Sunteți pe pagina 1din 64

Heat Treatment

Lecture 4
Dr. Marwa A. Abbas

Metallurgy and Materials Science dept.

abbas.marwa@outlook.com
Hardening
• Hardening is carried out by quenching a steel, that is cooling
it rapidly from a temperature above the transformation
temperature.

• Steel is quenched in water or brine for the most rapid cooling,


in oil for some alloy steels, and in air for certain higher alloy
steels.

• With this fast cooling rate, the transformation from austenite


to pearlite cannot occur and the new phase obtained by
quenching is called martensite.

• Martensite is a supersaturated metastable phase and have


body centered tetragonal lattice (bct) instead of bcc. After steel
is quenched, it is usually very hard and strong but brittle.
Martensite looks needle-like under microscope due to its fine
lamellar structure.
(a) The unit cell of BCT martensite is related to the FCC austenite unit cell.
(b) As the percentage of carbon increases, more interstitial sites are filled by the carbon atoms
and the tetragonal structure of the martensite becomes more pronounced.
HARDENING Heat Treatment
 The sample is heated above A3 | Acm to cause Austenization. The sample is then quenched at a
cooling rate higher than the critical cooling rate (i.e. to avoid the nose of the CCT diagram).
 The quenching process produces residual strains (thermal, phase transformation).
 The transformation to Martensite is usually not complete and the sample will have some
retained
Heat A3 | Acm → Austenization → Quench (higher than critical cooling rate)
Austenite.
above
 The Martensite produced is hard and brittle and tempering operation usually follows
hardening. This gives a good combination of strength and toughness.

910°C nin
g Acm
Har rde
d e ni
ng Ha
A3

723°C Full Annealing


A1


T

Wt% C
0.8 %
Quenching has been
known for thousands
of years. For example,
a series of such heat
treatments has been
used for making
Damascus steel and
Japanese Samurai
swords.
Martensite microstructure in high carbon steel.
Retained austenite (white) trapped between martensite needles
(black)
Quenching Media
• Depending on how fast steel must be quenched (from IT
diagram), the heat treater will determine type of
quenching required:
o Water (most severe)
o Oil
o Molten Salt
o Gas/ Air (least severe)
o Many phases in between!!!
Ex: add water/polymer to water reduces quench time!
Adding 10% sodium hydroxide or salt will have twice the
cooling rate!
Severity of quench values of some typical quenching conditions
 Note that we have a variety of quenching media at our disposal, with varying degrees of
cooling effect. The severity of quench is indicated by the ‘H’ factor (defined below), with an
ideal quench having a H-value of ∞.
Severity of Quench as indicated by the heat transfer equivalent H
f −1 f → heat transfer factor
Note that apart from the nature of the
H= [m ]
K K → Thermal conductivity quenching medium, the vigorousness of the
shake determines the severity of the quench.
When a hot solid is put into a liquid
Process Variable H medium, gas bubbles form on the surface of
Air No agitation 0.02 the solid (interface with medium). As gas
Oil quench No agitation 0.2 has a poor conductivity the quenching rate is

Increasing severity of quench


" Slight agitation 0.35 reduced. Providing agitation (shaking the
" Good agitation 0.5 solid in the liquid) helps in bringing the
" Vigorous agitation 0.7 liquid medium in direct contact with the
Water quench No agitation 1.0 solid; thus improving the heat transfer (and
" Vigorous agitation 1.5 the cooling rate). The H value/index
Brine quench compares the relative ability of various
No agitation 2.0
(saturated Salt water) media (gases and liquids) to cool a hot solid.
" Vigorous agitation 5.0 Ideal quench is a conceptual idea with a heat
Ideal quench ∞ transfer factor of ∞ (⇒ H = ∞).
Through hardening of the sample
 The surface of is affected by the quenching medium and experiences the best possible
cooling rate. The interior of the sample is cooled by conduction through the (hot) sample and
hence experiences a lower cooling rate. This implies that different parts of the same sample
follow different cooling curves on a CCT diagram and give rise to different microstructures.
 This gives to a varying hardness from centre to circumference. Critical diameter (dc) is that
diameter, which can be through hardened (i.e. we obtain 50% Martensite and 50% pearlite at
the centre of the sample).

Schematic showing variation


in cooling rate from surface
to interior leading to
different microstructures
Typical hardness test survey made along a
diameter of a quenched cylinder
Schematic of Jominy End Quench Test

Jominy hardenability test Variation of hardness along a Jominy bar


(schematic for eutectoid steel)
How to increase the hardenability?
 Hardenability should not be confused with the ability to obtain high hardness. A material
with low hardenability may have a higher surface hardness compared to another sample
with higher hardenability.
 A material with a high hardenability can be cooled relatively slowly to produce 50%
martensite (& 50% pearlite). A material with a high hardenability has the ‘nose’ of the CCT
curve ‘far’ to the right (i.e. at higher times). Such a material can be through hardened easily.

 Hardenability of plain carbon steel can


increased by alloying with most elements (it
is to be noted that this is an added advantage
as alloying is usually done to improve other
properties).
 However, alloying gives two separate ‘C-
curves’ for Pearlitic and Bainitic
transformations (e.g. figure to the right).
TTT diagram of low alloy steel (0.42% C, 0.78%
 This implies that the ‘nose’ of the Bainitic Mn, 1.79% Ni, 0.80% Cr, 0.33% Mo)
U.S.S. Carilloy Steels, United States Steel
transformation has to be avoided to get Corporation, Pittsburgh, 1948)

complete Martensite on quenching.


Effect of carbon
The effect of carbon content on the
hardness of martensite in steels.

Increasing carbon reduces the Ms and


Mf temperatures in plain-carbon
steels.
• The structure and properties of steel martensite depend on the
carbon content of the alloy.
• When the carbon content is low, the martensite grows in a ‘‘lath’’
shape, composed of bundles of flat, narrow plates that grow side by
side. This martensite is not very hard.

• At a higher carbon content, plate martensite grows, in which flat,


narrow plates grow individually rather than as bundles. The
hardness is much greater in the higher carbon, plate martensite
structure, partly due to the greater distortion, or large c/a ratio, of
the crystal structure.

(a) Lath martensite in low-carbon steel. (b) Plate martensite in high carbon Steel.
Retained Austenite
• There is a large volume expansion when martensite forms from
austenite. As the martensite plates form during quenching, they
surround and isolate small pools of austenite, which deform to
accommodate the lower density martensite.

• However, for the remaining pools of austenite to transform, the


surrounding martensite must deform. Because martensite resists the
transformation, either the existing martensite cracks or the austenite
remains trapped in the structure as retained austenite.

• Retained austenite can be a serious problem. Martensite softens and


becomes more ductile during tempering. After tempering, the
retained austenite cools below the Ms and Mf temperatures and
transforms to martensite, since the surrounding tempered martensite
can deform. But now the steel contains more of the hard, brittle
martensite! A second tempering step may be needed to eliminate the
martensite formed from the retained austenite. Retained austenite is
also more of a problem for high-carbon steels. why?
Formation of quench cracks caused by residual stresses produced during
quenching. The figure illustrates the development of stresses as the
austenite transforms to martensite during cooling.
Example
Unusual combinations of properties can be obtained by
producing a steel whose microstructure contains 50% ferrite
and 50% martensite; the martensite provides strength and the
ferrite provides ductility and toughness. Design a heat
treatment to produce a dual phase steel in which the
composition of the martensite is 0.60% C.
Quench Rate
• In practice, undesired micro-constituents may form during the
quenching process. For example, pearlite may form as the steel
cools past the nose of the curve, because the time of the nose is
less than one second in plain-carbon steels.
• The rate at which the steel cools during quenching depends on
several factors. First, the surface always cools faster than the
center of the part. In addition, as the size of the part increases,
the cooling rate at any location is slower. Finally, the cooling
rate depends on the temperature and heat transfer
characteristics of the quenching medium
• Continuous Cooling Transformation Diagrams We can develop
a continuous cooling
• transformation (CCT) diagram by determining the
microstructures produced in the steel
• at various rates of cooling. The CCT curve for a 1080 steel is
shown in Figure 13-14.
The CCT diagram (solid lines) for a 1080 steel compared with the
TTT diagram (dashed lines).
• The CCT diagram differs from the TTT diagram in that longer
times are required for transformations to begin and no bainite
region is observed.

• If we cool a 1080 steel at 5C/s, the CCT diagram tells us that


we obtain coarse pearlite; we have annealed the steel.
• Cooling at 35C/s gives fine pearlite and is a normalizing heat
treatment.
• Cooling at 100C/s permits pearlite to start forming, but the
reaction is incomplete and the remaining austenite changes to
martensite.
• We obtain 100% martensite and thus are able to perform a
quench and temper heat treatment, only if we cool faster than
140C/s.
• Other steels, such as the low-carbon steel have more
complicated CCT diagrams.
In various handbooks, you can find a compilation of TTT and CCT diagrams for
different grades of steels.
• The addition of alloys to steel decreases the cooling rate
required to produce hardness. A decrease in the cooling
rate is an advantage, since it lessens the danger of cracking
and warping.

• Pure iron, wrought iron, and extremely low-carbon steels


have very little hardening properties and are difficult to
harden by heat treatment. Cast iron has limited capabilities
for hardening. When you cool cast iron rapidly, it forms
white iron, which is hard and brittle. And when you cool it
slowly, it forms grey iron, which is soft but brittle under
impact.
• In plain carbon steel, the maximum hardness obtained by
heat treatment depends almost entirely on the carbon content
of the steel. As the carbon content increases, the hardening
ability of the steel increases; however, this capability of
hardening with an increase in carbon content continues only
to a certain point.
• In practice, 0.80% carbon is required for maximum hardness.
When you increase the carbon content beyond 0.80%, there is
no increase in hardness, but there is an increase in wear
resistance. This increase in wear resistance is due to the
formation of hard cementite.
• When you alloy steel to increase its hardness, the alloys make
the carbon more effective in increasing hardness and
strength. Because of this, the carbon content required to
produce maximum hardness is lower than it is for plain
carbon steels. Usually, alloy steels are superior to carbon
steels.
Heat Treatment

Lecture 5
Dr. Marwa A. Abbas

Metallurgy and Materials Science dept.

abbas.marwa@outlook.com
Tempering of Steel
• Martensite is not an equilibrium
phase. This is why it does not
appear on the Fe-Fe3C phase
diagram.

• When martensite is heated below


the eutectoid temperature, the
thermodynamically stable α and
Fe3C phases precipitate. This
process is called tempering.

• The decomposition of martensite


causes the strength and hardness Effect of tempering
temperature on the properties
of the steel to decrease while the of an eutectoid steel.
ductility and impact properties
are improved.
The effect of tempering temperature on the mechanical
properties of a 1050 steel.
Tempered martensite in steel
• At low tempering temperatures, the martensite may form two
transition phases: a lower carbon martensite and a very fine
non-equilibrium ε-carbide, or Fe2.4C. The steel is still strong,
brittle, and perhaps even harder than before tempering.

• At higher temperatures, the stable α and Fe3C form and the


steel becomes softer and more ductile.

• If the steel is tempered just below the eutectoid temperature,


the Fe3C becomes very coarse and the dispersion-
strengthening effect is greatly reduced.

• By selecting the appropriate tempering temperature, a wide


range of properties can be obtained. The product of the
tempering process is a microconstituent called tempered
martensite
Tempering
α ' ( BCT ) α ( BCC )
Fe3C
→
Temper
+
Martensite Ferrite Cementite

 A sample with martensitic microstructure is hard but brittle. Hence


after quenching the sample (or component) is tempered. Maternsite
being a metastable phase decomposes to ferrite and cementite on
heating (providing thermal activation).
 Tempering is carried out just below the eutectoid temperature (heat
→ wait→ slow cool).
 In reality the microstructural changes which take place during
tempering are very complex.
 The time temperature cycle for tempering is chosen so as to optimize
strength and toughness. E.g. tool steel has a as quenched hardness of
Rc65, which is tempered to get a hardness of Rc45-55.
Example

A rotating shaft that delivers power from an electric


motor is made from a 1050 steel. Its yield strength
should be at least 145,000 psi, yet it should also have
at least 15% elongation in order to provide
toughness. Design a heat treatment to produce this
part.
Residual Stresses
• Residual Stresses and Cracking Residual stresses are
produced because of the volume change or because of cold
working. stress-relief anneal can remove or minimize these
residual stresses due to cold working.

• Stresses are also induced because of thermal expansion and


contraction.

• During quenching, the surface of the quenched steel cools


rapidly and transforms to martensite. When the austenite in
the center later transforms, the hard surface is placed in
tension, while the center is compressed. If the residual
stresses exceed the yield strength, quench cracks form at
the surface.
Quench cracks

Formation of quench cracks caused by residual stresses produced during


quenching.
Marquenching/Martempering

• Cooling to just above the


Ms and holding until the
temperature equalizes in
the steel, then
subsequent quenching
permits all of the steel to
transform to martensite
at about the same time.
This heat treatment is
called marquenching or
martempering.

The marquenching heat treatment, designed to


reduce residual stresses and quench cracking.
Austempering
The isothermal transformation heat treatment used to produce
bainite, called austempering, simply involves austenitizing the
steel, quenching to some temperature below the nose of the
TTT curve, and holding at that temperature until all of the
austenite transforms to bainite.

The austempering and isothermal anneal heat treatments in a 1080 steel.


MARTEMPERING & AUSTEMPERING
 These processes have been developed to avoid residual stresses generated during quenching.
 In both these processes Austenized steel is quenched above Ms (say to a temperature T1) for
homogenization of temperature across the sample.
 In Martempering the steel is then quenched and the entire sample transforms simultaneously
to martensite. This is followed by tempering.
 In Austempering instead of quenching the sample, it is held at T1 for it to transform to
bainite. 800
Eutectoid temperature
723 Austenite

Pearlite
600
α + Fe3C
500 Pearlite + Bainite
T →

400 Bainite

T1
300
Ms
200 Austempering
Mf
100
Martempering Martensite

0.1 1 10 102 103 104 105


t (s) →
Interrupting the Isothermal Transformation

• Complicated microstructures are produced by interrupting


the isothermal heat treatment.

• For example, we could austenitize the 1050 steel at 800C,


quench to 650C and hold for 10 s (permitting some ferrite
and pearlite to form), then quench to 350C and hold for 1 h
(3600 s). Whatever unstable austenite remained before
quenching to 350C transforms to bainite. The final structure
is ferrite, pearlite, and bainite.

• We could complicate the treatment further by interrupting


the treatment at 350C after 1 min (60 s) and quenching. Any
austenite remaining after 1 min at 350C forms martensite.
The final structure now contains ferrite, pearlite, bainite,
and martensite.
Producing complicated structures by interrupting the isothermal heat
treatment of a 1050 steel.
Heat Treatment

Lecture 6
Dr. Marwa A. Abbas

Metallurgy and Materials Science dept.

abbas.marwa@outlook.com
Effect of carbon content on tempered steels

Effect of carbon content and tempering temperature on room temperature


hardness of three molybdenum steels. Tempering time: 1 h at temperature
Effect of alloying elements
• The main purpose of adding alloying elements to steel is to
increase its hardenability, i.e, its capability to form martensite
upon quenching from above its critical temperature.

• The general effect of alloying elements on tempering is a


retardation of the rate of softening, especially at the higher
tempering temperatures.
• Thus, to reach a given hardness in a given period of time, alloy
steels require higher tempering temperatures than do carbon
steels.

• Non-carbide forming elements such as Ni, Si, Al, P and Mn


remain essentially in solution in the ferrite and have only a
minor effect on tempered hardness. Hardening due to the
presence of these elements occurs mainly through solid-solution
hardening of the ferrite or matrix grain size control.
• The carbide forming elements (Cr, Mo, W, V, Ta, Nb, and Ti)
retard the softening process by the formation of alloy carbides.
The effect of the carbide-forming elements is minimal at low
tempering temperatures where Fe3C forms; however, at higher
temperatures, alloy carbides are formed, and hardness
decreases slowly with tempering temperature.

• The carbide-forming elements retard the coalescence of


cementite during tempering and form numerous small
carbide particles. Under certain conditions, such as with
highly alloyed steels, hardness may actually increase. This
effect is known as secondary hardening. As the alloy content
increases, the magnitude of the secondary-hardening effect
increases.
Influence of molybdenum content on the softening of quenched 0.35% C steels
with increasing tempering temperature.
• Synergistic effects of various combinations of alloying
elements can occur: Cr tends to produce secondary hardening
at a lower temperature than does Mo, and the combination of
Cr and Mo produces a rather flat tempering curve, with the
peak hardness occurring at a lower temperature than when
only Mo is present.

Variation of RT hardness with


tempering temperature for H11
steel.
Other Alloying Effects
• In addition to ease of hardening and secondary hardening,
alloying elements produce a number of other effects. The
higher tempering temperatures used for alloy steels permit
greater relaxation of residual stresses and improve the
properties.

• Furthermore, alloy steels requires using a less drastic quench


so that quench cracking is minimized unless the quenching
rate is too severe. The higher hardenability of alloy steels may
also permit the use of lower carbon content to achieve a given
strength level but with improved ductility and toughness.

• Residual Elements (elements not intentionally added to steel)


such as tin, phosphorus, antimony, and arsenic are known to
cause embrittlement.
Temper Embrittlement
• When carbon or low-alloy steels are cooled slowly from
tempering above 575 °C or are tempered for extended times
between 375 and 575 °C, a loss in toughness occurs compared
to that resulting from normal tempering cycles and relatively
fast cooling rates.

• The cause of temper embrittlement is believed to be the


precipitation of compounds containing trace elements such as
tin, arsenic, antimony, and phosphorus, along with chromium
and/or manganese.
• The intergranular nature of the fracture suggests that the
embrittlement occurs at the prior austenite grain boundaries.

• Although manganese and chromium cannot be restricted, a


reduction of the other elements and quenching from above
575°C are the most effective remedies for this type of
embrittlement.
Effect of temper embrittlement on notch toughness.

Variation in Charpy V-notch impact energy with temperature for 5140 steel
hardened and tempered at 620 °C. one series of specimens was quenched from
the tempering temperature, and the other was furnace cooled.
De-embrittlement procedure

Steels that have been embrittled because of temper


embrittlement can be de-embrittled by:
• Heating to about 575 °C and holding a few minutes, then
• Subsequent rapid cooling or quenching.

• The time for de-embrittlement depends on the alloying


elements present and the temperature of reheating.

• De-embrittlement is accompanied by a redistribution of


impurities at the grain boundaries.
Blue Brittleness
• The heating of plain carbon steels or some alloy steels to the
temperature range of 230 to 370 °C may result in increased
tensile and yield strength, as well as decreased ductility and
impact strength. This embrittling phenomenon is caused by
precipitation hardening and is called blue brittleness because
it occurs within the blue heat range.

• If susceptible steels are heated within the 230 to 370 °C (450


to 700 °F) range, they may be embrittled and thus should not
be used in applications in which they are subjected to impact
loads.
Tempered martensite embrittlement
• Martensitic (Quenched) steels can become embrittled if the
wrong temper temperature is used.

• Both intergranular and transgranular fracture modes may


be observed in tempered martensite embrittlement.

• The combination of the segregation of impurities such as


phosphorus to the austenitic grain boundaries during
austenitizing and the formation of cementite at prior
austenitic grain boundaries during tempering are
responsible for the intergranular fracture mode of tempered
martensite embrittlement.

• The transgranular fracture mode is caused by the formation


of cementite between parallel martensitic lathes during
tempering.
Room-temperature
Charpy V-notch impact
energy vs tempering
temperature for 4130,
4140, and 4150 steels
austenitized at 900 °C
and tempered 1 h at
temperatures shown.
Austempering advantages

Austempering of steel offers several potential advantages:

• Increased ductility, toughness, and strength at a given


hardness .

• Reduced distortion, which lessens subsequent machining


time, stock removal, sorting, inspection, and scrap.

• The shortest overall time cycle to through-harden within the


hardness range of 35 to 55 HRC, with resulting savings in
energy and capital investment.
Steels for Austempering

The selection of steel for austempering must be based on


transformation characteristics as indicated in time-
temperature transformation (TTT) diagrams.

Three important considerations are:

• The location of the nose of the TTT curve and the speed of
the quench being utilized

• The time required for complete transformation of austenite


to bainite at the austempering temperature

• The location of the Ms point


1080 carbon steel has limited
suitability for austempering.

Cooling from the austenitizing


temperature to the austempering
bath must be accomplished in about
1 s to avoid the nose of the TTT
curve, and thus prevent
transformation to pearlite during
cooling.

Depending on the temperature,


isothermal transformation in the
bath is completed within a period
ranging from a few minutes to
about 1 h.
Because of the rapid cooling rate
required, austempering of 1080 can
be successfully applied only to thin
sections of about 5 mm (0.2 in.)
maximum.
5140 low-alloy steel is well suited
to austempering, as indicated by
the TTT curve.

Approximately 2 s are allowed to


bypass the nose of the curve, and
transformation to bainite is
completed within 1 to 10 min at
315 to 400 °C.

Parts made of 5140 steel or of


other steels with similar
transformation characteristics are
adaptable to austempering in
larger section sizes than are
feasible for 1080 steel because of
the greater time allowed for
bypassing the nose of the curve.
• Some steels, although they have sufficient carbon or alloy
content to be hardenable, are impractical for austempering
because transformation at the nose of the TTT curve starts in
less than 1 s, thus making it virtually impossible to quench
other than thin sections in molten salt without forming
some pearlite, or they require excessively long periods of
time for transformation.

• The chemical composition of the steel is the major


determinant of the martensite-start (Ms) temperature.
Carbon is the most significant variable affecting the Ms.
• However, carbide-forming elements (such as Mo and V) can
tie up the carbon as alloy carbides and prevent complete
solution of carbon.
Ms = 538 - (361 × %C)
The Ms temperature, in degrees Centigrade, of a -(39 × % Mn)
completely austenitized steel can be approximately -(19 × % Ni)
calculated by means of the following formula: -(39 × % Cr)
Austenitizing temperature has a significant effect on the time
at which transformation begins. As the austenitizing
temperature is increased above normal (for a specific steel), the
nose of the TTT curve shifts to the right because of grain
coarsening.

Section Thickness Limitations


Maximum thickness of section is important in determining
whether or not a part can be successfully austempered. For
1080 steel, a section thickness of about 5 mm (0.2 in.) is the
maximum that can be austempered to a fully bainitic
structure.
Carbon steels of lower carbon content will be restricted to a
proportionately lesser thickness. Lower-carbon
steels containing boron, however, can be successfully
austempered in heavier sections. In some alloy steels, section
thicknesses up to about 25 mm (1 in.) can be austempered to
fully bainitic structures.
The Grossmann hardenability test
• To determine hardenability according to Grossmann's
method, a number of cylindrical steel bars of different
diameters are hardened in a given cooling medium.

• Using metallographic examination, the bar that has 50%


martensite at its centre is singled out and the diameter of this
bar is designated as the critical diameter (D0), in inches.

• The cooling intensites of the different cooling media have


been determined and are called the H-factors.

• Using the H value of the cooling medium into consideration,


the D0-value can be converted to the ideal diameter Di which
is defined as the bar diameter which, when the surface is
cooled at an infinitely rapid rate (H = oo ), will yield a
structure, at the centre, containing 50% martensite.
The value of Di obtained is hence a measure of the hardenability
of the steel and is independent of the cooling medium.

Charts showing the correlation between


critical diameter D0, ideal diameter Di and H-
value. The lower diagram is an enlargement
of the lower, left-hand portion of the upper
diagram
Calculation of Di-values from the chemical
composition
• The hardenability may be calculated from the composition of
low-alloy and medium-alloy steels, taking into account only
the amount of each element in solution at the austenitizing
temperature.
• The austenite grain size must also be taken into consideration.
The smaller the grain size the lower is the hardenability.

• The computation starts from the C content and the grain size.
By means of Figure 1, a 'base' hardenability characteristic for Di
is obtained. For the other alloying elements the curves in
Figure 2 indicate the multiplying factor that corresponds to
each alloy content.
• Figure 1 is applicable to C contents above 0·8%, but only on the
assumption that all carbides are in solution at the austenitizing
temperature.
Figure 2. Multiplying factors for different
alloying elements for harden ability
calculations
Figure 1. The ideal diameter as a
function of the carbon content and
austenite grain size for plain carbon
steels
Examples of hardenability calculations:

1. Steel SS 2225, ASTM grain size 7, has the following composition:

From Figure 1, the base value of Di is 0· 17 in. On multiplying this value with the
appropriate factors we obtain:
Di = 0·17 X 1·2 X 3·3 X 3·4 X 1·6 = 3·7 in

2. Steel S S 2541 (BS 816M40), AS TM grain size 6, the following composition:

Di= 0·22 X 1·2 X 3·3 X 4·0 X l·5 X1·6 = 8·4 in

The Di-values obtained can be converted to D0 by Figure 1. For example, by


quenching in oil with moderate agitation (H = 0·40) the critical diameter of steel S
S 2225 is D0 = 2 in and of steel SS 2541, D0 = 6·4 in.
Values of Di, calculated as above, are only approximate but they are useful as a
means of comparing the hardenability of different grades or heats.
• Moser and Legat proposed a more precise relationship
between grain size and C content has been obtained and
this is presented in Figure 3. The revised multiplying
factors by Moser and Legat are shown in Figure 4.

Figure 3. Relationship between base Figure 4. Multiplying factors for


hardenability, carbon content and grain size as calculating the hardenability as
obtaining in actual practice influenced by Mo, Mn, Cr, Si and Ni

S-ar putea să vă placă și