You are on page 1of 20

Heat Treatment of Steels

A steel is usually defined as an alloy of iron and carbon with the carbon content
between a few hundreds of a percent up to about 2 wt%. Other alloying elements can
amount in total to about 8 wt% in low alloy steels and higher in more highly alloyed
steels such as tool steels and stainless steels. The strength and hardness of steels
depend on their carbon and other alloying elements. In general, increase in carbon and
alloy contents increase in their hardness and strength. However, for a fixed chemical
composition, the mechanical, physical and chemical properties of steels can be
changed to a wide scale by heat treatment. Heat treatment is a process of heating and
cooling. Depending on the heating rate, holding period at the selected high temperature
and subsequent cooling rate from the hold down (high) temperature (Figure 1) control
the microstructures as well as the properties of steels.

Figure 1: Heat treatment cycles.

There are several types of heat treatment, however, the most common heat
treatments used for the modification of steel properties are:

A. Annealing B. Normalizing,
C. Hardening D. Tempering

Annealing
Annealing is a heat treatment process where a material is exposed to a moderate
to elevated temperature for an extended time period and then slowly cooled, generally
in the furnace. Ordinarily, annealing is carried out to (1) relieve stresses; (2) increase
softness, ductility and toughness, and/or (3) produce a specific microstructure. A variety
of annealing heat treatments are possible. Any annealing process consists of three
stages: (1) heating to the desired temperature, (2) holding or “soaking” at that
temperature and (3) slow cooling, usually to room temperature. Time is an important
parameter in these procedures. During heating and cooling, there exists temperature
gradients between the outside and interior portions of the piece; their magnitudes
depend on the size and geometry of the piece. If the rate of temperature change is too
great, temperature gradients and internal stresses may be induced that may lead to
warping or even cracking. Also, the actual annealing time must be long enough to allow
for any necessary transformation reactions. Annealing might be of the following four
categories:
A. Process annealing B. Stress relief annealing
C. Spheroidizing D. Full annealing

Process Annealing
Process annealing is a heat treatment that is used to negate the effects of cold
work, that is, to soften and increase the ductility of a previously strain-hardened metal. It
is commonly utilized during fabrication procedures that require extensive plastic
deformation, to allow a continuation of deformation without fracture or excessive energy
consumption in the production process. As this type of annealing is practiced during
industrial process, so it is named as process annealing. Process annealing is a very
useful heat treatment for production of thin wire and sheet products, where low to
medium carbon steels are used, Figure 2.

Figure 2: Various annealing treatment for carbon steel (left) and effect of process
annealing on microstructures of steel (right).

Here, recovery and recrystallization processes are allowed to occur in the place of
elongated and deformed grains, Figure 2 (right). Ordinarily a fine grained microstructure
is desired, therefore, the heat treatment is terminated before appreciable grain growth.
Since the material stays in the same phase throughout the process, the only change
that occurs is the size, shape and distribution of the grain structure. This process is
cheaper than either full annealing or normalizing since the material is not heated to a
very high temperature or cooled in a furnace. It has been mentioned that this annealing
heat treatment is usually performed during thin wire and sheet production, where
surface oxidation at high temperature might deteriorate the properties of the product. As
a result, process annealing is usually carried out at relatively lower temperature. If high
temperature is essential during any process annealing nonoxidizing atmosphere should
be ensured.

Stress Relief Annealing


Internal residual stresses may develop in metal pieces in response to the following:
(1) plastic deformation processes such as machining and grinding; (2) nonuniform
cooling of a piece that was processed or fabricated at an elevated temperature, such as
a weld or a casting and (3) a phase transformation that is induced upon cooling wherein
parent and product phases have different densities. Distortion and warpage may result if
these residual stresses are not removed. They may be eliminated by a stress relief
annealing heat treatment in which the piece is heated to the recommended
temperature, held there long enough to attain a uniform temperature and finally cooled
to room temperature in air. The annealing temperature is ordinarily a relatively low one
such that effects resulting from cold working and other heat treatments are not affected.
The stress relief heat treatment is usually used for medium carbon steels and the
temperature selected for this is below the lower critical temperature the steel. Some
products that need stress relief annealing are shown below (Figure 3).

Figure 3: Some products that need stress relief heat treatment.

Spheroidizing
Medium and high carbon steels having a microstructure containing even coarse
pearlite may still be too hard to conveniently machine or plastically deform. These
steels, and in fact any steel, may be heat treated or annealed to develop the spheroidite
structure. Spheroidized steels have a maximum softness and ductility and are easily
machined or deformed. The spheroidizing heat treatment, during which there is a
coalescence of the Fe3C to form the spheroid particles, Figure 4.
Figure 4: Microstructures of high carbon steel before (left and middle) and after
spherodizing annealing (right).

Spherodizing heat treatment can be carried out in several ways as follows:


 Heating the alloy at a temperature just below the eutectoid line (lower critical line)
in the region of the Fe-C phase diagram. If the precursor microstructure contains
pearlite, spheroidizing times will ordinarily range between 15 and 25 h.
 Heating to a temperature just above the eutectoid temperature and then either
cooling very slowly in the furnace, or holding at a temperature just below the
eutectoid temperature.
 Heating and cooling alternately within about of the eutectoid line.

Spheroidizing annealing is usually used for high carbon steels (carbon > 0.6%). To
some degree, the rate at which spheroidite forms depends on prior microstructure. For
example, it is slowest for coarse pearlite and with the fineness of pearlite grains the
spherodization rate increases. Also, prior cold work increases the spheroidizing reaction
rate.

Full Annealing
It is a heat treatment technique used for hypoeutectoid steels (low and medium
carbon steels) that will be machined or will experience extensive plastic deformation
during a forming operation or for as cast products where the mechanical property is
inferior because of coarse grained structure with inhomogeneous chemical
compositions, Figure 5. This heat treatment is also used to relieve internal stresses
induced by rolling, forging or to remove coarseness of grains in the as cast products. In
general, the alloy is treated by heating to a temperature of about 40-50oC above the A3
line to form austenite fully. In the case of hyper eutectoid steel, this heat treatment is not
usually practiced, where spheroidizing annealing is preferred. However, if necessary, a
temperature 40-50oC above the A1 line is selected (to form austenite and Fe3C phases).
It is held at this temperature for sufficient time for all the material to transform into
austenite or austenite-cementite as the case may be. It is then slowly cooled at a rate of
about 20ºC/hr in a furnace to about 50ºC into the ferrite-cementite range. At this point, it
can be cooled in room temperature air with natural convection. The grain structure
might be coarse pearlite with ferrite or cementite (depending on whether hypo or hyper
eutectoid). The steel becomes soft and ductile. It is possible to calculate upper and
lower critical temperatures using the actual chemical composition of the steel. The
following equations will give an approximate critical temperature for a hypoeutectoid
steel:
Ac1(°C) = 723 - 20.7(%Mn) - 16.9(%Ni) + 29.1(%Si) - 16.9(%Cr)
Standard deviation = ± 11.5 °C
Ac3(°C) = 910 - 203%C - 15.2(%Ni) + 44.7(%Si) + 104(%V) + 31.5(%Mo)
Standard deviation = ± 16.7 °C

Figure 5 : Effect of full annealing on microstrutural changes in steel.

When the steel is heated well above the upper critical temperature large austenite
crystals form. Slow cooling gives rise to the Widmanstätten type of structure, with its
characteristic lack of both ductility and resistance to shock. This is known as an
overheated structure and it can be refined by reheating the steel to just above the upper
critical point. Surface decarburization usually occurs during the overheating.
Normalizing
Normalizing is accomplished by heating at least 55-60oC (Figure 6) above the
upper critical temperature for all steels and after soaking at the normalizing temperature
for specific time period, the object is cooled in still air. The soaking time should be
chosen properly so that the alloy is completely transformed to austenite.

Figure 6: Normalizing heat treatment for steel.

Soak periods for normalizing are typically one hour per inch of cross-sectional
area, but not less than two hours holding at the austenitizing temperature. It is important
to remember that the mass of the part or the work load can have a significant influence
on the cooling rate and thus on the resulting microstructure. Thin pieces cool faster and
are harder after normalizing than thicker ones. By contrast, after furnace cooling in an
annealing process, the hardness of the thin and thicker sections will be of almost
similar.

Normalizing vs. Annealing


Normalizing differs from annealing in that the metal is heated to a higher
temperature and then removed from the furnace for air cooling rather than furnace
cooling that is performed in the case of annealing. This is because normalizing is a final
heat treatment in most cases, whereas annealing in an intermediate heat treatment.
However, in practical applications, for many manufacturing engineers, there is often a
great deal of confusion as to when to specify normalizing and when to call out
annealing. There is a logical reason for this because, in many instances, the procedure
for normalizing and that of annealing are one and the same. For example, very-low-
carbon steel can be almost fully annealed by heating above the transformation range
and cooling in air. In normalizing, the cooling rate is slower than that of a quench-
temper operation, but faster than that used in annealing. As a result of this intermediate
cooling rate, the parts will possess hardness and strength somewhat greater than if
annealed but somewhat less than if quenched and tempered. The slower cooling rate
means normalized sections will not be as highly stressed as quenched sections. Thus,
normalizing is a treatment where a moderate increase in strength is achieved without
undue increase in stress.
Annealing and normalizing do not present a significant difference on the ductility of
low-carbon steels. As the carbon content increases, however, annealing heat treatment
ensures significantly higher level of ductility (20%). On the other hand, the ductility of
the normalized high-carbon steels continues to drop to the 1–2% level (Figure 7).

Figure 7: Comparison of tensile properties of annealed and normalized steels.

Normalizing also improves microstructural homogeneity and refines the grain


sizes, Figure 8. For hypereutectoid steels, normalizing breaks down the continuity of the
cementite networks and induces higher strength as well as toughness. This type of
grain refinement also contributes in strengthening the steel. As a result, in general,
normalized steel a fixed chemical composition is stronger than that in the annealed
condition.
Annealed 0.4%C Steel Normalized 0.4%C Steel

Annealed 1.0%C Steel Normalized 1.0%C Steel


Figure 8: Microstructures of various annealed and normalized steels.

Another important difference in heat treatment parameter of normalizing and


annealing is the temperature selection for hypereutectoid steel as seen in the diagram,
which is above the upper critical temperature for normalizing. But for annealing this
temperature is above the lower critical temperature. The main reason is that annealing
is usually performed as an intermediate heat treatment to make the subsequent
operations as machining, homogenization of chemical compositions, removal residual
stress, etc easy. Annealing at just above the lower critical temperature provide the
necessary solution. As normalizing is a final heat treatment, normalizing above lower
critical temperature makes the product soft. On the other hand normalizing at this
temperature might also form network type cementite, which is not beneficial for any load
bearing component as network type cementite deteriorates both the strength and
ductility of hyper eutectoid steels.

Hardening of Steels
It has been mentioned that strength or hardness of steel is a function of the carbon
as well as other alloying element contents in it. However, the strength and hardness of
the steel can be significantly increased by changing the microstructures and also crystal
structures. We know that steel at room temperature has BCC structure along with
ferrite-pearlite or pearlite-cementite microstructures depending on the carbon content of
the steel. When steel is heated to above the upper critical temperature, it exhibits FCC
structure with single phase austenitic microstructure, Figure 9.
Figure 9: Various crystal structures and microstructures of steels.

Figure 10: Crystal structures of BCC, FCC and BCT (martensite)

We know that the empty space in BCC crystal is 32%, whereas it is 26% in FCC.
However, carbon solubility in FCC is much higher than that in BCC, which is 0.2% in
high temperature ferrite (delta ferrite having BCC crystal). From this difference in carbon
solubility is just not for temperature difference. There is no atom inside the body of FCC
unit cell, i.e. it has one big gap that can accommodate carbon atoms easily. On the
other hand, because of presence of one atom inside the body of BCC unit cell, the total
32% empty space is at different locations rather than at a single location, which is not
capable to accommodate carbon atom easily. This is the reason, why C solubility in
FCC cell is higher than that of BCC cell. Higher temperature means higher solubility,
which is the reason why delta ferrite has higher C solubility than ferrite. For hardening
of steel, it is heated to austenitic region where it gains FCC structure. From the Fe-C
diagram, it is clear that solubility of carbon in austenite is higher than that of the ferrite.
In austenite, the carbon content becomes the average carbon content of double phase
structures at room temperature. When the steel with austenite phase (FCC) is suddenly
cooled, all the carbon cannot come out from its body, which induces severe stress in the
structures. Because of this severe stress caused by entrapped carbon in it, one side of
the cubic structure becomes elongated. This deformed BCC structure is called BCT
(Figure 10) structure and the resultant microstructure is termed as martensite, which is
very strong and brittle.
Martensite is a nonequilibrium single-phase structure that results from a
diffusionless transformation of austenite. The martensitic transformation occurs when
the quenching rate is rapid enough to prevent carbon diffusion. Any diffusion
whatsoever will result in the formation of ferrite and cementite phases. Since the
martensitic transformation does not involve diffusion, it occurs almost instantaneously;
the martensite grains nucleate and grow at a very rapid rate-the velocity of sound within
the austenite matrix. Thus the martensitic transformation rate, for all practical purposes,
is time independent. The martensite grains thus formed are of plate-like or needle-like
appearance, Figure 11.

Figure 11: Microstructures of martensite.


CR

Figure 12: Effect of cooling rate on microstructures of steel during cooling from
high temperature.

Being a nonequilibrium phase, martensite does not appear on the iron–iron carbide
phase diagram. The austenite-to-martensite transformation is, however, represented on
the isothermal transformation diagram. Martensite is usually formed by drastic cooling
(quenching). Depending on cooling rate there might be other phases present, Figure 12.
The minimum cooling rate that might ensure 100% martensite in the quenched steel is
call critical cooling rate for that steel. Here it is to be mentioned that CR is dependent on
chemical compositions of the steel component. Not only chemical compositions, shape,
size and geometry of the steel products also control their hardenability. Thinner
products harden easily. At the same time, sharp edges of a product harden more easily,
Figure 13.

Figure 13: Steel products of various shapes, sizes and geometries


Since the martensitic transformation is diffusionless and instantaneous, it is not
depicted in this diagram as the pearlitic and bainitic reactions are. The beginning of this
transformation is represented by a horizontal line designated M(start), Figure 12. Two
other horizontal and dashed lines, labelled M(50%) and M(90%), indicate percentages
of the austenite-to-martensite transformation. Above the M(start) line, there three S type
curves. Before outer one (left) and above the horizontal line, austenite phase will remain
stable. After right most line, 100% austenite will be transformed to either pearlite (P or
P+F) or bainite (B) depending on cooling rate (location at where the cooling curves
touch the transformation curve (S curves). For hardening, in the ferrite-pearlite
structures martensite needs to be formed, which is achieved by rapid cooling (called
quenching). Martensite formation can be achieved by two different cooling modes,
namely continuous (Figure 12) and continuous cooling (Figure 14).

Figure 14: Isothermal cooling during quenching.

It important to note that iron–carbon alloys containing less than about 0.25 wt%
carbon are not normally heat treated to form martensite because required cooling rate is
too rapid, which is not practically possible in most of the cases. However, other alloying
elements such as chromium, nickel, molybdenum, manganese, silicon and tungsten, etc
help to achieve CR within the possible practical cooling rate.
Hardenability
The influence of alloy compositions on the ability of a steel alloy to transform to
martensite for a particular quenching treatment is related to a parameter called
hardenability. For every different steel alloy there is a specific relationship between the
mechanical properties and the cooling rate. “Hardenability” is a term that is used to
describe the ability of an alloy to be hardened by the formation of martensite as a result
of a given heat treatment. Hardenability is not “hardness,” which is the resistance to
indentation; rather, hardenability is a qualitative measurement of depth up to which hard
martensitic structure is formed, Figure 15.

Figure 15: Variation in hardness from surface to the core of hardened object.

A steel alloy that has a high hardenability is one that hardens or forms martensite,
not only at the surface, but to a large degree throughout the entire interior. The
quenched end is cooled most rapidly and exhibits the maximum hardness; 100%
martensite is the product at this position for most steels. Cooling rate decreases with
distance from the quenched end and the hardness also decreases, Figure 15. With
diminishing cooling rate more time is allowed for carbon diffusion and the formation of a
greater proportion of the softer pearlite, which may be mixed with martensite and
bainite. Thus, steel that is highly hardenable will retain large hardness values for
relatively long distances; a low hardenable one will not.

Figure 16: Effect of alloying elements on Ms temperatures of steels.

It has been observed that increase in the carbon content also increases the
hardenability. Most of other alloying elements which enter into solid solution in austenite
lower the martensite start temperature (Ms), with the exception of Co and Al (Figure 15).
However, the interstitial solutes carbon and nitrogen have a much larger effect than the
metallic solutes. It has been found that 1 wt% of carbon lowers the Mf (martensite finish
temperature) by over 300°C. Note that above 0.7 wt% C the Mf temperature is below
room temperature and consequently higher carbon steels quenched into water will
normally contain substantial amounts of retained austenite. The hardenability curves of
some common alloys are shown in the Figure 17.
1040: 0.4%C, 5140: 0.4%C, 0.85Cr

8640: 0.4%C, 0.55Ni, 0.5Cr, 0.2Mo

4140: 0.4%C, 1.0Cr, 0.2Mo

4340: 0.4%C, 1.85Ni, 0.8Cr, 0.25%Mo)

Figure 17: Effect of alloying elements on hardenability of steels.

Quenching Media
Quenching is the act of rapidly cooling the hot steel to harden it. Depending on
chemical compositions of the steel part to be hardened, its shape and size, required
level of hardness, quenching can be performed in various media some of which are
discussed in below.

Water: Quenching can be done by plunging the hot steel in water. The water adjacent
to the hot steel vaporizes and there is no direct contact of the water with the steel. This
slows down cooling until the bubbles break and allow water contact with the hot steel.
As the water contacts and boils, a great amount of heat is removed from the steel. With
good agitation, bubbles can be prevented from sticking to the steel and thereby prevent
soft spots. Water is a good rapid quenching medium, provided good agitation is done.
However, water is corrosive with steel and the rapid cooling can sometimes cause
distortion or cracking.

Salt Water: Salt water is a more rapid quench medium than plain water, because the
bubbles are broken easily and allow for rapid cooling of the part. However, salt water is
even more corrosive than plain water and hence must be rinsed off immediately after
quenching.

Oil: Oil is used when a slower cooling rate is desired. Since oil has a very high boiling
point, the transition from start of martensite formation to the finish is slow and this
reduces the likelihood of cracking. Oil quenching results in fumes, spills and sometimes
a fire hazard.

Polymer Quench: Polymer quenches that will produce a cooling rate in between water
and oil. The cooling rate can be altered by varying the components in the mixture-as
these are composed of water and some glycol polymers. Polymer quenches are
capable of producing repeatable results with less corrosion than water and less of a fire
hazard than oil. But, these repeatable results are possible only with constant monitoring
of the chemistry. The cooling rates of steel component in various cooling medium are
shown in Figure 18.

Figure 18: Effect of quenching media on cooling rate of steel component.

Cryogenic Quench
Cryogenics or deep freezing is done to make sure there is no retained austenite
during quenching. The amount of martensite formed at quenching time is a function of
the lowest temperature encountered. At any given temperature of quenching there is a
certain amount of martensite and the balance is untransformed austenite. This
untransformed austenite is very brittle and can cause loss of strength or hardness,
dimensional instability or cracking. Quenches are usually done to take the high
temperature components to room temperature. Most medium carbon steels and low
alloy steels undergo transformation to 100% martensite at room temperature. However,
high carbon and high alloy steels retain a fraction of austenite at room temperature. To
eliminate retained austenite, the quench temperature has to be lowered. This is the
reason to use cryogenic quenching.

Tempering
For hardening, steel is quenched from austenite phase to form martensitic
structure. Essentially, martensite is a highly supersaturated solid solution of carbon in
iron with high level of residual stress, which is very brittle having a poor toughness. So,
components with martensitic structures of high residual stress have a little application.
As a result, it is essential to reduce residual stress and brittleness of the component,
which is done by tempering. Tempering is a heat treatment process which is done
below the lower critical temperature. Depending on the geometry of the components
and their chemical compositions and targeted hardness, soaking time is selected.
During tempering, martensite rejects carbon in the form of finely divided carbide phases.
The end result of tempering is a fine dispersion of carbides in a ferrite matrix, which
often bears little structural similarity to the original as-quenched martensite. It should be
noted that, in many steels, the martensite reaction does not go to completion on
quenching, resulting in varying amounts of retained austenite which does not remain
stable during the tempering process, Figure 19.

Figure 19: As quenched steels with martensite and retained austenite of various
morphologies.

Tempering of Plain Carbon Steels


The as-quenched martensite possesses a complex structure. This occurs in the
first-formed martensite, i.e. the martensite formed near Ms, which has the opportunity of
tempering during the remainder of the quench. This phenomenon, which is referred to
as auto-tempering, is clearly more likely to occur in steels with a high Ms. On reheating
as-quenched martensite, the tempering takes place in four distinct but overlapping
stages:
Up to 250°C, precipitation of a-iron carbide; partial loss of tetragonality in
martensite between 200 and 300°C, decomposition of retained austenite between 200
and 350°C, replacement of iron carbide by cementite; martensite loses tetragonality
above 350°C, cementite coarsens and spheroidizes; recrystallization of ferrite.
It is useful to define a fourth stage of tempering in which the cementite particles
undergo a coarsening process and essentially lose their crystallographic morphology,
becoming spheroidized. The coarsening commences between 300 and 400°C, while
spheroidization takes place increasingly up to 700°C. At the higher end of this range of
temperature the martensite lath boundaries are replaced by more equiaxed ferrite grain
boundaries by a process which is best described as recrystallization. The final result is
an equiaxed array of ferrite grains with coarse spheroidized particles of Fe 3C partly, but
not exclusively, in the grain boundaries.
The spheroidization of the Fe3C rods is encouraged by the resulting decrease in
surface energy. The particles, which preferentially grow and spheroidize are located
mainly at interlath boundaries and prior austenite boundaries, although some particles
remain in the matrix. The boundary sites are preferred because of the greater ease of
diffusion in these regions. The original martensite lath boundaries remain stable up to
about 600°C, but in the range 350-600°C, there is considerable rearrangement of the
dislocations within the laths and at those lath boundaries, which are essentially low
angle boundaries. In the as-quenched state, martensite, in addition to being very hard,
is so brittle that it cannot be used for most applications. Moreover, martensite having
high level of residual stress induced during its transformation during quenching have a
weakening effect. The ductility and toughness of martensite may be enhanced and
these internal stresses relieved by a heat treatment known as tempering. Tempering is
accomplished by heating a martensitic steel up to a temperature below the eutectoid for
a specified time period.
We know martensite is a supersaturated single-phase BCT structure. During
tempering it transforms to the tempered martensite, composed of the stable ferrite and
cementite phases, as indicated on the iron–iron carbide phase diagram. The
microstructure of tempered martensite consists of extremely small and uniformly
dispersed cementite particles embedded within ferrite and martensite matrix depending
on the tempering temperature. As a result, tempering might reduce very little to a
significant decrease in hardness with compatible change in microstructures, Figure 20.

Figure 20: Micrographs showing the effect of tempering: Very low temperature tempering
(left), moderate temperature tempering (middle) and high temperature tempering (right).
Depending on tempering temperature and time of tempering, formed tempered
martensite may be nearly as hard and strong as martensite, but with substantially
enhanced ductility and toughness. The effects of tempering temperature on hardness
andtensileproperties of steel are shown in Figures 21 and 22.

Figure 21: Effects of tempering temperatures on hardness values.

Figure 22: Effects of tempering temperature on tensile properties of steel.


Defects in Hardening
Uneven heating is the cause of most of the defects in hardening. Cracks of a
circular form, from the corners or edges of a tool, indicate uneven heating in hardening.
Cracks of a vertical nature and dark-colored fissures indicate that the steel has been
burned, Figure 23.

Figure 23: Various defects after hardening heat treatment.

You might also like