed at lower
temperatures to minimize or prevent surface oxidation.
4.11. Annealing
The annealing temperature may be higher than the material’s recrystallization temperature, depending on the degree
of cold work. For example, the recrystallization temperature for copper ranges between 200° and 300°C, whereas the
annealing temperature required to fully recover the original properties ranges from 260° to 650°C, depending on the
degree of prior cold work.
Annealing occurs by the diffusion of atoms within a solid material, so that the material progresses towards its
equilibrium state. Heat increases the rate of diffusion by providing the energy needed to break bonds. The movement
of atoms has the effect of redistributing and eradicating the dislocations in metals and (to a lesser extent) in
ceramics. This alteration to existing dislocations allows a metal object to deform more easily, increasing its ductility.
The amount of process-initiating Gibbs free energy in a deformed metal is also reduced by the annealing process. In
practice and industry, this reduction of Gibbs free energy is termed stress relief.
Gibbs free energy: The maximum amount of non-expansion work that can be extracted from a thermodynamically
closed system.
The relief of internal stresses is a thermodynamically spontaneous process; however, at room temperatures, it is a
very slow process. The high temperatures at which annealing occurs serve to accelerate this process.
4.11. Annealing
Full annealing is a term applied to ferrous alloys. The steel is heated to above A1 or A3 (Fig. 4.21) and the cooling
takes place slowly [typically at 10°C per hour], in a furnace, after which it is turned off. The cooling rate of the steel
has to be sufficiently slow so as to not let the austenite transform into bainite or martensite, but rather have it
completely transform to pearlite and ferrite or cementite. The structure obtained through full annealing is coarse
pearlite, which is soft and ductile and has small, uniform grains.
A full anneal typically results in the second most
ductile state a metal can assume for metal alloy. Its
purpose is to originate a uniform and stable
microstructure that most closely resembles the metal's
phase diagram equilibrium microstructure, thus
letting the metal attain relatively low levels of
hardness, yield strength and ultimate strength with high
plasticity and toughness.
4.11. Annealing
Normalizing. To avoid excessive softness from annealing of steels, the cooling cycle may be done completely in
still air. This process is called normalizing, to indicate that the part is heated to a temperature above A3 or Acm in
order to transform the structure to austenite. Normalizing results in somewhat higher strength and hardness, and
lower ductility, than does full annealing (Fig. 4.22).
The structure obtained is fine pearlite, with small, uniform grains.
Normalizing is generally carried out to refine the grain structure,
obtain uniform structure (homogenization), decrease residual
stresses, and improve machinability.

Fig. 4.22. Hardness of steels in the quenched and normalized


conditions as a function of carbon content.
4.11. Annealing
Spheroidite annealing: hypereutectoid steel with a microstructure of
pearlite and cementite network generally gives poor machinability.
This poor machinability is due to the presence of high level of plate-
shape cementite in the microstructure. Since cementite is hard and
brittle, the cutting tool cannot cut through these plates. Instead, the
plates have to be broken. Therefore, the tool is subjected to continual
shock load by the cementite plates and results in a ragged surface
finish. A heat-treating process which will improve the machinability is
known as spheroidize annealing. This process will produce a
Fig. 4.12. Microstructure of eutectoid steel.
spheroidal or globular form of carbide in a ferritic matrix as shown
Spheroidite is formed by tempering the
in the figure. steel at 700°C.
4.11. Annealing
Spheroidite. When pearlite is heated to just below the eutectoid
temperature and then held at that temperature for a period of time,
such as at 700°C for a day, the cementite lamellae transform to
roughly spherical shapes. Unlike the lamellar shapes of cementite,
which act as stress raisers, spheroidites (spherical particles) have
smaller stress concentrations because of their rounded shapes.
Consequently, this structure has higher toughness and lower
hardness than the pearlite structure. It can be cold worked, because
the ductile ferrite has high toughness and the spheroidal carbide
Fig. 4.12. Microstructure of eutectoid steel.
particles prevent the initiation of cracks within the material.
Spheroidite is formed by tempering the
Spheroidizing annealing improves the cold workability and the steel at 700°C.
machinability of steels.
4.11. Annealing
Stress-relief Annealing. To reduce or eliminate residual stresses, a
workpiece is generally subjected to stress-relief annealing, or stress
relieving. The temperature and time required for this process depend
on the material and on the magnitude of the residual stresses present.
The residual stresses may have been induced during forming, machining,
or caused by volume changes during phase transformations.

• Stress-relief annealing is used to eliminate and/or minimize stresses


arising from plastic deformation during machining or forming
processes.
• Stress-relief annealing temperatures are relatively low, so that useful
effects of cold working are not eliminated. Therefore, the treatment
is not intended to produce significant changes in material
structures or mechanical properties and is normally restricted to
relatively low temperatures. The given figure shows that heating
temperature for the stress relief annealing is 625o−650oC.
4.11. Annealing
Stress Relieving. is applied to both ferrous and non-ferrous alloys and is intended to remove internal residual
stresses generated by prior manufacturing processes such as machining, cold rolling and welding. Without it,
subsequent processing may give rise to unacceptable distortion and/or the material can suffer from service problems
such as stress corrosion cracking.
Carbon steels and alloy steels can be given two forms of stress relief:
• Treatment at typically 150o−200oC relieves peak stresses after hardening without significantly reducing hardness
(e.g. case-hardened components, bearings, …)
• Treatment at typically 600o−680oC (e.g. after welding, machining, …) provides virtually complete stress relief.
Non-ferrous alloys are stress relieved at a wide variety of temperatures related to alloy type and condition. Alloys
that have been age-hardened are restricted to stress relieving temperatures below the ageing temperature.
Austenitic stainless steels are stress relieved below 480°C or above 900°C, temperatures in between reducing
corrosion resistance in grades that are not stabilized or low-carbon. Treatments above 900°C are often full solution
anneals.
4.11. Annealing
Tempering. If steels are hardened by heat treatment, tempering is used in order to reduce brittleness, increase
ductility and toughness, and reduce residual stresses. The term “tempering” is also used for glasses.
In tempering, the steel is heated to a specific temperature, depending on its composition, and then cooled at a
prescribed rate. The results of tempering for an oil-quenched AISI 4340 steel are shown in Fig. 4.23.
Alloy steels may undergo temper embrittlement, which is caused by the
segregation of impurities along the grain boundaries, at temperatures
between 480°C and 590°C.
Tempering is usually performed after hardening, to reduce some of the
excess hardness, and is done by heating the metal to some temperature below
the critical point A1 for a certain period of time, then allowing it to cool in
still air. The temperature determines the amount of hardness removed, and
depends on both the specific composition of the alloy and on the desired
properties in the finished product. For instance, very hard tools are often
tempered at low temperatures, while springs are tempered at much higher Fig. 4.23. Mechanical properties of oil-
temperatures (below A1). quenched 4340 steel as a function of
tempering temperature.
4.11. Annealing
Tempering. is a heat-treatment technique applied to ferrous alloys, such as steel or cast iron, to achieve greater
toughness by decreasing the hardness of the alloy. The reduction in hardness is usually accompanied by an increase
in ductility, thereby decreasing the brittleness of the metal.
Tempering is usually performed after quenching, which is rapid cooling of the metal to put it in its hardest state.

Tempering is accomplished by controlled heating of the quenched work-piece


to a temperature below its "lower critical temperature". This is also called the
lower transformation temperature or lower arrest (A1) temperature; the
temperature at which the crystalline phase of the alloy, called ferrite and
cementite, begin combining to form a single-phase solid solution referred to as
austenite. Heating above this temperature is avoided, so as not to destroy the Fig. Microscopic image of martensite,
very-hard, quenched microstructure, called martensite. a very hard microstructure formed
when steel is quenched. Tempering
reduces the hardness in the martensite
by transforming it into various forms
of tempered martensite.
4.11. Annealing
During tempering process, precise control of time and temperature is crucial to achieve the desired balance of
physical properties. Low tempering temperatures may only relieve the internal stresses, decreasing brittleness while
maintaining a majority of the hardness. Higher tempering temperatures tend to produce a greater reduction in the
hardness, sacrificing some yield strength and tensile strength for an increase in elasticity and plasticity.
However, in some low alloy steels, containing other elements like chromium and molybdenum, tempering at low
temperatures may produce an increase in hardness, while at higher temperatures the hardness will decrease. Many
steels with high concentrations of these alloying elements behave like precipitation hardening alloys, which
produces the opposite effects under the conditions found in quenching and tempering, and are referred to as marging
steels.
In carbon steels, tempering alters the size and distribution of carbides in the martensite, forming a microstructure
called "tempered martensite". Tempering is also performed on normalized steels and cast irons, to increase ductility,
machinability, and impact strength. Steel is usually tempered evenly, called "through tempering," producing a nearly
uniform hardness, but it is sometimes heated unevenly, referred to as "differential tempering," producing a variation
in hardness.
4.11. Annealing
Austempering. In austempering, the heated steel is quenched from the austenitizing temperature rapidly, to avoid
formation of ferrite or pearlite. It is held at a certain temperature until isothermal transformation from austenite to
bainite is complete. It is then cooled to room temperature, usually in still air and at a moderate rate in order to avoid
thermal gradients within the part.
The quenching medium most commonly used is molten
salt, at temperatures ranging from 160°C to 750°C
(depending on the type of desired bainite and to avoid
formation of martensite).
Austempering is often substituted for conventional
quenching and tempering, either to reduce the tendency
for cracking and distortion during quenching or to
improve ductility and toughness while maintaining
hardness. As compared to quenching and tempering,
because of the shorter cycle time involved in
austempering, this process is also economical.
4.11. Annealing
Austempering is an isothermal heat treatment process which is carried out on ferrous materials to produce improved
mechanical properties and reduced distortion.
Austempering produces bainitic microstructure in case of steels where as in case of cast iron the microstructure
consists of ausferrite (mixture of acicular ferrite and high carbon stabilised austenite).
Austempering consists of quenching from the austenitising temperature into a molten salt bath (nitrite salt bath)
which is held at a constant temperature called the austempering temperature. The temperature ranges from 240 to
400oC, depending upon the material being austempered and the properties desired.
The quenched parts are then kept immersed in the molten salt bath for
a certain amount of time. The holding times can vary from few
minutes to few hours, depending upon the material being austempered
and the properties desired. During the holding period, an isothermal
reaction takes place and the desired microstructure is obtained.
Austempering usually produces an uniform microstructure.
Austempered ductile iron castings are becoming increasing popular in
the automotive industry as they offer cost and weight competitive
advantages over the other conventionally used competing materials.
4.11. Annealing
Martempering (Marquenching). In martempering, steel or cast iron is first quenched from the austenitizing
temperature in a hot fluid medium, such as hot oil or molten salt. Next, it is held at that temperature until the
temperature is uniform throughout the part. It is cooled at a moderate rate, such as in air, in order to avoid excessive
temperature gradients within the part. Usually, the part is subsequently tempered, because the structure obtained is
otherwise primarily untempered martensite, and thus
not suitable for most applications.
4.11. Annealing
Martempering is also known as stepped quenching or interrupted quenching. In this process, steel is heated above
the upper critical point (above the transformation range) and then quenched in a salt, oil, or lead bath kept at a
temperature of 150-300°C. The workpiece is held at this temperature above martensite start (Ms) point until the
temperature becomes uniform throughout the cross-section of workpiece. After that, it is cooled in air or oil to room
temperature. The steel is then tempered. In this process, Austenite is transformed to martensite by step quenching, at
a rate fast enough to avoid the formation of ferrite, pearlite or bainite.
In the martempering process, austenitized metal part is immersed in a bath at a temperature just above the
martensite start temperature (Ms). By using interrupted quenching, the cooling is stopped at a point above the
martensite transformation region to ensure sufficient time for the center to cool to the same temperature as the
surface. The metal part is then removed from the bath and cooled in air to room temperature to permit the austenite
to transform to martensite. Martempering is a method by which the stresses and strains generated during the
quenching of a steel component can be controlled. In Martempering steel is heated to above the critical range to
make it all austenite.
Martempered steels have lower tendency to crack, distort, or develop residual stresses during heat treatment.
The drawback of this process is that the large section cannot be heat treated by this process.

You might also like