Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Heat treatment

Heat treatment is defined as ‘A‘A combination of heating and cooling operations, timed and applied to
a metal in the solid state in a way that will produce desired properties’, i.e. it is an operation or
combination of heating and cooling a solid metal or alloy to induce in it certain predetermined
pr
physical and mechanical properties. These properties are dependent on the microstructure of the
alloy, i.e, nature, shape, size distribution and the amount of micro-constituents
micro constituents (phases), which are
controlled by the changes in the alloy composi
composition
tion and mainly by heat treatment. Different heat
treatments bring out a very wide range and distribution of micro-constituents
micro constituents-the variations of which
result in very wide range of properties.

In an abstract way,, it is depicted below.

Purpose of heat treatment

To increase hardness, wear and abrasion resistance


To re-soften
soften after it has been hardened by heat treatment or cold working
To adjust the mechanical, physical and chemical properties such as hardness, yield strength, tensile
strength, electricall and magnetic properties, microstructure or corrosion resistance.
To reduce or eliminate internal residual stresses. Internal stresses lead to premature and brittle
failures of the components. They also reduce corrosion resistance and hence are not desirable.
desira
To induce controlled residual stresses; e.g. compressive stresses on the surface sharply increase the
fatigue life of components.
To decrease or increase the grain size
To eliminate gases, particularly hydrogen that embrittles steel. If the steel is hheld at some elevated
temperature for a short time, these gases will get diffused into the atmosphere.
To change the composition of the surface for steels, i.e, by diffusion of C, N, Si, etc, so as to increase
wear resistance, fatigue life or corrosion resistance
Heating of the component must be uniform and quick. Uniform heating is necessary to reduce distortion and cracking of
the components due to thermal shock. Uniform and Quick heating results in the following advantages
Less oxidation
Less carburization
Less grain growth
More production per furnace
Less cost
Less fuel power

Heat Treatment Techniques for steel

Fig : Iron carbon diagram


Annealing

Annealing, in general refers to heating the material to a predetermined temperature, soaking at this temperature and then
cooling it slowly, normally in a furnace by switching it off. The process consists of heating the steel to above A3
temperature for hypoeutectoid steels and above A1 temperature for hyper eutectoid steels by 30°C to 50°C, holding at this
temperature for a definite period and slow cooling to below A1 or to room temperature usually in the furnace. Due to the
slow cooling, eutectoid reaction occurs very nearly to accordance with the conditions represented by Fe-C diagram. To
ensure equalization of temperature throughout the cross section of the component and complete austenisation, a holding
(soaking) period of at least 20 min per cm of the thickest section is necessary.

Hyper eutectoid steels are always annealed from above A1 temperature and never from Acm temperature because of the
following reasons
• If slowly cooled from above Acm temperature, proeutectoid cementite separates along the grain boundaries of
austenite and completely envelopes the austenite grains which transform to pearlite at A1 temperature. The
structure of the steel at room temperature shows a continuous network of cementite areas all around the pearlite
regions. Such a cementite network provides easy fracture path and renders the steel brittle during deformation or
during service.
• Acm temperature is high and therefore, heating to above Acm temperature results in more oxidation and
decarburization of steel if the proper atmosphere is not maintained.
• Heavy grain coarsening of austenite occurs and this leads to deterioration of mechanical properties.

The aims of annealing the steel could be varying and can be broadly classified into the following categories
1. Improvement in the mechanical properties of the cast or hot worked steels
2. Homogenisation of segregated castings, and ingots of steels- homogenization annealing
3. Restore ductility of cold worked steels-Recrystallization annealing
4. Improve the machinability and cold formability of high carbon steels and alloy steels- Spheroidisation annealing

Spheroidise annealing
This heat treatment is given to high carbon and air hardening alloy steels to soften and to increase machinability. The
microstructure shows globules of cementite or carbides in the matrix of ferrite.
Holding just below A1 temperature
The steels are held for a longer time just below A1 temperature and the cementite from pearlite forms globules. The
process is very slow and requires more time for obtaining spheroidised structures. It can be accelerated by prior cold
working of steel.

Hardening and high temperature tempering


Due to tempering of hardened steels at higher temperatures for a long time, cementite globules are formed in the matrix
of ferrite from martensite.
Martensite→cementite (in globular form) + ferrite

Spheroideised structures are often soft and have excellent machinability. The time of hardening is greatly reduced when
cementite is in fine spheroidised form and uniformly distributed because less time is required to dissolve fine
spheroidised cementite than the almellar cementite to obtain homogeneous austenite. This reduces the soaking time
required during hardening, subsequently reducing the oxidation and decarburization. This is utiliozed in hardening of thin
sections such as razor blades and needles to reduce decarburization. The above materials have spheroidised structure
prior to hardening.

Subcritical annealing
In these processes of annealing, the cold worked steel is heated to some temperatures below the lower critical
temperatures (A1). They are used after cold working of steels to relieve internal stresses or to reduce the hardness or to
refine and modify the structure
Stress relief annealing
In this process, cold worked steel is heated to a temperature below the recrystallisation temperature and soaked for one or
two hours and cooled to room temperature in air. Due to this, internal stresses are partly relieved without the loss of
strength and hardness. It reduces the risk of distortion in machining in machining and also increases corrosion resistance.
Stress relief annealing is also carried out on components in which internal stresses are developed from other sources like
rapid cooling and phase changes.

Recrystallisation annealing

When a metal is cold-worked, by any of the many industrial shaping operations, changes occur in both its physical and
mechanical properties. While the increased hardness and strength which result from the working treatment may be of
importance in certain applications, it is frequently necessary to return the metal to its original condition to allow further
forming operations (e.g. deep drawing) to be carried out for applications where optimum physical properties, such as
electrical conductivity, are essential. The treatment given to the metal to bring about a decrease of the hardness and an
increase in the ductility is known as annealing. This usually means keeping the deformed metal for a certain time at a
temperature higher than about one-third the absolute melting point.
Cold working produces an increase in dislocation density; for most metals _ increases from the value of 1010–1012 lines
m_2 typical of the annealed state, to 1012–1013 after a few per cent deformation, and up to 1015–1016 lines m_2 in the
heavily deformed state. Such an array of dislocations gives rise to a substantial strain energy stored in the lattice, so that
the cold-worked condition is thermodynamically unstable relative to the undeformed one. Consequently, the deformed
metal will try to return to a state of lower free energy, i.e. a more perfect state. In general, this return to a more
equilibrium structure cannot occur spontaneously but only at elevated temperatures where thermally activated processes
such as diffusion, cross slip and climb takes place. Like all non-equilibrium processes the rate of approach to equilibrium
will be governed by an Arrhenius equation of the form
Rate D A exp [_Q/kT]
where the activation energy Q depends on impurity content, strain, etc.

The removal of the cold-worked condition occurs by a combination of three processes, namely: (1) recovery, (2)
recrystallization and (3) grain growth. These stages have been successfully studied using light microscopy, transmission
electron microscopy, or X-ray diffraction; mechanical property measurements (e.g. hardness); and physical property
measurements (e.g. density, electrical resistivity and stored energy). Figure 7.49 shows the change in some of these
properties on annealing. During the recovery stage the decrease in stored energy and electrical resistivity is accompanied
by only a slight lowering of hardness, and the greatest simultaneous change in properties occurs during the primary
recrystallization stage. However, while these measurements are no doubt striking and extremely useful, it is necessary to
understand them to correlate such studies with the structural changes by which they are accompanied.

7.8.2 Recovery
This process describes the changes in the distribution and density of defects with associated changes in physical and mechanical
properties which take place in worked crystals before recrystallization or alteration of orientation occurs. It will be remembered that
the structure of a cold-worked metal consists of dense dislocation networks, formed by the glide and interaction of dislocations, and,
consequently, the recovery stage of annealing is chiefly concerned with the rearrangement of these dislocations to reduce the lattice
energy and does not involve the migration of large-angle boundaries. This rearrangement of the dislocations is assisted by thermal
activation. Mutual annihilation of dislocations is one process. When the two dislocations are on the same slip plane, it is possible that
as they run together and annihilate they will have to cut through intersecting dislocations on other planes, i.e. ‘forest’ dislocations.

This recovery process will, therefore, be aided by thermal fluctuations since the activation energy for such a cutting process is small.
When the two dislocations of opposite sign are not on the same slip plane, climb or cross-slip must first occur, and both processes
require thermal activation. One of the most important recovery processes which leads to a resultant lowering of the lattice strain energy
is rearrangement of the dislocations into cell walls. This process in its simplest form was originally termed polygonization and is
illustrated schematically in Figure 7.50, whereby dislocations all of one sign align themselves into walls to form small-angle or
subgrain boundaries. During deformation a region of the lattice is curved, as shown in Figure 7.50a, and the observed curvature can be
attributed to the formation of excess edge dislocations parallel to the axis of bending. On heating, the dislocations form a sub-boundary
by a process of annihilation and rearrangement. This is shown in Figure 7.50b, from which it can be seen that it is the excess
dislocations of one sign which remain after the annihilation process that align themselves into walls. Polygonization is a simple form
of sub-boundary formation and the basic movement is climb whereby the edge dislocations change their arrangement from a horizontal
to a vertical grouping. This process involves the migration of vacancies to or from the edge of the half-planes of the dislocations.
Figure 7.50 (a) Random arrangement of excess parallel edge dislocations and (b) alignment into dislocation walls;

Recrystallization
The most significant changes in the structure
structure-sensitive
sensitive properties occur during the primary Recrystallization stage. In this stage the
deformed lattice is completely replaced by a new unstrained one by means of a nucleation and growth process, in which practically
stress free grains grow from nuclei formed in the deformed matrix. The orientation of the new grains differs considerably fro
from that of
the crystals they consume, so that the growth process must be regarded as incoherent, i.e. it takes place by the advance of large
large-angle
boundaries separating the new crystals from the strained matrix. During the growth of grains, atoms get transferred from one grain to
another across the boundary.

Grain growth
When primary recrystallization is complete (i.e. when the growing crystals have consumed all the strained material) the mater
material can
lower its energy further by reducing its total area of grain surface. With extensive annealing it is often found that gr
grain boundaries
straighten, small grains shrink and larger ones grow. The general phenomenon is known as grain growth, and the most important factor
tension T (D surface-free energy
governing the process is the surface tension of the grain boundaries. A grain boundary has a surface tension,
per unit area) because its atoms have a higher free energy than those within the grains. Consequently, to reduce this energy a
polycrystal will tend to minimize the area of its grain boundaries

Normalising
The purpose of the process is the same as that of annealing. For hyper eutectoid steels, the process
may also be used to eliminate the cementite network that may have formed due to slow cooling in the
the temperature from Acm to A1.
The process consisits of heating the steel to above upper critical temperatures (A3 for hypo eutectoid
steels and Acm for hyper eutectoid steels) by 30 to 50°C, holding long enough at this temperature for
homogeneous austenization and cooling to room temperature in still air/forced air. Due to air cooling
which Is slightly fast as compared to furnace cooling employed in full annealing, normalized
components show slightly different structure and properties than annealed components. Normalising
takes less time and is more convenient and economical than annealing and hence is a more common
heat treatment in industry. Full annealing is specifically used for complex shapes where even air
cooling may cause cracking or considerable warping of the components.

Hyper eutectoid steels are usually normalized from above Acm temperature and the cementite
seperates in the form of needles in the grains of austenite which transforms to pearlite at A1. This the
microstructure at room temperature shows innumerable needles of cementite in the matrix of pearlite.

Hardening

Purpose : to harden the steel to maximum level by austenite to martensite transformation and to
increase the wear resistance and cutting ability of steel.
The conventional hardening process consists of heating the steel to above A3 temperature for
hypoeutectoid steels and above A1 temperature for hyper eutectoid steels by 50°C, austenizing for
sufficient time and cooling with a rate just exceeding the critical cooling rate of that steel to room
temperature or below room temperature. Due to this usual diffusion transformations are stopped and
the austenite transforms to martensite by diffusionless process.
Hypo eutectoid steels are hardened from above A3 temperatures and not between A1 and A3
temperatures because the phases that exixts at this temperature are proeutectoid ferrite and austenite.
On quenching the steels, only austenite gets transformed to martensite with no change in ferrite. Such
steels show free ferrite in their microstructures and since ferrite is soft phase, the hardness of the
hardened steel gerts reduced. On the other hand, hyper eutectoid steels are always hardened from
temperatures between A1 and Acm temperatures. At this temperature, austenization is not complete
and some proeutectoid cementite will exist along with austenite. Such steels after hardening show
free cementite along with martensite in their microstructures. Since cementie being a hard phase, the
hardness of steels do not get reduced. More over the free cementite do not increse the brittleness of
steel
eel because usually it is fine, well distrbuted and partially spherodoised. Also the grain size remains
fine because cementite particels do not allow to coarsen the austenite.

A proper quenching medium should be used such that the component gets cooled at a rate just
exceeding the critical cooling rate of that steel. Faster cooling than the above also produces
martensite but the tendency of warping and cracking is more and and hence should be avoided. The
critical cooling rate of a steel largely depends on the alloying elements and to a lesser extent on the
carbon percent in the steel. Alloy steels have less critical cooling rate and hence some of the alloy
steels can be hardened by air cooling. Low carbon
carbon steels with very low carbon content cannot be
hardened by quenching because of their high critical cooling rate which is not possible to exceed
even by brine quenching.

Perhaps the most fascinating aspect of steel is that it may be strengthened to amazingly high levels
by quenching. The strength levels are higher than the strongest commercial alloys of aluminum,
copper and titanium by factors of roughly 4.7, 2.2 and 2.1. Steels are generally quenched by
immersing the hot metal into liquid coolants, such as water, oil or liquid salts. Increased strengths do
not occur unless the hot steel contains the austenite phase. The very rapid cooling prevents the
austenite from transforming
forming into the preferred ferrite + cementite structure. A new structure called
martensite* is formed instead, and this martensite phase is responsible for the very high strength
levels. Martensite As explained in Chapter 1, austenite has a face centered cubic (FCC) crystal
structure and ferrite has a body center cubic (BCC) crystal structure. The steel phase diagram shows
us that the FCC structure will dissolve way more carbon than the BCC structure. At the A1
temperature the %C that can dissolve in FCC iiron ron is higher than in BCC iron by the ratio of
0.77/0.02 = 38.5. As discussed above, carbon atoms are much smaller than Fe atoms and the
dissolved C atoms lie in the interstices (holes) between the larger Fe atoms. The FCC structure
dissolves more C atoms because some of the holes in this structure are larger than any of the holes in
the BCC structure. The sketch at the right shows γ iron in a 1060 steel (0.6 %C) transforming to α
iron as the interface (vertical line) moves to your right. After the interfacee has moved, say 1 inch, the
%C in that 1 inch region must drop from 0.6 % to 0.02 %C. At slow cooling rates the carbon is able
to move ahead of the interface into the γ iron along the direction of the dashed arrow by the diffusion
process to be discussed in Chapter 7. However, If one forces this transformation to occur very rapidly
by quenching, there is not enough time for the C atoms to get themselves rearranged and some or all
of them get trapped in the ferrite causing its composition to rise well above 0.02 %, which makes its
crystal structure become distorted from the BCC form. The resulting distorted crystal structure is
martensite. Fig. 4.10 compares the unit cell of BCC ferrite to that of the distorted unit cell of
martensite. What we find is that the unit cell of the martensite crystal is similar to the BCC unit cell
in that it has an atom at its center and one atom at each of the 8 corners. However, the unit cell is no
longer a cube. One of its edges, called the c axis in Fig. 4.10, is longer than the other two, called the a
axes. So the structure is called body centered tetragonal, BCT.
Using x-ray diffraction techniques we can measure the a and c lengths of the unit cell of martensite.
As shown in Fig. 4.11 it is found that as the %C dissolved in the martensite increases the c axis
becomes proportionately larger than the a axis. The increased carbon content in the martensite is
obtained by quenching austenites of higher %C levels. The results of Fig. 4.11 show that as the %C
goes up the resulting distortion from the cubic structure (c gets progressively bigger than a) increases due to
the trapped carbon in the BCT martensite structure. The strength and hardness of the martensite is
found to increase dramatically as the %C increases, as shown by the hardness data of Fig. 4.12. (If you
are not familiar with hardness measurements, see p 38). One way to rationalize this increased hardness is to think
of the chemical bonds holding the Fe atoms together as springs. As the %C increases the springs will
be extended by larger amounts thereby making it more difficult to further extend them, i.e., making
the structure harder.

Martensite structure
The growth of martensite laths, or plates, starting at grain boundaries in the original grain structure, produces a structure
like that shown in these light microscope images. In steel, the "as quenched" structure is very fine and the individual
laths are not well defined. After tempering fine carbide particles are present which are themselves, too small to resolve,
and the BCC iron structure is more easily etched showing the lath arrangement. Martensitic transformations also occur in
other alloy systems such as titanium producing similar structures.

Origin of Quench Cracks in steel


When thick cross sections are quenched drastically, cracks form on the surface of the steel. Distortions or shapes of thin
sections due to plastic deformation known as warping may also occur. Warping and cracking are certainly not desirable.
The presence of surface cracks reduces the fatigue life of the components.

A steep temperature gradient is produced from the surface to the centre of the steel on quenching. Superimposed on this,
are two effects that occur on cooling.
1. Thermal contraction of the steel
2. Volume expansion due to phase transformation (austenite to martensite)
The coefficient of linear thermal expansion for steels is about 11X10-6 K-1. When steels are quenched from 850°C to
25°C, the total contraction will be about 0.9% The % volume expansion during the austenite – martensite transformation
is roughly around 4%. The combined effect leads to volume expansion of the product.
When steels are quenched from the austenizing temperature, the surface of the steel comes more or less immediately to
the bath temperature, whereas the centre is still near the austenizing temperature. The transformation to martensite occurs
at the surface layers and is accompanied by a volume increase. This produces stresses at the centre. The yield strength
(yield stress is the minimum stress required to cause plastic deformation) of austenite is low at high temperatures. So, the
stresses are accompanied by plastic flow of the austenite. Eventually, the centre also cools down to room temperature,
undergoing the martensite transformation. The centre now expands and produces tensile stresses at the surface. The
centre now expands and produces tensile stresses at the surface. These stresses cannot be accommodated by plastic flow
of the surface layers, which already contain transformed martensite. Fresh |Martensite is hard and brittle that it cannot
flow plastically. The residual stress pattern obtained is shown in the figure. The outer layers are in tension and the centre
is in compression. Fig shows that the residual stress distribution is opposite, when only the surface layers are hardened.
In most cases the transformation stresses are dominant and the net residual stresses at the surface are tensile in nature.
Brittle martensite cannot support the tensile stresses and cracks. It should be noted that the residual stresses arise from
steep thermal gradient and the lack of simultaneous transformation throughout the cross section. If the cooling rates are
slower, the temperature gradient is less and the transformation occurs simultaneously, with little residual stresses. To
remove the residual stresses, a stress relieving treatment is usually done by reheating the steel to relatively low
temperature.

Cracks that form at the surface of a steel during quenching due to tensile residual stresses that are produced because of the
volume change that accompanies the austenite-to-martensite transformation
Annealing - A heat treatment process in which a material is heated to an elevated temperature, allowed to dwell there for a set amount of time
and then cooled with a controlled rate.
Stages of annealing:
• Heating to required temperature
• Holding (“soaking”) at constant temperature
• Cooling
The time at the high temperature (soaking time) is long enough to allow the desired transformation (diffusion, kinetics) to occur. Cooling is
done slowly to avoid warping/cracking of due to the thermal gradients and thermo-elastic stresses within the or even cracking the metal piece.

Purposes of annealing:
• Relieve internal stresses
• Increase ductility, toughness, softness
• Produce specific microstructure

Process Annealing - used to revert effects of workhardening (by recovery and recrystallization) and to increase ductility. Heating is
usually limited to avoid excessive grain growth and oxidation.
Stress Relief Annealing – used to eliminate/minimize stresses arising from Plastic deformation during
Machining,Non-uniform cooling, Phase transformations between phases with different densities
Stress relief annealing allows these stresses to relax.
Annealing temperatures are relatively low so that useful effects of cold working are not eliminated.
To produce recrystallization one must heat the cold worked metal up to a temperature called the
recrystallization temperature. The Recrystallization temperature depends on the amount of prior
deformation that was used. More deformation increases the defect density and lowers the
recrystallization temperature. However, as shown in Fig. 8.7, the drop in Recrystallization
temperature stops after around a 50 % reduction. This results because the increase in defect density
saturates at a maximum value for deformations larger than around 50 %. Therefore, if one talks about
the recrystallization temperature of a given alloy, it is assumed that the deformation used is roughly
50% or more. Another complication is the time for recrystallization.
Recrystallization will occur faster at higher temperatures. So data such as that of Fig. 8.7 are
generally presented for the 1 hour recrystallization temperature. That temperature is defined as the
temperature where recrystallization is complete in roughly 1 hour. An important practical question
when trying to produce recrystallization is how much cold work is needed to cause recrystallization
to occur. The data of Fig. 8.7 show that in low carbon steels recrystallization will occur with only
around 5% cold work, but the temperature needed is significantly higher than required with larger
amounts of cold work.

Illustration: The temperature at which a metal recrystallizes depends on its melting point. Tin melts at a low
temperature, 232oC (450 oF), and therefore in alloys such as soft solder recrystallization can occur at room temperature.
Take a thick copper wire and bend it back and forth and it will harden at the bend and break there. Do the same for the
wire from a roll of soft solder and you find you can continue to bend it almost indefinitely with no hardening or breaking.
The soft solder is recrystallizing in your hands at room temperature, producing new low defect grains that are soft, while
the copper retains its original grains and they progressively become harder and brittle as the defect density increases on
bending.

TEMPERING

After the hardening treatment is applied, steel is often harder than needed and is too brittle for most practical uses. Also,
severe internal stresses are set up during the rapid cooling from the hardening temperature. To relieve the internal stresses
and reduce brittleness, you should temper the steel after it is hardened. Tempering consists of heating the steel to a
specific temperature (below its hardening temperature), holding it at that temperature for the required length of time, and
then cooling it, usually instill air. The resultant strength, hardness, and ductility depend on the temperature to which the
steel is heated during the tempering process. The purpose of tempering is to reduce the brittleness imparted by hardening
and to produce definite physical properties within the steel. Tempering always follows, never precedes, the hardening
operation. Besides reducing brittleness, tempering softens the steel. That is unavoidable, and the amount of hardness that
is lost depends on the temperature that the steel is heated to during the tempering process. That is true of all steels except
high-speed steel. Tempering increases the hardness of high-speed steel. Tempering is always conducted at temperatures
below the low-critical point of the steel. In this respect, tempering differs from annealing, normalizing, and hardening in
which the temperatures are above the upper critical point. When hardened steel is reheated, tempering begins at 212°F
and continues as the temperature increases toward the low-critical point. By selecting a definite tempering temperature,
you can predetermine the resulting hardness and strength. The minimum temperature time for tempering should be 1
hour. If the part is more than 1 inch thick, increase the time by 1 hour for each additional inch of thickness. Normally, the
rate of cooling from the tempering temperature has no effect on the steel. Steel parts are usually cooled in still air after
being removed from the tempering furnace; however, there are a few types of steel that must be quenched from the
tempering temperature to prevent brittleness. These blue brittle steels can become brittle if heated in certain temperature
ranges and allowed to cool slowly. Some of the nickel chromium steels are subject to this temper brittleness. Steel may
be tempered after being normalized, providing there is any hardness to temper. Annealed steel is impossible to temper.
Tempering relieves quenching stresses and reduces hardness and brittleness. Actually, the tensile strength of a hardened
steel may increase as the steel is tempered up to a temperature of about 450°F. Above this temperature it starts to
decrease. Tempering increases softness, ductility, malleability and impact resistance. Again, high-speed steel is an
exception to the rule. High-speed steel increases in hardness on tempering, provided it is tempered at a high temperature
(about 1550°F). Remember, all steel should be removed from the quenching bath and tempered before it is completely
cold. Failure to temper correctly results in a quick failure of the hardened part.

The cooling rate of an object depends on many things. The size, composition, and initial temperature of the part and final
properties are the deciding factors in selecting the quenching medium. A quenching medium must cool the metal at a rate
rapid enough to produce the desired results. Mass affects quenching in that as the mass increases, the time required for
complete cooling also increases. Even though parts are the same size, those containing holes or recesses cool more
rapidly than solid objects. The composition of the metal determines the maximum cooling rate possible without the
danger of cracking or warping. This critical cooling rate, in turn, influences the choice of the quenching medium. The
cooling rate of any quenching medium varies with its temperature; therefore, to get uniform results, you must keep the
temperature within prescribed limits. The absorption of heat by the quenching medium also depends, to a large extent, on
the circulation of the quenching medium or the movement of the part. Agitation of the liquid or the part breaks up the gas
that forms an insulating blanket between the part and the liquid. Normally, hardening takes place when you quench a
metal. The composition of the metal usually determines the type of quench to use to produce the desired hardness. For
example, shallow-hardened low-alloy and carbon steels require severer quenching than deep-hardened alloy steels that
contain large quantities of nickel, manganese, or other elements. Therefore, shallow-hardening steels are usually
quenched in water or brine, and the deep-hardening steels are quenched in oil. Sometimes it is necessary to use a
combination quench, starting with brine or water and finishing with oil. In addition to producing the desired hardness, the
quench must keep cracking, warping, and soft spots to a minimum.

The volume of quenching liquid should be large enough to absorb all the heat during a normal quenching operation
without the use of additional cooling. As more metals are quenched, the liquid absorbs the heat and this temperature rise
causes a decrease in the cooling rate. Since quenching liquids must be maintained within definite temperature ranges,
mechanical means are used to keep the temperature at prescribed levels during continuous operations.

Liquid Quenching
The two methods used for liquid quenching are called still-bath and flush quenching. Instill-bath quenching, you cool the
metal in a tank of liquid. The only movement of the liquid is that caused by the movement of the hot metal, as it is being
quenched. For flush quenching, the liquid is sprayed onto the surface and into every cavity of the part at the same time to
ensure uniform cooling. Flush quenching is used for parts having recesses or cavities that would not be properly
quenched by ordinary methods. That assures a thorough and uniform quench and reduces the possibilities of distortion.
Quenching liquids must be maintained at uniform temperatures for satisfactory results. That is particularly true for oil. To
keep the liquids at their proper temperature, they are usually circulated through water-cooled Figure 2-3.—Portable
quench tank. coils. Self-contained coolers are integral parts of large quench tanks. A typical portable quench tank is
shown in figure 2-3. This type can be moved as needed to various parts of the heat-treating shop. Some tanks may have
one or more compartments. If one compartment contains oil and the other water, the partition must be liquid-tight to
prevent mixing. Each compartment has a drain plug, a screen in the bottom to catch scale and other foreign matter, and a
mesh basket to hold the parts. A portable electric pump can be attached to the rim of the tank to circulate the liquid. This
mechanical agitation aids in uniform cooling.

Water
Water can be used to quench some forms of steel, but does not produce good results with tool or other alloy steels. Water
absorbs large quantities of atmospheric gases, and when a hot piece of metal is quenched, these gases have a tendency to
form bubbles on the surface of the metal. These bubbles tend to collect in holes or recesses and can cause soft spots that
later lead to cracking or warping. The water in the quench tank should be changed daily or more often if required. The
quench tank should be large enough to hold the part being treated and should have adequate circulation and temperature
control. The temperature of the water should not exceed 65°F.
When aluminum alloys and other nonferrous metals require a liquid quench, you should quench them in clean water. The
volume of water in the quench tank should be large enough to prevent a temperature rise of more than 20°F during a
single quenching operation.

Brine is the result of dissolving common rock salt in water. This mixture reduces the absorption of atmospheric
gases that, in turn, reduces the amount of bubbles. As a result, brine wets the metal surface and cools it more rapidly than
water. In addition to rapid and uniform cooling, the brine removes a large percentage of any scale that may be present.
The brine solution should contain from 7% to 10% salt by weight or three-fourths pound of salt for each
gallon of water. The correct temperature range for a brine solution is 65°F to 100°F. Low-alloy and carbon steels can be
quenched in brine solutions; however, the rapid cooling rate of brine can cause cracking or stress in high-carbon or low-
alloy steels that are uneven in cross section. Because of the corrosive action of salt on nonferrous
metals, these metals are not quenched in brine.

The importance of tempering will be illustrated by describing a simple experiment that can be done if
one wants to obtain a hands-on appreciation of heat treatment. The experiment is done with two
lengths (each around 6 to 12 inches) of drill rod having a diameter of 1/8 inches. Drill rod is either 1095
steel or W1 tool steel, which have essentially the same composition. One inch long sections of the
two rods are simultaneously heated with a propane torch to an orange color for around 20 seconds
and then immediately quenched in a glass of water. The temperature produces austenite over a short
length of the rods and the quench converts it to martensite + retained austenite. The first rod is placed
in a vice with the discolored heated section about 1/2 inch from the grips and struck firmly with a
hammer. The piece will break across the center of the heated section in a brittle manner, just like a
piece of glass. The discolored heated section of the second rod is polished shinny with a fine emery
or sand paper. It is now gently heated with the torch until it turns a deep blue color, this is the
tempering step. Bending this rod in the vice or with a set of pliers produces the result shown in Fig.
10.1 Martensite has formed in the heated zone, and if you look carefully at a severely bent rod you
will be able to see the boundary between the martensite and non martensite, as shown at the arrows
in Fig. 10.1. The non martensite region has the as-received spheroidized structure of Fig. 4.23. This
structure is soft and the metal flow within it on bending produces a surface distortion that terminates
at the hard martensite zone and reveals the transition boundary.

If the tempering temperature is hot enough, the rod will not break through the martensite region even
when this region is bent to over 90 degrees by hammering in the vice. If it is too hot the martensite
region will become softer than one might like. Upon heating the polished steel rod, an oxide layer
forms on its surface and the thickness of the oxide layer produces various colors due to an
interference effect of light rays bouncing off the bottom and top of the oxide layer. The oxide
thickness is controlled by the temperature of the steel and hence a series of different colors occurs at
the relatively low temperatures used for tempering. The temperatures corresponding to the various
colors are know as temper colors and Table 10.1 presents the correlation between colors and
temperatures. The color temperature has been used for millennia by blacksmiths when tempering
steels

The experiment shows that in the as-quenched condition the steel is way too brittle to be useful for
applications other than those which require extreme hardness with no bending, such as files. Hence,
virtually all quenched steels are tempered. Figure 10.2 shows qualitatively how the quenching and
tempering operation changes the stress-strain characteristics of the steel. The tempering operation
sacrifices the high strength of the steel to gain back improvements in ductility and toughness.

Several things are going on in the steel during the tempering process that result in the loss of strength
and the gain in ductility and toughness. The first thing that happens is the relief of the
high degree of volume strain in the steel produced by the formation of the higher volume per atom
martensite phase. Then a series of internal structure changes occurs which is generally partitioned
into the 3 stages of tempering.

Stage 1: This first stage consists of the formation of very small carbides in the martensite, so small
(around 10 nm) they can only be seen in an electron microscope. These first formed carbides are
metastable carbides (do not appear on the equilibrium phase diagram), epsilon carbide (Fe2.4C) in
hypoeutectoid steels and Hagg (Fe2.2C) and eta (Fe2C) carbides in hypereutectoid steels. Stage 2:
This stage is simply the decomposition of any retained austenite into carbides and ferrite. It is only
important in high carbon steels where % retained austenite is significant. Stage 3: This stage occurs
at the highest tempering temperatures and here the metastable carbides are replaced with small
particle of cementite, the stable carbide of steels.
As explained in Chapter 4, toughness is a better measure of the ability of a steel to avoid failure in
practice than is ductility; and, toughness is evaluated with impact tests, such as the Charpy test. The
Izod test is similar to the Charpy test and Fig. 10.4 presents a summary of scatter bands of data from
such tests on steels of two different levels of %C. These curves illustrate two important
characteristics of heat treated steels:

Quenching a piece of hot steel in water or oil produces copious amounts of vapor around the piece,
often called the "vapor blanket". In water the blanket is steam and in oil it is vaporized oil.
The presence of this vapor phase around the hot steel leads to the very complex mode of heat
transfer that occurs during quenching. There has been considerable research on this problem in the
latter decades of the 20th century and a method developed in England has now been adopted as an
international standard for characterizing quench fluids, ISO 9950. The test utilizes an Inconel 600
alloy cylinder (basically the same Ni-Cr alloy as used for the heating elements of your electric stove), 12.5 mm in
diameter by 60 mm long. A metal clad type K thermocouple (see Appendix A) is fitted into a hole along
its center and the output of this thermocouple is monitored during the quench. The output produces a
cooling curve such as that shown for an oil bath in Fig. 12.10. The heat transfer during the quench
can be partitioned into 3 stages, conventionally called A, B and C. Initially, during the A stage, heat
transfer is relatively slow as the heat must pass through the vapor blanket that initially surrounds the
immersed sample. Notice that the B stage begins with a rapid increase in the rate at which the
temperature drops. When the oil (or water) begins to penetrate through the vapor blanket it contacts
the hot steel and immediately boils. The heat required to boil the liquid is removed from the steel and
this mode of heat transfer is extremely efficient. It is called "nucleate boiling heat transfer", hence the
name of stage B shown on Fig. 12.10. When the boiling stops heat is transferred directly to the liquid
touching the steel causing its temperature to increase which, in turn, drops its density. Hence, this
liquid rises and is replaced by colder liquid contacting the steel piece. The motion of the liquid is
called convection and hence the name of stage C shown on Fig. 12.10.

For evaluating quenching intensity we are mainly interested in how fast the temperature is falling
and, hence, the most useful parameter is the cooling rate, which will have units of oC/sec. (or oF/sec.).
Therefore, it has become common to characterize the quenching power of a quenchant with a plot of
cooling rate versus temperature of the Inconel rod, and Fig. 12.11 is presented to show you how such
curves are related to the simple cooling curve of Fig. 12.10. In order to obtain the cooling rate at 400
oC, one constructs a tangent line to the cooling curve at this temperature as shown. Then, moving
Δ Δ
from the bottom of this line to the top one measures the rise, T, and the run, t. The ratio of the
Δ Δ
rise divided by the run, T/ t (called the slope in geometry classes) will have units of oC/sec. and will be
the cooling rate when the Inconel center is at 400 oC. With computer software it is a simple matter to
determine the cooling rate at each temperature and the solid line is the corresponding cooling curve
with the cooling rates in oC/sec. given along the top of the diagram. Notice that the maximum
cooling rate is 70 oC/sec. and it occurs at around 650 oC for this oil. There is an excellent discussion
of water, oil and polymer quenchants in reference [12.8], and you will see on pages 78 to 81 that
these cooling rate curves are now widely used to characterize the quenching power of the various
quenching oils and polymer quenchants that are commercially available. Figure 12.12 presents
cooling rate curves for several common quenching fluids. Notice that the maximum cooling rate
occurs at different temperatures and varies from a high of 285 oC/sec. for salt water to a low of 65
oC/sec. for a normal oil.

For a given quenchant, the speed of the quenching process will depend on the temperature of the bath
as well as any agitation used on the work piece during the quenching operation. One can slow down
the quench rate of water significantly by heating the water to elevated temperatures. The cooling rate
curves given in [12.8] show that the maximum cooling rate for water is decreased from 225 oC/sec.
for a bath at 20 oC (68 oF) to only 90 oC/sec. for a bath at 80 oC (176 oF), a decrease of 60 %. The
cooling rate data given in [12.8] for hot oils show that the maximum cooling rate is less sensitive to
bath temperature and varies in a more complicated manner. As the bath temperature is changed from
200 to 150 to 50 oC (392, 302, 122 oF), a maximum cooling rate of 83 oC/sec. occurs at the intermediate
bath temperature of 150 oC and drops to 80 oC/sec. at the higher temperature of 200 oC, and to 75
oC/sec. at the lowest bath temperature of 50 oC.

You might also like