Download as pptx, pdf, or txt
Download as pptx, pdf, or txt
You are on page 1of 65

Video of steel making:

https://www.youtube.com/watch?v=9l7JqonyoKA
Eutectoid reaction

y↔α+ Fe3C
In order to understand the transformation processes, consider a steel of the
eutectoid composition. 0.8%carbon, being slow cooled along line x-x‘.
• At the upper temperatures, only austenite is present, with the 0.8% carbon being
dissolved in solid solution within the FCC. When the steel cools through 723°C,
several changes occur simultaneously.
• The iron wants to change crystal structure from the FCC austenite to the BCC
ferrite, but the ferrite can only contain 0.02% carbon in solid solution.
• The excess carbon is rejected and forms the carbon-rich intermetallic known as
cementite.
Phases in Fe–Fe3C Phase Diagram

α-ferrite - solid solution of C in BCC Fe Alpha Phase


• Stable form of iron at room temperature.
• The maximum solubility of C is 0.022 wt%
• Transforms to FCC γ-austenite at 912 °C

γ-austenite - solid solution of C in FCC Fe Gamma Phase


• The maximum solubility of C is 2.14 wt %.
• Transforms to BCC δ-ferrite at 1395 °C
• Is not stable below the eutectic temperature
(727 ° C) unless cooled rapidly.
Austenite A face centred cubic (FCC) phase found in all steels. In most carbon and
low alloy steels it is present only above 760oC. In austenitic stainless steels their
chemical composition stabilises austenite to room temperature and even cryogenic
temperatures.
Austenite is non-magnetic.
Austenitic Stainless Steel Stainless Steels that contain a minimum of 18%
Chromium and sufficient Nickel, or a combination of Nickel,
Manganese and Nitrogen, to stabilise the face centred cubic (FCC)
phase austenite down to cryogenic temperatures – This phase is
normally present only above 760oC in most steels. These steels are, except in very
particular circumstances, nonmagnetic and have good ductility but relatively high
work hardening rates. They have excellent corrosion resistance to most
environments, although susceptible to stress corrosion cracking (SCC) above 60oC.
Pitting and crevice corrosion can occur in chloride environments except with
Type/Grade 316 where the Molybdenum content inhibits, but cannot always fully
prevent, these problems.They cannot be hardened by heat treatment - The
δ-ferrite solid solution of C in BCC Fe Delta Phase
• The same structure as α-ferrite
• Stable only at high T, above 1394 °C
• Melts at 1538 °C

Fe3C (iron carbide or cementite)


• This intermetallic compound is metastable, it
remains as a compound indefinitely at room T, but
decomposes (very slowly, within several years)
into α-Fe and C (graphite) at 650 - 700 °C

Fe-C liquid solution


A few comments on Fe–Fe3C system
C is an interstitial impurity in Fe. It forms a solid solution
with α, γ, δ phases of iron
Maximum solubility in BCC (Body Centred Cubic) α-ferrite is limited
(max. 0.022 wt% at 727 °C) - BCC has relatively small interstitial
Positions.
Maximum solubility in FCC (Face Centred Cubic) austenite is 2.14 wt% at
1147°C - FCC has larger interstitial positions.
Mechanical properties: Cementite is very hard and brittle - can
strengthen steels. Mechanical properties also depend on the
microstructure, that is, how ferrite and cementite are mixed.
Magnetic properties: α -ferrite is magnetic below 768 °C,
austenite is non-magnetic.
Classification. Three types of ferrous alloys:
• Iron: less than 0.008 wt % C in α−ferrite at room T
• Steels: 0.008 - 2.14 wt % C (usually < 1 wt % )
α-ferrite + Fe3C at room T
• Cast iron: 2.14 - 6.7 wt % (usually < 4.5 wt %)
Follow Green arrow on next slide

Follow Blue arrow on next slide


Microstructure of eutectoid steel
When alloy of eutectoid composition (0.76 wt % C) is cooled slowly it forms
pearlite, a lamellar or layered structure of two phases: α-ferrite and cementite
(Fe3C)
The layers of alternating phases in pearlite are formed for the same reason as
layered structure of eutectic structures:
redistribution C atoms between ferrite (0.022 wt%) and cementite (6.7 wt%) by
atomic diffusion.
Mechanically, pearlite has properties intermediate to soft, ductile ferrite and hard,
brittle cementite.
Martensite
• Martensite forms when austenite is rapidly cooled
(quenched) to room T.
• It forms nearly instantaneously when the required low temperature is reached.
The austenite-martensite does not involve diffusion → no thermal activation is
needed, this is called an athermal transformation.
• Each atom displaces a small (sub-atomic) distance to transform FCC γ-Fe
(austenite) to martensite which has a Body Centered Tetragonal (BCT) unit cell (like
BCC,
but one unit cell axis is longer than the other two).
• Martensite is metastable - can persist indefinitely at room temperature, but will
transform to equilibrium phases on annealing at an elevated temperature.
• Martensite can coexist with other phases and/or microstructures in Fe-C system
• Since martensite is metastable non-equilibrium phase, it does not appear in
phase Fe-C phase diagram.
Mechanical properties of martensite
Of the various microstructures in steel alloys Martensite is the hardest, strongest
and the most brittle.
The strength of martensite is not related to microstructure.
Rather, it is related to the interstitial C atoms hindering dislocation motion (solid
solution hardening) and to the small number of slip systems.
Tempered Martensite
Martensite is so brittle that it needs to be modified for practical applications. This
is done by heating it to 250-650 o C for some time (tempering) which produces
tempered martensite, an extremely fine-grained and well dispersed cementite
grains in a ferrite matrix.
Tempered martensite is less hard/strong as compared to regular martensite but has
enhanced ductility (ferrite phase is ductile).
Mechanical properties depend upon cementite particle size: fewer, larger particles
means less boundary area and softer, more ductile material.
Particle size increases with higher tempering temperature and/or longer time
(more C diffusion) - therefore softer, more ductile material.
Material Properties
Mechanical Properties The properties of a material that reveal its elastic or plastic
behaviour under an applied load. They govern its suitability for any mechanical
application. The usual properties considered are modulus of elasticity, yield or
0.2%/1.0% proof strength, ultimate tensile strength, elongation and fatigue limit.
Related terms: Tensile Strength

Ductility The ability of a metal or alloy to deform without cracking or failing under
tensile loads. Materials with low ductility exhibit brittle behaviour and fail at more
modest loads. Ductility typically decreases at lower temperatures and most
materials will become brittle below a certain temperature called the Ductile-Brittle
Transition.

Brittleness Defined as the lack of ductility normally associated with hardness.


Malleability Is a property similar to ductility. If a material can be easily beaten or
rolled into plate form it is said to be malleable.

Elasticity If all the strain in a stressed material disappears upon removal of the
stress the material is elastic. (It returns to the unstressed size)

Plasticity If none of the strain in a stressed material disappears upon removal of


the stress the material is plastic. (It does not return to unstressed size)

Hardness A material’s resistance to erosion or wear will indicate the hardness of


the material.

Strength The greater the load which can be carried the stronger the material.

Toughness A material’s ability to sustain variable load conditions without failure is


a measure of a material’s toughness.
Manganese (Mn): Manganese is added to steel to improve hot working properties
and increase strength, toughness and hardenability. Manganese, like nickel, is
an Austenite forming element and has been used as a substitute for nickel in the
AISI200 Series of Austenitic Stainless Steels, e.g. AISI 202 as a substitute for AISI
304.
Chromium (Cr): Chromium is added to steel to increase resistance to oxidation. This
resistance increases as more chromium is added. 'Stainless Steels
have a minimum of 10.5% Chromium (traditionally 11 or 12%). This gives a very
marked degree of general corrosion resistance when compared to steels with a
lower percentage of Chromium. The corrosion resistance is due to the formation of
a self-repairing passive layer of Chromium Oxide on the surface of the
stainless steel.
Nickel (Ni): Nickel is added in large amounts, over about 8%, to high Chromium
stainless steels to form the most important class of corrosion and heat resisting
steels. These are the Austenitic stainless steels, typified by 18-8 (304/1.4301),
where the tendency of Nickel to form Austenite is responsible for
a great toughness (impact strength) and high strength at both high and low
temperatures. Nickel also greatly improves resistance to oxidation and corrosion.
Molybdenum (Mo): Molybdenum, when added to chromium-nickel austenitic
steels, improves resistance to pitting and crevice corrosion especially in chlorides
and sulphur containing environments.
Nitrogen (N): Nitrogen has the effect of increasing the Austenite stability of
stainless steels and is, as in the case of Nickel, an Austenite forming element. Yield
strength is greatly improved when nitrogen is added to stainless steels as is
resistance to pitting corrrosion.
Copper (Cu): Copper is normally present in stainless steel as a residual element.
However, it is added to a few alloys to produce precipitation hardening properties
or to enhance corrosion resistance particularly in sea water environments and
sulphuric acid.
Titanium (Ti): Ttitanium is added for carbide stabilization especially when the
material is to be welded. It combines with carbon to form titanium
carbides, which are quite stable and hard lo dissolve in steel, which tends to
minimise the occurrence of intergranular corrosion. Adding approximately 0.25 /
0.60% titanium causes the carbon to combine with titanium in preference to
chromium, preventing a tieup of corrosion-resisting chromium as inter-granular
carbides and the accompanying loss of corrosion resistance at the grain
boundaries. However, the use of titanium has gradually decreased over recent
years due to the ability of steelmakers to deliver stainless steels with very low
carbon contents that are readily weldable without stabilisation.
Phosphorus (P): Phosphorus is usually added with sulphur, to improve
machinability. The Phosphorus present in Austenitic stainless steels increases
strength. However, it has a detrimental effect on corrosion resistance and increases
the tendency of the material to crack during welding.
Sulphur (S): When added in small amounts Sulphur improves machinability.
However, like Phosphorous it has a detrimental effect on corrosion resistance and
weldability.
Selenium (Se): Selenium was previously used as an addition to improve
machinability.
Niobium / Colombium (Nb): Niobium is added to steel in order to stabilise carbon,
and, as such, performs in the same way as described for Titanium. Niobium also
has the effect of strengthening steels and alloys for high temperature service.
SiIicon (Si): Silicon is used as a deoxidising (killing) agent in the melting of steel and
as a result most steels contain a small percentage of Silicon.
Cobalt (Co): Cobalt becomes highly radioactive when exposed to the intense
radiation of nuclear reactors, and, as a result, any stainless steel that is in nuclear
service will have a Cobalt restriction, usually approximately 0.2% maximum. This
problem is emphasized because there is normally a residual Cobalt content in the
Nickel used in producing Austenitic stainless steels.
Calcium (Ca): Small additions are used to improve machiniability, without the
detrimental effects on other properties caused by Sulphur, Phosphorus and
Selenium.
Stress Strain Curve for Mild Steel
 
If tensile force is applied to a steel bar, it will have some elongation. If the force is
small enough, the ratio of the stress and strain will remain proportional. This can
be seen in the graph as a straight line between zero and point A – also called
the limit of proportionality. If the force is greater, the material will experience
elastic deformation, but the ratio of stress and strain will not be proportional. This
is between points A and B, known as the elastic limit.
Beyond the elastic limit, the mild steel will experience plastic deformation. This
starts the yield point – or the rolling point – which is point B, or the upper yield
point. As seen in the graph, from this point on the correlation between the stress
and strain is no longer on a straight trajectory. It curves from point C (lower yield
point), to D (maximum ultimate stress), ending at E (fracture stress).
Now, we’ll look at each individual measure on the graph above and explain how
each is derived.
Stress: If an applied force causes a change in the dimension of the material, then
the material is in the state of stress. If we divide the applied force (F) by the cross-
sectional area (A), we get the stress.
The symbol of stress is σ (Greek letter sigma). For tensile (+) and compressive (-)
forces. The standard international unit of stress is the pascal (Pa), where 1 Pa = 1
N/m2. The formula to derive the stress number is σ = F/A.
For tensile and compressive forces, the area taken is perpendicular to the applied
force. For sheer force, the area is taken parallel to the applied force. The symbol for
shear stress is tau (τ).
Strain: Strain is the change in the dimension (L-L0) with respect to the original. It is
denoted by the symbol epsilon (ε). The formula is ε = (L-L0) / L0. For a shear force,
strain is expressed by γ (gamma)
Elasticity: Elasticity is the property of the material which enables the material to
return to its original form after the external force is removed.
Plasticity: This is a property that allows the material to remain deformed without
fracture even after the force is removed.
The definitions below are important for understanding the Stress-Strain
interactions as seen in the graph.
 Hooke’s Law: Within the proportional limit (straight line between zero and A),
strain is proportionate to stress.
 Young’s modulus of elasticity: Within the proportional limit, stress = E × strain. E is
a proportionality constant known as the modulus of elasticity or Young’s modulus
of elasticity. Young’s modulusis a measure of the ability of a material to withstand
changes in length when under lengthwise tension or compression. E has the same
unit as the unit of stress because the strain is dimensionless. The formula is E
= σ / ε Pa.
2.1.6 For plates, strip and sections, the test specimens are to be machined to the
dimensions shown in Fig. 2.2.3 or Fig. 2.2.4. Where the capacity of the available
testing machine is insufficient to allow the use of a test specimen of full
thickness, this may be reduced by machining one of the rolled surfaces.
Alternatively, for materials over 40 mm thick, test specimens of circular cross-
section machined to the dimensions shown in Fig. 2.2.1 may be used. The axes of
these test specimens are to be located at approximately one quarter of the
thickness from one of the rolled surfaces.
2.1.11 For aluminium alloy plates and sections of thick-ness, a, less than or equal to
12,5 mm; the dimensions of rectangular cross-sectioned test specimens are to be
as shown in Fig. 2.2.3. The rectangular cross-sectioned test specimen surfaces
should remain as rolled/extruded. Where the thickness, a, is greater than 12,5 mm
the test specimens are to be of round type as shown in Fig. 2.2.2.
3.1.1 Impact tests are to be of the Charpy V-notch type. The test specimens are to
be machined to the dimensions and tolerances given in Table 2.3.1 and are to be
carefully checked for dimensional accuracy.
Engineering materials don’t reach theoretical strength when they are tested in the
laboratory. Therefore, the performance of the material in service is not same as it
is expected from the material, hence, the design of a component frequently
implores the engineer to minimize the possibility of failure. However, the level of
performance of components in service depends on several factors such as inherent
properties of materials, load or stress system, environment and maintenance. The
reason for failure in engineering component can be attributed to design
deficiencies, poor selection of materials, manufacturing defects, exceeding design
limits and overloading, inadequate maintenance etc.
Mechanical Failure
The usual causes of mechanical failure in the component or system are:
• Misuse or abuse
• Assembly errors
• Manufacturing defects
• Improper or inadequate maintenance
• Design errors or design deficiencies
• Improper material or poor selection of materials
• Improper heat treatments
• Unforeseen operating conditions
• Inadequate quality assurance
• Inadequate environmental protection/control
• Casting discontinuities.
The general types of mechanical failure include:
• Failure by fracture due to static overload, the fracture being either brittle or
ductile.
• Buckling in columns due to compressive overloading.
• Yield under static loading which then leads to misalignment or overloading on
other components.
• Failure due to impact loading or thermal shock.
• Failure by fatigue fracture.
• Creep failure due to low strain rate at high temperature.
• Failure due to the combined effects of stress and corrosion.
• Failure due to excessive wear.
Failure Due to Fracture
Fracture is described in various ways depending on the behavior of material under
stress upon the mechanism of fracture or even its appearance. The fracture can be
classified either as ductile or brittle depending upon whether or not plastic
deformation of the material before any catastrophic failure. A brief description of
both types of fracture is given below.
Ductile Fracture
Ductile fracture is characterized by tearing of metal and significant plastic deformation. The
ductile fracture may have a gray, fibrous appearance. Ductile fractures are associated with
overload of the structure or large discontinuities. This type of fracture occurs due to error in
design, incorrect selection of materials, improper manufacturing technique and/or handling.
Figure 2.2 shows the features of ductile fracture. Ductile metals experience observable plastic
deformation prior to fracture. Ductile fracture has dimpled, cup and cone fracture appearance.
The dimples can become elongated by a lateral shearing force, or if the crack is
in the opening (tearing) mode. The fracture modes (dimples, cleavage, or
intergranular fracture) may be seen on the fracture surface and it is possible all
three modes will be present of a given fracture face.

Brittle Fracture
Brittle fracture is characterized by rapid crack propagation with low energy release
and without significant plastic deformation. Brittle metals experience little or no
plastic deformation prior to fracture. The fracture may have a bright granular
appearance. The fractures are generally of the flat type and chevron patterns may
be present. Materials imperfection, sharp corner or notches in the component,
fatigue crack etc. Brittle fracture displays either cleavage (transgranular) or
intergranular fracture. This depends upon whether the grain boundaries are
stronger or weaker than the grains. This type of fracture is associated with non-
metals such as glass, concrete and thermosetting plastics. In metals, brittle fracture
occurs mainly when BCC and HCP (Hexagonal Close Packed) crystals are present.
In polymeric material, initially the crack grows by the growth of the voids along
the midpoint of the trend which then coalesce to produce a crack followed by the
growth of voids ahead of the advancing crack tip. This part of the fracture surface
shows as the rougher region. Prior to the material yielding and necking formation,
the material is quite likely to begin to show a cloudy appearance. This is due to
small voids being produced within the material. Ceramics are brittle materials,
whether glassy or crystalline. Typically fractured ceramic shows around the origin
of the crack a mirror-like region bordered by a misty region containing numerous
micro cracks. In some cases, the mirror-like region may extend over the entire
surface. The difference between ductile fracture and brittle fracture is shown in
Table 2.1.
Ferritic steel, as well as other BCC metals, suffer from the fact that at low enough
temperatures they will break in a brittle fashion. What this means in terms of the
ideas of the tensile test presented above, is that the % elongation to failure is close
to zero.
As the temperature is lowered one finds that over a small temperature range the
BCC metals suddenly begin to fail in this brittle mode. An average temperature of
the small range, called the DBTT (ductile brittle transition temperature), is often
chosen to characterize the temperature where the transition occurs.
The simple tensile test will detect this transition, but, unfortunately, it detects DBTT
values well below those that occur in complex steel parts. The tensile test applies
stress in only one direction. In complex steel parts, the applied stress will act in all
3 possible directions, a situation called a triaxial stress state. The DBTT turns out to
be raised by a triaxial stress state.
A triaxial stress state will develop at the base of a notch when a notched sample is
broken in a tensile machine, and such tests are called notched tensile tests.
Charpy Impact Test. this is used as an
indicator test for deciding steel grades
for class.

However, it is more useful to break the sample with an impact test, where the load
is applied much more rapidly than in a tensile machine, as the combination of the
notch geometry and the high load rate produces values of DBTT close to the
temperature where brittle failure begins to occur in complex steel parts.
Failure Due to Fatigue
Metal fatigue is caused by repeated cycling of the load. It is a progressive localized
damage due to fluctuating stresses and strains on the material. Metal fatigue
cracks initiate and propagate in regions where the strain is most severe. Slide
shows typical S–N curve for the fatigue strength of a metal.
The process of fatigue consists of three stages:
• Initial crack formation
• Progressive crack growth across the part
• Final but sudden fracture of the remaining cross section.
Prevention of Fatigue Failure
The most effective method of improving fatigue performance is improvements in
design. The following design guideline is effective in controlling or preventing
fatigue failure:
Schematic of S–N curve showing increase in fatigue life with decreasing stresses
• Eliminate or reduce stress raisers by streamlining the part or component.
• Avoid sharp surface tears resulting from punching, stamping, shearing, or other
processes.
• Prevent the development of surface discontinuities during processing.
• Reduce or eliminate tensile residual stresses caused by manufacturing.
• Improve the details of fabrication and fastening procedures.
Failure Due to Creep
Creep occurs under certain load at elevated temperature normally above 40 % of
melting temperature of the material. Boilers, gas turbine engines, and ovens are
some of the examples whereby the components experiences creep phenomenon.
An understanding of high temperature materials behaviour over a period of time is
beneficial in evaluating failures of component due to creep. Failures involving
creep are usually easy to identify due to the deformation that occurs. Failures may
appear ductile or brittle manner due to creep. Cracking may be either
transgranular or intergranular, if creep testing is done at a constant temperature
and load, actual components may experience damage or failure at various
temperatures and loading conditions.
In a creep test, a constant load is applied to a tensile specimen maintained at a
constant temperature. Strain is then measured over a period of time. The slope of
the curve, shown in slide is the strain rate of the test during stage II or the creep
rate of the material. Primary creep (known as stage I) is a period of decreasing
creep rate. Primary creep is a period of primarily transient creep.
During this period deformation takes place and the resistance to creep increases
until stage II. Secondary creep (or stage II) is a period of approximate constant
creep rate. Stage II is referred to as steady state creep. Tertiary creep (stage III)
occurs when there is a reduction in cross sectional area due to necking or effective
reduction in area due to internal void formation. Subsequently, increase in creep
rate leading to the creep fracture or stress rupture.
Strain rate (typical creep curve) of material under creep test

You might also like