You are on page 1of 65

Steel

From Wikipedia, the free encyclopedia


"Steel worker" redirects here. For other uses, see Steel (disambiguation) and Steel worker
(disambiguation).


The steel cable of a colliery winding tower
Steels and other ironcarbon alloy
phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E
Steel is an alloy of iron, with carbon, which may contribute up to 2.1% of its weight. Carbon, other
elements, and inclusions within iron act as hardening agents that prevent the movement
of dislocations that naturally exist in the iron atom crystal lattices. Varying the amount of alloying
elements, their form in the steel either as solute elements, or as precipitated phases, retards the
movement of those dislocations that make iron so ductile and so weak, and so it controls qualities
such as the hardness, ductility, and tensile strength of the resulting steel. Steel can be made
stronger than pure iron, but only by trading away ductility, of which iron has an excess.
Although steel had been produced in bloomery furnaces for thousands of years, steel's use
expanded extensively after more efficient production methods were devised in the 17th century
for blister steel and then crucible steel. With the invention of the Bessemer process in the mid-19th
century, a new era of mass-produced steel began. This was followed by Siemens-Martin
process and then Gilchrist-Thomas process that refined the quality of steel. With their introductions,
mild steel replaced wrought iron.
Further refinements in the process, such as basic oxygen steelmaking (BOS), further lowered the
cost of production, while increasing the quality of the metal and largely replaced earlier methods.
Today, steel is one of the most common materials in the world, with more than 1.3 billion tons
produced annually. It is a major component in buildings, infrastructure, tools, ships, automobiles,
machines, appliances, and weapons. Modern steel is generally identified by various grades defined
by assorted standards organizations.
Contents
[hide]
1 Definitions and related materials
2 Material properties
o 2.1 Heat treatment
3 Steel production
4 History of steelmaking
o 4.1 Ancient steel
o 4.2 Wootz steel and Damascus steel
o 4.3 Modern steelmaking
4.3.1 Processes starting from bar iron
4.3.2 Processes starting from pig iron
5 Steel industry
6 Recycling
7 Contemporary steel
8 Uses
o 8.1 Historical
o 8.2 Long steel
o 8.3 Flat carbon steel
o 8.4 Stainless steel
o 8.5 Low-background steel
9 See also
10 References
o 10.1 Bibliography
11 Further reading
12 External links
Definitions and related materials[edit]
The carbon content of steel is between 0.002% and 2.1% by weight. Too little carbon content leaves
(pure) iron quite soft, ductile, and weak. Carbon contents higher than those of steel make an alloy
commonly called pig iron that is brittle and not malleable. Alloy steel is steel to which additional
alloying elements have been intentionally added to modify the characteristics of steel. Common
alloying elements include: manganese, nickel, chromium, molybdenum, boron, titanium, vanadium,
and niobium.
[1]
Additional elements may be present in steel: manganese, phosphorus, sulfur, silicon,
and traces of oxygen, nitrogen, and aluminium.
Alloys with a higher than 2.1% carbon content, depending on other element content and possibly on
processing, are known as cast iron. Cast iron is not malleable even when hot, but it can be formed
by casting as it has a lower melting point than steel and good castability properties.
[1]
Steel is also
distinguishable from wrought iron (now largely obsolete), which may contain a small amount of
carbon but large amounts of slag. Note that the percentages of carbon and other elements quoted
are on a weight basis.
Material properties[edit]


Iron-carbon phase diagram, showing the conditions necessary to form different phases
Iron is found in the Earth's crust only in the form of an ore, usually an iron oxide, such
as magnetite,hematite etc. Iron is extracted from iron ore by removing the oxygen by combining it
with a preferred chemical partner such as carbon that is lost to the atmosphere as carbon dioxide.
This process, known assmelting, was first applied to metals with lower melting points, such as tin,
which melts at approximately 250 C (482 F) and copper, which melts at approximately 1,100 C
(2,010 F). In comparison, cast iron melts at approximately 1,375 C (2,507 F).
[2]
Small quantities of
iron were smelted in ancient times, in the solid state, by heating the ore buried in a charcoal fire and
welding the metal together with a hammer, squeezing out the impurities. With care, the carbon
content could be controlled by moving it around in the fire.
All of these temperatures could be reached with ancient methods that have been used since
the Bronze Age. Since the oxidation rate of iron increases rapidly beyond 800 C (1,470 F), it is
important that smelting take place in a low-oxygen environment. Unlike copper and tin, liquid or solid
iron dissolves carbon quite readily. Smelting results in an alloy (pig iron) that contains too much
carbon to be called steel.
[2]
The excess carbon and other impurities are removed in a subsequent
step.
Other materials are often added to the iron/carbon mixture to produce steel with desired
properties. Nickeland manganese in steel add to its tensile strength and make the austenite form of
the iron-carbon solution more stable, chromium increases hardness and melting temperature,
and vanadium also increases hardness while making it less prone to metal fatigue.
[3]

To inhibit corrosion, at least 11% chromium is added to steel so that a hard oxide forms on the metal
surface; this is known as stainless steel. Tungsten interferes with the formation of cementite,
allowingmartensite to preferentially form at slower quench rates, resulting in high speed steel. On the
other hand, sulfur, nitrogen, and phosphorus make steel more brittle, so these commonly found
elements must be removed from the steel melt during processing.
[3]

The density of steel varies based on the alloying constituents but usually ranges between 7,750 and
8,050 kg/m
3
(484 and 503 lb/cu ft), or 7.75 and 8.05 g/cm
3
(4.48 and 4.65 oz/cu in).
[4]

Even in a narrow range of concentrations of mixtures of carbon and iron that make a steel, a number
of different metallurgical structures, with very different properties can form. Understanding such
properties is essential to making quality steel. At room temperature, the most stable form of iron is
the body-centered cubic (BCC) structure called ferrite or -iron. It is a fairly soft metal that can
dissolve only a small concentration of carbon, no more than 0.021 wt% at 723 C (1,333 F), and
only 0.005% at 0 C (32 F). At 910C pure iron transforms into a face-centered cubic (FCC)
structure, called austenite or -iron. The FCC structure of austenite can dissolve considerably more
carbon, as much as 2.1%
[5]
(38 times that of ferrite) carbon at 1,148 C (2,098 F), which reflects the
upper carbon content of steel, beyond which is cast iron.
[6]

When steels with less than 0.8% carbon, known as a hypoeutectoid steel, are cooled,
the austenitic phase (FCC) of the mixture attempts to revert to the ferrite phase (BCC), resulting in
an excess of carbon. One way for carbon to leave the austenite is for it to precipitate out of solution
as cementite, leaving behind iron that is low enough in carbon to take the form of ferrite, resulting in
a ferrite matrix with cementite inclusions. Cementite is a hard and brittle intermetallic compound with
the chemical formula of Fe
3
C. At theeutectoid, 0.8% carbon, the cooled structure takes the form
of pearlite, named for its resemblance to mother of pearl. It is a lamellar structure of ferrite and
cementite. For steels that have more than 0.8% carbon, the cooled structure takes the form of
pearlite and cementite.
[7]

Perhaps the most important polymorphic form of steel is martensite, a metastable phase that is
significantly stronger than other steel phases. When the steel is in an austenitic phase and
then quenched rapidly, it forms into martensite, as the atoms "freeze" in place when the cell structure
changes from FCC to BCC. Depending on the carbon content, the martensitic phase takes different
forms. Below approximately 0.2% carbon, it takes an ferrite BCC crystal form, but at higher carbon
content it takes a body-centered tetragonal(BCT) structure. There is no thermal activation energy for
the transformation from austenite to martensite. Moreover, there is no compositional change so the
atoms generally retain their same neighbors.
[8]

Martensite has a lower density than does austenite, so that the transformation between them results
in a change of volume. In this case, expansion occurs. Internal stresses from this expansion
generally take the form of compression on the crystals of martensite and tension on the remaining
ferrite, with a fair amount of shear on both constituents. If quenching is done improperly, the internal
stresses can cause a part to shatter as it cools. At the very least, they cause internal work
hardening and other microscopic imperfections. It is common for quench cracks to form when steel
is water quenched, although they may not always be visible.
[9]

Heat treatment[edit]
Main article: Heat treating carbon steel
There are many types of heat treating processes available to steel. The most common
are annealing and quenching and tempering. Annealing is the process of heating the steel to a
sufficiently high temperature to soften it. This process goes through three
phases: recovery, recrystallization, and grain growth. The temperature required to anneal steel
depends on the type of annealing and the constituents of the alloy.
[10]

Quenching and tempering first involves heating the steel to the austenite phase then quenching it
in water or oil. This rapid cooling results in a hard but brittle martensitic structure.
[8]
The steel is then
tempered, which is just a specialized type of annealing. In this application the annealing (tempering)
process transforms some of the martensite into cementite, or spheroidite to reduce internal stresses
and defects, which ultimately results in a more ductile and fracture-resistant steel.
[11]

Steel production[edit]
Main article: Steelmaking
See also: List of countries by steel production


Iron ore pellets for the production of steel
When iron is smelted from its ore by commercial processes, it contains more carbon than is
desirable. To become steel, it must be melted and reprocessed to reduce the carbon to the correct
amount, at which point other elements can be added. This liquid is thencontinuously cast into long
slabs or cast into ingots. Approximately 96% of steel is continuously cast, while only 4% is produced
as ingots.
[12]

The ingots are then heated in a soaking pit and hot rolled into slabs, blooms, or billets. Slabs are hot
or cold rolled into sheet metal or plates. Billets are hot or cold rolled into bars, rods, and wire.
Blooms are hot or cold rolled into structural steel, such as I-beams andrails. In modern steel mills
these processes often occur in one assembly line, with ore coming in and finished steel coming
out.
[13]
Sometimes after a steel's final rolling it is heat treated for strength, however this is relatively
rare.
[14]

History of steelmaking[edit]



Ferrite (iron)
From Wikipedia, the free encyclopedia
Main article: Allotropes of iron
Steels and other ironcarbon alloy phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E


Iron-carbon phase diagram, showing the conditions under which ferrite () is stable.
Ferrite, also known as -ferrite (-Fe) or alpha iron, is a materials science term for pure iron, with
a body-centered cubic B.C.Ccrystal structure. It is this crystalline structure which
gives steel and cast iron their magnetic properties, and is the classic example of
aferromagnetic material.
[1]

It has a strength of 280 N/mm
2[citation needed]
and a hardness of approximately 80 Brinell.
[2]

Mild steel (carbon steel with up to about 0.2 wt% C) consist mostly of ferrite, with increasing
amounts of pearlite (a fine lamellar structure of ferrite and cementite) as the carbon content is
increased. Since bainite (shown as ledeburite on the diagram at the bottom of this page) and pearlite
each have ferrite as a component, any iron-carbon alloy will contain some amount of ferrite if it is
allowed to reach equilibrium at room temperature. The exact amount of ferrite will depend on the
cooling processes the iron-carbon alloy undergoes as it cools from liquid state.
In pure iron, ferrite is stable below 910 C (1,670 F). Above this temperature the face-centred
cubic form of iron, austenite (gamma-iron) is stable. Above 1,390 C (2,530 F), up to the melting
point at 1,539 C (2,802 F), the body-centred cubic crystal structure is again the more stable form
of delta-ferrite (-Fe). Ferrite above the critical temperature A
2
(Curie temperature) of 771 C
(1,044 K; 1,420 F), where it is paramagnetic rather than ferromagnetic, is beta ferrite or beta iron (-
Fe). The term beta iron is seldom used because it is crystallographically identical to, and its phase
field contiguous with, -Fe.
Only a very small amount of carbon can be dissolved in ferrite; the maximum solubility is about 0.02
wt% at 723 C (1,333 F) and 0.005% carbon at 0 C (32 F).
[3]
This is because carbon dissolves in
iron interstitially, with the carbon atoms being about twice the diameter of the interstitial "holes", so
that each carbon atom is surrounded by a strong local strain field. Hence theenthalpy of mixing is
positive (unfavourable), but the contribution of entropy to the free energy of solution stabilises the
structure for low carbon content. 723 C (1,333 F) also is the minimum temperature at which iron-
carbon austenite (0.8 wt% C) is stable; at this temperature there is a eutectoid reaction between
ferrite, austenite and cementite.


Austenite, also known as gamma phase iron (-Fe), is a metallic, non-magnetic allotrope of iron or
a solid solution of iron, with analloying element.
[1]
In plain-carbon steel, austenite exists above the
critical eutectoid temperature of 1,000 K (1,340 F; 730 C); other alloys of steel have different
eutectoid temperatures. It is named after Sir William Chandler Roberts-Austen (18431902).
[2]

Contents
[hide]
1 Allotrope of iron
2 Austenitization
3 Austempering
4 Behavior in plain carbon-steel
5 Behavior in cast iron
6 Stabilization
7 Austenite transformation and Curie point
8 Thermo-optical emission
9 See also
10 References
11 External links
Allotrope of iron[edit]
From 912 to 1,394 C (1,674 to 2,541 F) alpha iron undergoes a phase transition from body-centred
cubic (BCC) to the face-centred cubic (FCC) configuration of gamma iron, also called austenite. This
is similarly soft and ductile but can dissolve considerably more carbon (as much as 2.04% by mass
at 1,146 C (2,095 F)). This gamma form of iron is exhibited by the most commonly used type
ofstainless steel for making hospital and food-service equipment.
Austenitization[edit]
Austenitization means to heat the iron, iron-based metal, or steel to a temperature at which it
changes crystal structure from ferrite to austenite.
[3]
An incomplete initial austenitization can leave
undissolved carbides in the matrix.
[4]

For some irons, iron-based metals, and steels, the presence of carbides may occur during the
austenitization step. The term commonly used for this is two-phase austenitization.
[5]

Austempering[edit]
Main article: Austempering
Austempering is a hardening process that is used on iron-based metals to promote better
mechanical properties. The metal is heated into the austenite region of the iron-cementite phase
diagram and then quenched in a salt bath or other heat extraction medium that is between
temperatures of 300375 C (572707 F). The metal is annealed in this temperature range until the
austenite turns to bainite or ausferrite (bainitic ferrite + high-carbon austenite).
[6]

By changing the temperature for austenitization, the austempering process can yield different and
desired microstructures.
[7]
A higher austenitization temperature can produce a higher carbon content
in austenite, whereas a lower temperature produces a more uniform distribution of austempered
structure.
[7]
The carbon content in austenite as a function of austempering time has been
established.
[8]

Behavior in plain carbon-steel[edit]
As austenite cools, it often transforms into a mixture of ferrite and cementite as the carbon diffuses.
Depending on alloy composition and rate of cooling, pearlite may form. If the rate of cooling is very
swift, the alloy may experience a large lattice distortion known as martensitic transformation in which
it transforms into a BCT-structure instead of into cubic latticed ferrite and cementite. In industry, this
is a very important case, as the carbon is not able to diffuse due to the cooling speed, which results
in the formation of hardmartensite. The rate of cooling determines the relative proportions of
martensite, ferrite, and cementite, and therefore determines the mechanical properties of the
resulting steel, such as hardness and tensile strength. Quenching (to induce martensitic
transformation), followed by tempering will transform some of the brittle martensite into tempered
martensite. If a low-hardenability steel is quenched, a significant amount of austenite will be retained
in the microstructure.
Behavior in cast iron[edit]
Heating white hypereutectic cast iron above 727 C (1,341 F) causes the formation of austenite in
crystals of primary cementite.
[9]
This austenisation of white iron occurs in primary cementite at the
interphase boundary with ferrite.
[9]
When the grains of austenite form in cementite, they occur as
lamellar clusters oriented along the cementite crystal layer surface.
[9]
Austenite is formed by
withdrawal of carbon atoms from cementite into ferrite.
[9]

Stabilization[edit]
The addition of certain alloying elements, such as manganese and nickel, can stabilize the austenitic
structure, facilitating heat-treatment of low-alloy steels. In the extreme case of austenitic stainless
steel, much higher alloy content makes this structure stable even at room temperature. On the other
hand, such elements as silicon, molybdenum, andchromium tend to de-stabilize austenite, raising
the eutectoid temperature.
Austenite is only stable above 910 C (1,670 F) in bulk metal form. However, the use of a face-
centered cubic (fcc) or diamond cubic substrate allows the epitaxial growth of fcc transition
metals.
[10]
The epitaxial growth of austenite on the diamond (100) face is feasible because of the
close lattice match and the symmetry of the diamond (100) face is fcc. More than a monolayer of -
iron can be grown because the critical thickness for the strained multilayer is greater than a
monolayer.
[10]
The determined critical thickness is in close agreement with theoretical prediction.
[10]

Austenite transformation and Curie point[edit]
In many magnetic alloys, the Curie point, the temperature at which magnetic materials cease to
behave magnetically, occurs at nearly the same temperature as the austenite transformation. This
behavior is attributed to the paramagnetic nature of austenite, while both martensite and ferrite are
strongly ferromagnetic.
Thermo-optical emission[edit]
During heat treating, a blacksmith causes phase changes in the iron-carbon system in order to
control the material's mechanical properties, often using the annealing, quenching, and tempering
processes. In this context, the color of light, or "blackbody radiation," emitted by the workpiece is an
approximate gauge of temperature. Temperature is often gauged by watching the color
temperature of the work, with the transition from a deep cherry-red to orange-red (815 C (1,499 F)
to 871 C (1,600 F)) corresponding to the formation of austenite in medium and high-carbon steel.
In the visible spectrum, this glow increases in brightness as temperature increases, and when
cherry-red the glow is near its lowest intensity and may not be visible in ambient light. Therefore,
blacksmiths usually austenize steel in low-light conditions, to help accurately judge the color of the
glow.
Maximum carbon solubility in austenite is 2.03% C at 1,420 K (1,150 C).


Cementite
From Wikipedia, the free encyclopedia
Steels and other ironcarbon alloy
phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E
Cementite, also known as iron carbide, is a chemical compound of iron and carbon, with the
formula Fe
3
C (or Fe
2
C:Fe). By weight, it is 6.67% carbon and 93.3% iron. It has
an orthorhombic crystal structure.
[1]
It is a hard, brittle material,
[1]
normally classified as aceramic in
its pure form, though it is more important in metallurgy.
Contents
[hide]
1 Metallurgy
2 Pure form
3 Other iron carbides
4 References
o 4.1 Bibliography
5 External links
Metallurgy[edit]


Orthorhombic Fe3C. Iron atoms are blue.


The iron-carbon phase diagram
In the ironcarbon system (i.e. plain-carbon steels and cast irons) it is a common constituent
because ferrite can contain at most 0.02wt% of uncombined carbon. Therefore, in carbon steels and
cast irons that are slowly cooled a portion of the elements is in the form of cementite.
[2]
It forms
directly from the melt in the case of white cast iron. In carbon steel, it either forms
from austenite during cooling or from martensite during tempering. An intimate mixture withferrite,
the other product of austenite, forms a lamellar structure called pearlite.
Cementite is thermodynamically unstable, eventually being converted to ferrite and graphite.
Pure form[edit]
Cementite changes from ferromagnetic to paramagnetic at its Curie temperature of approximately
480 K.
[3]

A natural iron carbide (containing minor amounts of nickel and cobalt) occurs in iron meteorites and
is called cohenite after the German mineralogist Emil Cohen, who first described it.
[4]

Other iron carbides[edit]
Two other forms of metastable iron carbide have been identified in tempered steel. Epsilon ()
carbide, hexagonal close-packed Fe
2-3
C, precipitates in plain-carbon steels of carbon content >
0.2%, tempered at 100-200C. Non-stoichiometric -carbide dissolves above ~200C, where Hgg
carbides and cementite begin to form. Hagg carbide,monoclinic Fe
5
C
2
, precipitates in hardened tool
steels tempered at 200-300C.
[5][6]



Graphite
From Wikipedia, the free encyclopedia
For other uses, see Graphite (disambiguation).
Graphite

Graphite specimen
General
Category Native element mineral
Formula C
(repeating unit)
Strunz
classification
01.CB.05a
Crystal symmetry Hexagonal dihexagonal dipyramidal
H-M symbol: (6/m 2/m 2/m)
Space group: P 6
3
/mmc
Unit cell a = 2.461 , c = 6.708 ; Z = 4
Identification
Color Iron-black to steel-gray; deep blue in
transmitted light
Crystal habit Tabular, six-sided foliatedmasses, granular to
compacted masses
Crystal system Hexagonal
Twinning Present
Cleavage Basal perfect on {0001}
Fracture Flaky, otherwise rough when not on cleavage
Tenacity Flexible non-elastic, sectile
Mohs scalehardness 12
Luster Metallic, earthy
Streak Black
Diaphaneity Opaque, transparent only in extremely thin
akes
Density 2.092.23 g/cm
3

Optical properties Uniaxial ()
Pleochroism Strong
Solubility Molten Ni
Other
characteristics
strongly anisotropic, electric conductor,
greasy feel, readily marks
References
[1][2][3]

Graphite /rfat/ is made almost entirely of carbon atoms, and as with diamond, is
a semimetal native element mineral, and anallotrope of carbon. Graphite, meaning "writing stone",
was named by Abraham Gottlob Werner in 1789 from the Ancient Greek (graph), "to
draw/write", for its use in pencils, where it is known as lead (not to be confused with the metallic
element lead). Graphite is the most stable form of carbon under standard conditions. Therefore, it is
used in thermochemistry as the standard statefor defining the heat of formation of carbon
compounds. Graphite may be considered the highest grade of coal, just above anthraciteand
alternatively called meta-anthracite, although it is not normally used as fuel because it is difficult to
ignite.
Contents
[hide]
1 Types or varieties
2 Occurrence
3 Properties
o 3.1 Structure
o 3.2 Other properties
4 History of natural graphite use
o 4.1 Other names
5 Uses of natural graphite
o 5.1 Refractories
o 5.2 Batteries
o 5.3 Steelmaking
o 5.4 Brake linings
o 5.5 Foundry facings and lubricants
o 5.6 Pencils
o 5.7 Other uses
o 5.8 Expanded graphite
o 5.9 Intercalated graphite
6 Uses of synthetic graphite
o 6.1 Invention of a process to produce synthetic graphite
o 6.2 Electrodes
o 6.3 Powder and scrap
o 6.4 Neutron moderator
o 6.5 Other uses
7 Graphite mining, beneficiation, and milling
8 Graphite recycling
9 See also
10 References
11 Further reading
12 External links
Types or varieties[edit]
There are three principal types of natural graphite, each occurring in different types of ore deposit:
Crystalline flake graphite (or flake graphite for short) occurs as isolated, flat, plate-like particles
with hexagonal edges if unbroken and when broken the edges can be irregular or angular;
Amorphous graphite: very fine flake graphite is sometimes called amorphous in the trade;
[4]

Lump graphite (also called vein graphite) occurs in fissure veins or fractures and appears as
massive platy intergrowths of fibrous or acicular crystalline aggregates, and is probably
hydrothermal in origin.
[5]

Highly ordered pyrolytic graphite or highly oriented pyrolytic graphite (HOPG) refers to graphite with
an angular spread between the graphite sheets of less than 1. This highest-quality synthetic form is
used in scientific research, in particular, as a standard for scanner calibration of scanning probe
microscopes.
[6][7]
The name "graphite fiber" is also sometimes used to refer to carbon fiber orcarbon
fiber-reinforced polymer.
Occurrence[edit]


Graphite output in 2005
Graphite occurs in metamorphic rocks as a result of the reduction of sedimentary carbon compounds
during metamorphism. It also occurs in igneous rocks and in meteorites.
[3]
Minerals associated with
graphite include quartz, calcite, micas and tourmaline. In meteorites it occurs with troilite and silicate
minerals.
[3]
Small graphitic crystals in meteoritic iron are called cliftonite.
[5]

According to the United States Geological Survey (USGS), world production of natural graphite in
2012 was 1,100 thousand tonnes (kt), of which the following major exporters are: China (750
kt), India (150 kt), Brazil (75 kt), North Korea (30 kt) and Canada (26 kt). Graphite is not mined in
the United States, but U.S. production of synthetic graphite in 2010 was 134 kt valued at $1.07
billion.
[8]

Properties[edit]
Structure[edit]
Graphite has a layered, planar structure. In each layer, the carbon atoms are arranged in
a honeycomb lattice with separation of 0.142 nm, and the distance between planes is
0.335 nm.
[9]
The two known forms of graphite, alpha (hexagonal) and beta (rhombohedral), have
very similar physical properties, except that the graphene layers stack slightly differently.
[10]
The
hexagonal graphite may be either flat or buckled.
[11]
The alpha form can be converted to the beta
form through mechanical treatment and the beta form reverts to the alpha form when it is heated
above 1300 C.
[12]


Scanning tunneling microscope image of graphite surface atoms


Graphite's unit cell


Animated view of the unit cell in three layers of graphene (note that this is a slightly different unit cell from
the one to the left)


Ball-and-stick model of graphite (two graphene layers)


Side view of layer stacking


Plane view of layer stacking
Rotating graphite stereogram
Other properties[edit]
The acoustic and thermal properties of graphite are highly anisotropic, since phonons propagate
quickly along the tightly-bound planes, but are slower to travel from one plane to another.
Graphite is an electric conductor, consequently, useful in such applications as arc lamp electrodes. It
can conduct electricity due to the vast electron delocalization within the carbon layers (a
phenomenon called aromaticity). These valence electrons are free to move, so are able to conduct
electricity. However, the electricity is primarily conducted within the plane of the layers. The
conductive properties of powdered graphite
[13]
allows its use as pressure sensor in carbon
microphones.
Graphite and graphite powder are valued in industrial applications for their self-lubricating and
dry lubricating properties. There is a common belief that graphite's lubricating properties are solely
due to the loose interlamellar coupling between sheets in the structure.
[14]
However, it has been
shown that in a vacuum environment (such as in technologies for use in space), graphite is a very
poor lubricant.
[citation needed]
This observation led to the hypothesis that the lubrication is due to the
presence of fluids between the layers, such as air and water, which are naturally adsorbed from the
environment. This hypothesis has been refuted by studies showing that air and water are not
absorbed.
[15]
Recent studies suggest that an effect called superlubricity can also account for
graphite's lubricating properties. The use of graphite is limited by its tendency to facilitate pitting
corrosion in some stainless steel,
[16][17]
and to promote galvanic corrosion between dissimilar metals
(due to its electrical conductivity). It is also corrosive to aluminium in the presence of moisture. For
this reason, the US Air Force banned its use as a lubricant in aluminium aircraft,
[18]
and discouraged
its use in aluminium-containing automatic weapons.
[19]
Even graphite pencil marks on aluminium
parts may facilitate corrosion.
[20]
Another high-temperature lubricant, hexagonal boron nitride, has
the same molecular structure as graphite. It is sometimes called white graphite, due to its similar
properties.
When a large number of crystallographic defects bind these planes together, graphite loses its
lubrication properties and becomes what is known as pyrolytic graphite. It is also highly anisotropic,
and diamagnetic, thus it will float in mid-air above a strong magnet. If it is made in a fluidized bed at
10001300 C then it is isotropic turbostratic, and is used in blood contacting devices like
mechanical heart valves and is called (pyrolytic carbon), and is not diamagnetic. Pyrolytic graphite,
and pyrolytic carbon are often confused but are very different materials.
[citation needed]

Natural and crystalline graphites are not often used in pure form as structural materials, due to their
shear-planes, brittleness and inconsistent mechanical properties



Graphite
From Wikipedia, the free encyclopedia
For other uses, see Graphite (disambiguation).
Graphite

Graphite specimen
General
Category Native element mineral
Formula C
(repeating unit)
Strunz
classification
01.CB.05a
Crystal symmetry Hexagonal dihexagonal dipyramidal
H-M symbol: (6/m 2/m 2/m)
Space group: P 6
3
/mmc
Unit cell a = 2.461 , c = 6.708 ; Z = 4
Identification
Color Iron-black to steel-gray; deep blue in
transmitted light
Crystal habit Tabular, six-sided foliatedmasses, granular to
compacted masses
Crystal system Hexagonal
Twinning Present
Cleavage Basal perfect on {0001}
Fracture Flaky, otherwise rough when not on cleavage
Tenacity Flexible non-elastic, sectile
Mohs scalehardness 12
Luster Metallic, earthy
Streak Black
Diaphaneity Opaque, transparent only in extremely thin
akes
Density 2.092.23 g/cm
3

Optical properties Uniaxial ()
Pleochroism Strong
Solubility Molten Ni
Other
characteristics
strongly anisotropic, electric conductor,
greasy feel, readily marks
References
[1][2][3]

Graphite /rfat/ is made almost entirely of carbon atoms, and as with diamond, is
a semimetal native element mineral, and anallotrope of carbon. Graphite, meaning "writing stone",
was named by Abraham Gottlob Werner in 1789 from the Ancient Greek (graph), "to
draw/write", for its use in pencils, where it is known as lead (not to be confused with the metallic
element lead). Graphite is the most stable form of carbon under standard conditions. Therefore, it is
used in thermochemistry as the standard statefor defining the heat of formation of carbon
compounds. Graphite may be considered the highest grade of coal, just above anthraciteand
alternatively called meta-anthracite, although it is not normally used as fuel because it is difficult to
ignite.
Contents
[hide]
1 Types or varieties
2 Occurrence
3 Properties
o 3.1 Structure
o 3.2 Other properties
4 History of natural graphite use
o 4.1 Other names
5 Uses of natural graphite
o 5.1 Refractories
o 5.2 Batteries
o 5.3 Steelmaking
o 5.4 Brake linings
o 5.5 Foundry facings and lubricants
o 5.6 Pencils
o 5.7 Other uses
o 5.8 Expanded graphite
o 5.9 Intercalated graphite
6 Uses of synthetic graphite
o 6.1 Invention of a process to produce synthetic graphite
o 6.2 Electrodes
o 6.3 Powder and scrap
o 6.4 Neutron moderator
o 6.5 Other uses
7 Graphite mining, beneficiation, and milling
8 Graphite recycling
9 See also
10 References
11 Further reading
12 External links
Types or varieties[edit]
There are three principal types of natural graphite, each occurring in different types of ore deposit:
Crystalline flake graphite (or flake graphite for short) occurs as isolated, flat, plate-like particles
with hexagonal edges if unbroken and when broken the edges can be irregular or angular;
Amorphous graphite: very fine flake graphite is sometimes called amorphous in the trade;
[4]

Lump graphite (also called vein graphite) occurs in fissure veins or fractures and appears as
massive platy intergrowths of fibrous or acicular crystalline aggregates, and is probably
hydrothermal in origin.
[5]

Highly ordered pyrolytic graphite or highly oriented pyrolytic graphite (HOPG) refers to graphite with
an angular spread between the graphite sheets of less than 1. This highest-quality synthetic form is
used in scientific research, in particular, as a standard for scanner calibration of scanning probe
microscopes.
[6][7]
The name "graphite fiber" is also sometimes used to refer to carbon fiber orcarbon
fiber-reinforced polymer.
Occurrence[edit]


Graphite output in 2005
Graphite occurs in metamorphic rocks as a result of the reduction of sedimentary carbon compounds
during metamorphism. It also occurs in igneous rocks and in meteorites.
[3]
Minerals associated with
graphite include quartz, calcite, micas and tourmaline. In meteorites it occurs with troilite and silicate
minerals.
[3]
Small graphitic crystals in meteoritic iron are called cliftonite.
[5]

According to the United States Geological Survey (USGS), world production of natural graphite in
2012 was 1,100 thousand tonnes (kt), of which the following major exporters are: China (750
kt), India (150 kt), Brazil (75 kt), North Korea (30 kt) and Canada (26 kt). Graphite is not mined in
the United States, but U.S. production of synthetic graphite in 2010 was 134 kt valued at $1.07
billion.
[8]

Properties[edit]
Structure[edit]
Graphite has a layered, planar structure. In each layer, the carbon atoms are arranged in
a honeycomb lattice with separation of 0.142 nm, and the distance between planes is
0.335 nm.
[9]
The two known forms of graphite, alpha (hexagonal) and beta (rhombohedral), have
very similar physical properties, except that the graphene layers stack slightly differently.
[10]
The
hexagonal graphite may be either flat or buckled.
[11]
The alpha form can be converted to the beta
form through mechanical treatment and the beta form reverts to the alpha form when it is heated
above 1300 C.
[12]


Scanning tunneling microscope image of graphite surface atoms


Graphite's unit cell


Animated view of the unit cell in three layers of graphene (note that this is a slightly different unit cell from
the one to the left)


Ball-and-stick model of graphite (two graphene layers)


Side view of layer stacking


Plane view of layer stacking
Rotating graphite stereogram
Other properties[edit]
The acoustic and thermal properties of graphite are highly anisotropic, since phonons propagate
quickly along the tightly-bound planes, but are slower to travel from one plane to another.
Graphite is an electric conductor, consequently, useful in such applications as arc lamp electrodes. It
can conduct electricity due to the vast electron delocalization within the carbon layers (a
phenomenon called aromaticity). These valence electrons are free to move, so are able to conduct
electricity. However, the electricity is primarily conducted within the plane of the layers. The
conductive properties of powdered graphite
[13]
allows its use as pressure sensor in carbon
microphones.
Graphite and graphite powder are valued in industrial applications for their self-lubricating and
dry lubricating properties. There is a common belief that graphite's lubricating properties are solely
due to the loose interlamellar coupling between sheets in the structure.
[14]
However, it has been
shown that in a vacuum environment (such as in technologies for use in space), graphite is a very
poor lubricant.
[citation needed]
This observation led to the hypothesis that the lubrication is due to the
presence of fluids between the layers, such as air and water, which are naturally adsorbed from the
environment. This hypothesis has been refuted by studies showing that air and water are not
absorbed.
[15]
Recent studies suggest that an effect called superlubricity can also account for
graphite's lubricating properties. The use of graphite is limited by its tendency to facilitate pitting
corrosion in some stainless steel,
[16][17]
and to promote galvanic corrosion between dissimilar metals
(due to its electrical conductivity). It is also corrosive to aluminium in the presence of moisture. For
this reason, the US Air Force banned its use as a lubricant in aluminium aircraft,
[18]
and discouraged
its use in aluminium-containing automatic weapons.
[19]
Even graphite pencil marks on aluminium
parts may facilitate corrosion.
[20]
Another high-temperature lubricant, hexagonal boron nitride, has
the same molecular structure as graphite. It is sometimes called white graphite, due to its similar
properties.
When a large number of crystallographic defects bind these planes together, graphite loses its
lubrication properties and becomes what is known as pyrolytic graphite. It is also highly anisotropic,
and diamagnetic, thus it will float in mid-air above a strong magnet. If it is made in a fluidized bed at
10001300 C then it is isotropic turbostratic, and is used in blood contacting devices like
mechanical heart valves and is called (pyrolytic carbon), and is not diamagnetic. Pyrolytic graphite,
and pyrolytic carbon are often confused but are very different materials.
[citation needed]

Natural and crystalline graphites are not often used in pure form as structural materials, due to their
shear-planes, brittleness and inconsistent mechanical properties


Pearlite
From Wikipedia, the free encyclopedia
For the amorphous volcanic glass, see perlite.

This article needs additional citations for verification. Please help improve this article by adding
citations to reliable sources. Unsourced material may be challenged and removed. (November 2010)
Steels and other ironcarbon alloy phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E


SEM micrograph of etched pearlite, 2000X.


Atom probe tomography of pearlite. The red dots indicate the positions of carbon atoms. Iron atoms are not shown.
The nanotube is shown for size reference.


Pearlite occurs at the eutectoid of the iron-carbon phase diagram (near the lower left).
Pearlite is a two-phased, lamellar (or layered) structure composed of alternating layers of alpha-
ferrite (88 wt%) and cementite (12 wt%) that occurs in some steels and cast irons. In fact, the
lamellar appearance is misleading since the individual lamellae within a colony are connected in
three dimensions; a single colony is therefore an interpenetrating bicrystal of ferrite and cementite. In
an iron-carbon alloy, during slow cooling pearlite forms by a eutectoid reaction as austenite cools
below 727 C (1,341 F) (the eutectoid temperature). Pearlite is a common microstructure occurring
in many grades of steels.
The eutectoid composition of austenite is approximately 0.76% carbon; steel with less carbon
content will contain a corresponding proportion of relatively pure ferrite crystallites that do not
participate in the eutectoid reaction and cannot transform into pearlite. Likewise steels with higher
carbon contents will form cementite before reaching the eutectoid point. The proportion of ferrite and
cementite forming above the eutectoid point can be calculated from the iron/ironcarbide
equilibrium phase diagram using the lever rule.
Steels with pearlitic (eutectoid composition) or near-pearlitic microstructure (near-eutectoid
composition) can be drawn into thin wires. Such wires, often bundled into ropes, are commercially
used as piano wires, ropes for suspension bridges, and as steel cord for tire reinforcement. High
degrees of wire drawing (logarithimic strain above 3) leads to pearlitic wires with yield strengths of
several Giga Pascals. It makes pearlite one of the strongest structural bulk materials on
earth.
[1]
Some hypereutectoid pearlitic steel wires, when cold wire drawn to true (logarithmic) strains
above 5, can even show a maximal tensile strength above 6 GPa. Although pearlite is used in many
engineering applications, the origin of its extreme strength is not well understood. It has been
recently shown that cold drawing not only strengthens pearlite by refining the lamellae structure, but
also simultaneously causes partial chemical decomposition of cementite and even a structural
transition from crystalline to amorphous cementite. The deformation-induced decomposition and
microstructural change of cementite is closely related to several other phenomena such as a strong
redistribution of carbon and other alloy elements like Si and Mn in both the cementite and the ferrite
phase; a variation of the deformation accommodation at the phase interfaces due to a change in the
carbon concentration gradient at the interfaces; and mechanical alloying.
[2]

Pearlite was first identified by Henry Clifton Sorby and initially named sorbite, however the similarity
of microstructure to nacre and especially the optical effect caused by the scale of the structure made
the alternative name more popular.
Bainite is a similar structure with lamellae much smaller than the wavelength of visible light and thus
lacks this pearlescent appearance. It is prepared by more rapid cooling. Unlike pearlite, whose
formation involves the diffusion of all atoms, bainite grows by a displacive transformation
mechanism.
Eutectoid steel[edit]
Eutectoid steel can in principle be transformed completely into pearlite; hypoeutectoid steels can
also be completely pearlitic if transformed at a temperature below the normal eutectoid.
[3]
Pearlite
can be hard and strong but is not particularly tough. It can be wear resistant because of a strong
lamellar network of ferrite and cementite. Examples of applications include cutting tools, high
strength wires, knives,chisels, and nails.
References[edit]
1. Jump up^


Bainite
From Wikipedia, the free encyclopedia
Steels and other ironcarbon alloy phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E


Bainite in draw quality semi-killed steel
Bainite is an acicular microstructure (not a phase) that forms in steels at temperatures of 250550
C (depending on alloy content).
[1]
First described by E. S. Davenport and Edgar Bain, it is one of the
decomposition products that may form when austenite(the face centered cubic crystal structure
of iron) is cooled past a critical temperature. This critical temperature is 1000K (727 C, 1340 F) in
plain carbon steels. Davenport and Bain originally described the microstructure as being similar in
appearance to tempered martensite.
A fine non-lamellar structure, bainite commonly consists of cementite and dislocation-rich ferrite. The
high concentration of dislocations in the ferrite present in bainite makes this ferrite harder than it
normally would be.
[2]

The temperature range for transformation to bainite (250550 C) is between those
for pearlite and martensite. When formed during continuous cooling, the cooling rate to form bainite
is more rapid than that required to form pearlite, but less rapid than is required to form martensite (in
steels of the same composition). Most alloying elements will lower the temperature required for the
maximum rate of formation of bainite, though carbon is the most effective in doing so.
[1]

The microstructures of martensite and bainite at first seem quite similar. This is a consequence of
the two microstructures sharing many aspects of their transformation mechanisms. However,
morphological differences do exist that require a TEM to see. Under alight microscope, the
microstructure of bainite appears darker than martensite due to its low reflectivity.
Bainite is an intermediate of pearlite and martensite in terms of hardness. For this reason, the
bainitic microstructure becomes useful in that no additional heat treatments are required after initial
cooling to achieve a hardness value between that of pearlitic and martensitic steels.
[3]

Contents
[hide]
1 History
2 Formation
o 2.1 Displacive theory
o 2.2 Diffusive theory
3 Morphology
o 3.1 Upper bainite
o 3.2 Lower bainite
4 Incomplete bainite transformation
5 References
6 External links
History[edit]
In the 1920s Davenport and Bain discovered a new steel microstructure which they provisionally
called martensite-troostite, due to it being intermediate between the already known low-
temperature martensite phase and what was then known as troostite (now fine-pearlite).
[4]
This
microstructure was subsequently named bainite by Bain's colleagues at the United States Steel
Corporation
[5]
although it took some time for the name to be taken up by the scientific community
with books as late as 1947 failing to mention bainite by name.
[4]
Bain and Davenport also noted the
existence of two distinct forms: 'upper-range' bainite which formed at higher temperatures and
'lower-range' bainite which formed near the martensite start temperature (these forms are now
known as upper- and lower-bainite respectively). The early terminology was further confused by the
overlap, in some alloys, of the lower-range of the pearlite reaction and the upper-range of the bainite
with the additional possibility of proeutectoid ferrite.
[4]

Formation[edit]


Illustration of a continuous cooling transformation (cct) diagram for steel
At 900 C a typical low-carbon steel is composed entirely of austenite, the high temperature phase of
iron. Below around 700 C (727 C in eutectic iron) the austenite is thermodynamically unstable and,
under equilibrium conditions, it will undergo a eutectoidreaction and form pearlite an interleaved
mixture of ferrite and cementite (Fe
3
C). In addition to the thermodynamic considerations indicated by
the phase diagram, the phase transformations in steel are heavily influenced by the chemical
kinetics. This leads to the complexity of steel microstructures which are a strongly influenced by the
cooling rate. This can be illustrated by a continuous cooling transformation (CCT) diagram which
plots the time required to form a phase when a sample is cooled at a specific rate thus showing
regions in time-temperature space from which the expected phase fractions can be deduced for a
given thermal cycle.
If the steel is cooled slowly the transformation will agree with the equilibrium predictions and pearlite
will dominate the microstructure with some fraction of proeutectoid ferrite or cementite depending on
the chemical composition. However, the transformation from austenite to pearlite is a time-
dependent reconstructive reaction which requires the large scale movement of the iron and carbon
atoms. While the interstitial carbon diffuses readily even at moderate temperatures the self-diffusion
of iron becomes extremely slow at temperatures below 600 C until, for all practical purposes, it
stops. As a consequence a rapidly cooled steel may reach a temperature where pearlite can no
longer form despite the reaction being incomplete and the remaining austenite being
thermodynamically unstable.
Austenite that is cooled very rapidly can form martensite, without any diffusion of either iron or
carbon, by the shear of the austenite's face-centered crystal structure into a distorted body-centered
tetragonal structure. This non-equilibrium phase can only form at low temperatures, where the
driving force for the reaction is sufficient to overcome the considerable lattice strain imposed by the
transformation. The transformation is essentially time-independent with the phase fraction depending
only the degree of cooling below the critical martensite start temperature.
[6]
Further, it occurs without
the diffusion of either substitutional or interstitial atoms and so martensite inherits the composition of
the parent austenite.
Bainite occupies a region between these two process in a temperature range where iron self-
diffusion is limited but there is insufficient driving force to form martensite. In contrast to pearlite,
where the ferrite and cementite grow cooperatively, bainite forms by the transformation of carbon-
supersaturated ferrite with the subsequent diffusion of carbon and the precipitation of carbides. A
further distinction is often made between so-called lower-bainite, which forms at temperatures closer
to the martensite start temperature, and upper-bainite which forms at higher temperatures. This
distinction arises from the diffusion rates of carbon at the temperature at which the bainite is forming.
If the temperature is high then the carbon will diffuse rapidly away from the newly formed ferrite and
form carbides in the carbon-enriched residual austenite between the ferritic plates leaving them
carbide-free. At low temperatures the carbon will diffuse more sluggishly and may precipitate before
it can leave the bainitic ferrite. There is some controversy over the specifics of bainite's
transformation mechanism; both theories are represented below.
Displacive theory[edit]
One of the theories on the specific formation mechanism for bainite is that it occurs by a shear
transformation, as in martensite. The transformation is said to cause a stress-relieving effect, which
is confirmed by the orientation relationships present in bainitic microstructures.
[2]
There are,
however, similar stress-relief effects seen in transformations that are not considered to be
martensitic in nature, but the term 'similar' does not imply identical. The relief associated with bainite
is an invariantplane strain with a large shear component. The only diffusion that occurs by this
theory is during the formation of the carbide phase (usually cementite) between the ferrite plates.
Diffusive theory[edit]
The diffusive theory of bainite's transformation process is based on short range diffusion at the
transformation front. Here, random and uncoordinated thermally activated atomic jumps control
formation and the interface is then rebuilt by reconstructive diffusion. The mechanism is not able to
explain the shape nor surface relief caused by the bainite transformation.
[2]

Morphology[edit]
Typically bainite manifiests as aggregates, termed sheaves, of ferrite plates (sub-units) separated by
retained austenite, martensite or cementite.
[7]
While the sub-units appear separate when viewed on
a 2-dimensional section they are in fact interconnected in 3-dimensions and usually take on a
lenticular plate or lath morphology. The sheaves themselves are wedge-shaped with the thicker end
associated with the nucleation site.
The thickness of the ferritic plates is found to increase with the transformation temperature.
[8]
Neural
network models have indicated that this is not a direct effect of the temperature per se but rather a
result of the temperature dependence of the driving force for the reaction and the strength of the
austenite surrounding the plates.
[8]
At higher temperatures, and hence lower undercooling, the
reduced thermodynamic driving force causes a decrease in the nucleation rate which allows
individual plates to grow larger before they physically impinge on each other. Further, the growth of
the plates must be accommodated by plastic flow in the surrounding austenite which is difficult if the
austenite is strong and resists the plate's growth.
Upper bainite[edit]
So-called "upper bainite" forms around 400550 C in sheaves. These sheaves contain several laths
of ferrite that are approximately parallel to each other and which exhibit a Kurdjumov-Sachs
relationship with the surrounding austenite, though this relationship degrades as the transformation
temperature is lowered. The ferrite in these sheaves has a carbon concentration below 0.03%,
resulting in carbon-rich austenite around the laths.
[1]

The amount of ferrite that forms between the laths is based on the carbon content of the steel. For a
low carbon steel, typically discontinuous "stringers" or small particles of cementite will be present
between laths. For a higher carbon steel, the stringers become continuous along the length of the
adjacent laths.
[1]

Lower bainite[edit]
Lower bainite forms between 250 and 400 C and takes a more acicular form than upper bainite.
There are not nearly as many low angle boundaries between laths in lower bainite. In lower bainite,
the habit plane in ferrite will also shift from <111> towards <110> as transformation temperature
decreases.
[1]
In lower bainite, cementite nucleates on the interface between ferrite and austenite.
Incomplete bainite transformation[edit]
Early research on bainite found that at a given temperature only a certain volume fraction of the
austenite would transform to bainite with the remainder decomposing to pearlite after an extended
delay. This was the case despite the fact that a complete austenite to pearlite transformation could
be achieved at higher temperatures where the austenite wasmore stable. The fraction of bainite that
could form increased as the temperature decreased. This was ultimately explained by accounting for
the fact that when the bainitic ferrite formed the supersaturated carbon would be expelled to the
surrounding austenite thus thermodynamically stabilising it against further transformation.
[9]




Ledeburite
From Wikipedia, the free encyclopedia
Steels and other ironcarbon alloy phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E


Iron-carbon phase diagram, showing the iron-carbon phase diagram (near the lower left).
In iron and steel metallurgy, ledeburite is a mixture of 4.3% carbon in iron and is a eutectic mixture
of austenite and cementite. Ledeburite is not a type of steel as the carbon level is too high although
it may occur as a separate constituent in some high carbon steels, it is mostly found with cementite
or pearlite in a range of cast irons.
It is named after the metallurgist Karl Heinrich Adolf Ledebur (18371906). He was the
first professor of metallurgy at theBergakademie Freiberg. He discovered ledeburite in 1882.
Ledeburite arises when the carbon content is between 2.06% and 6.67%. The eutectic mixture
of austenite and cementite is 4.3% carbon, Fe
3
C:2Fe, with a melting point of 1147 C.
Ledeburite-II (at ambient temperature) is composed of cementite-I with recrystallized secondary
cementite (which separates fromaustenite as the metal cools) and (with slow cooling) of pearlite. The
pearlite results from the eutectoidal decay of the austenite that comes from the ledeburite-I at
723 C. During more rapid cooling, bainite can develop instead of pearlite, and with very rapid
coolingmartensite can develop.



Tempering (metallurgy)
From Wikipedia, the free encyclopedia


Differentially tempered steel. The various colors produced indicate the temperature to which the steel was heated.
Light-straw indicates 204 C (399 F) and light blue indicates 337 C (639 F).
[1][2]

Tempering is a process of heat treating, which is used to increase the toughness of iron-based
alloys. Tempering is usually performed after hardening, to reduce some of the excess hardness, and
is done by heating the metal to some temperature below the critical temperature for a certain period
of time, then allowed to cool in still air. The exact temperature determines the amount of hardness
removed, and depends on both the specific composition of the alloy and on the desired properties in
the finished product. For instance, very hard tools are often tempered at low temperatures,
while springs are tempered to much higher temperatures. In glass, tempering is performed by
heating the glass and then quickly cooling the surface, increasing the toughness.
Contents
[hide]
1 Introduction
2 History
3 Terminology
4 In carbon steel
o 4.1 Quenched-steel
o 4.2 Normalized steel
o 4.3 Welded steel
o 4.4 Quench and self-temper
o 4.5 Blacksmithing
4.5.1 Tempering colors
4.5.2 Differential tempering
o 4.6 Interrupted quenching
4.6.1 Austempering
4.6.2 Martempering
o 4.7 Physical processes
4.7.1 Embrittlement
5 In alloy steels
6 In cast-iron
o 6.1 White tempering
o 6.2 Black tempering
7 Precipitation hardening alloys
8 See also
9 References
10 Further reading
11 External links
Introduction[edit]


Photomicrograph of martensite, a very hard microstructure formed when steel is quenched. Tempering reduces the
hardness in the martensite by transforming it into various forms of tempered martensite.
Tempering is a heat treatment technique applied to ferrous alloys, such as steel or cast iron, to
achieve greater toughness by decreasing the hardness of the alloy. The reduction in hardness is
usually accompanied by an increase in ductility, thereby decreasing the brittlenessof the metal.
Tempering is usually performed after quenching, which is rapid cooling of the metal to put it in its
hardest state. Tempering is accomplished by controlled heating of the quenched work-piece to a
temperature below its "lower critical temperature". This is also called the lower transformation
temperature or lower arrest (A
1
) temperature; the temperature at which the crystalline phases of the
alloy, calledferrite and cementite, begin combining to form a single-phase solid solution referred to
as austenite. Heating above this temperature is avoided, so as not to destroy the very-hard,
quenched microstructure, called martensite.
[3]

Precise control of time and temperature during the tempering process is critical to achieve the
desired balance of physical properties. Low tempering temperatures may only relieve some of the
internal stresses, decreasing brittleness while maintaining a majority of the hardness. Higher
tempering temperatures tend to produce a greater reduction in the hardness, sacrificing some yield
strength and tensile strength for an increase in elasticity and plasticity. However, in some low alloy
steels, containing other elements like chromium and molybdenum, tempering at low temperatures
may produce an increase in hardness, while at higher temperatures the hardness will decrease.
Many steels with high concentrations of these alloying elements behave like precipitation hardening
alloys, which produce the opposite effects under the conditions found in quenching and tempering,
and are referred to as maraging steels.
[3]

In carbon steels, tempering alters the size and distribution of carbides in the martensite, forming a
microstructure called "tempered martensite". Tempering is also performed onnormalized steels and
cast irons, to increase ductility, machinability, and impact strength.
[3]
Steel is usually tempered
evenly, called "through tempering," producing a nearly uniform hardness, but it is sometimes heated
unevenly, referred to as "differential tempering," producing a variation in hardness.
[4]

History[edit]
Tempering is an ancient heat-treating technique. The oldest known example of tempered martensite
is a pick axe which was found in Galilee, dating from around 1200 to 1100 BC.
[5]
The process was
used throughout the ancient world, from Asia to Europe and Africa. Many different methods and
cooling baths for quenching have been attempted during ancient times, from quenching in urine,
blood, or metals like mercury or lead, but the process of tempering has remained relatively
unchanged over the ages. Tempering was often confused with quenching and, often, the term was
used to describe both techniques. In 1889, Sir William Chandler Roberts-Austen wrote, "There is still
so much confusion between the words "temper," "tempering," and "hardening," in the writings of
even eminent authorities, that it is well to keep these old definitions carefully in mind. I shall employ
the word tempering in the same sense as softening."
[6]

Terminology[edit]
In metallurgy, one may encounter many terms that have very specific meanings within the field, but
may seem rather vague when viewed from outside. Terms such as "hardness," "impact resistance,"
"toughness," and "strength" can carry many different connotations, making it sometimes difficult to
discern the specific meaning. Some of the terms encountered, and their specific definitions are:
Strength: This is resistance to permanent deformation and breaking. Strength, in metallurgy, is
still a rather vague term, so is usually divided into yield strength (strength beyond which
deformation becomes permanent), tensile strength (the ultimate breaking strength), and shear
strength (resistance to transverse, or cutting forces).
Toughness: Resistance to fracture, as measured by the Charpy test. Toughness often increases
as strength decreases.
Hardness: Hardness is often used to describe strength or rigidity but, in metallurgy, the term is
usually used to describe resistance to scratching or abrasion.
Brittleness: Brittleness describes a material's tendency to break before bending or deforming
either elastically or plastically. Brittleness increases with decreased toughness, but is greatly
affected by internal stresses as well.
Plasticity: The ability to mold, bend or deform in a manner that does not spontaneously return to
its original shape. This is proportional to the ductility or malleability of the substance.
Elasticity: Also called flexibility, this is the ability to deform, bend, compress, or stretch and return
to the original shape once the external stress is removed. Elasticity is related to the Young's
modulus of the material.
Impact resistance: Usually synonymous with high-strength toughness, it is the ability resist
shock-loading with minimal deformation.
Wear resistance: Usually synonymous with hardness, this is resistance
to erosion, ablation, spalling, or galling.
Structural integrity: The ability to withstand a maximum-rated load while resisting fracture,
resisting fatigue, and producing a minimal amount of flexing or deflection, to provide a
maximum service life.
In carbon steel[edit]
Very few metals react to heat treatment in the same manner, or to the same extent, that carbon
steel does, and carbon steel heat treating behavior can vary radically depending on alloying
elements. Steel can be softened to a very malleable state through annealing, or it can be hardened
to a state nearly as rigid and brittle as glass by quenching. However, in its hardened state, steel is
usually far too brittle, lacking the structural integrity to be useful for most applications. Tempering is a
method used to decrease the hardness, thereby increasing the ductility of the quenched steel, to
impart some springiness and malleability to the metal. This allows the metal to bend before breaking.
Depending on how much temper is imparted to the steel, it may bend elastically (the steel returns to
its original shape once the load is removed), or it may bend plastically (the steel does not return to
its original shape, resulting in permanent deformation), before fracturing. Tempering is used to
precisely balance the mechanical properties of the metal, such as shear strength, yield
strength, hardness, ductility and tensile strength, to achieve any number of a combination of
properties, making the steel useful for a wide variety of applications. Tools such as hammers and
wrenches require good resistance to abrasion, impact resistance, and resistance to deformation.
Springs do not require as much rigidity, but must deform elastically before breaking. Automotive
parts tend to be a little less rigid, but need to deform plastically before breaking.
Except in rare cases where maximum rigidity and hardness are needed, such as the untempered
steel used for files, quenched steel is almost always tempered to some degree. However, steel is
sometimes annealed through a process called normalizing, leaving the steel only partially softened.
Tempering is sometimes used on normalized steels to further soften it, increasing the malleability
and machinability for easier metalworking. Tempering may also be used on welded steel, to relieve
some of the stresses and excess hardness created in the heat affected zone around the weld.
[3]

Quenched-steel[edit]
Tempering is most often performed on steel that has been heated above its upper critical (A
3
)
temperature and then quickly cooled, in a process called quenching, using methods such as
immersing the red-hot steel in water, oil, or forced-air. The quenched-steel, being placed in, or very
near, its hardest possible state, is then tempered to incrementally decrease the hardness to a point
more suitable for the desired application. The hardness of the quenched-steel depends on both
cooling speed and on the composition of the alloy. Steel with a high carbon-content will reach a
much harder state than steel with a low carbon-content. Likewise, tempering high-carbon steel to a
certain temperature will produce steel that is considerably harder than low-carbon steel that is
tempered at the same temperature. The amount of time held at the tempering temperature also has
an effect. Tempering at a slightly elevated temperature for a shorter time may produce the same
effect as tempering at a lower temperature for a longer time. Tempering times vary, depending on
the carbon content, size, and desired application of the steel, but typically range from a few minutes
to a few hours.
Tempering quenched-steel at very low temperatures, between 66 and 148 C (151 and 298 F), will
usually not have much effect other than a slight relief of some of the internal stresses. Tempering at
higher temperatures, from 148 to 205 C (298 to 401 F), will produce a slight reduction in hardness,
but will primarily relieve much of the internal stresses. Tempering in the range of 260 and 340 C
(500 and 644 F) causes a decrease in ductility and an increase in brittleness, and is referred to as
the "tempered martensite embrittlement" (TME) range. Except in the case of blacksmithing, this
range is usually avoided. Steel requiring more strength than toughness, such as tools, are usually
not tempered above 205 C (401 F). Instead, a variation in hardness is usually produced by varying
only the tempering time. When increased toughness is desired at the expense of strength, higher
tempering temperatures, from 370 to 540 C (698 to 1,004 F), are used. Tempering at even higher
temperatures, between 540 and 600 C (1,004 and 1,112 F), will produce excellent toughness, but
at a serious reduction in the strength and hardness. At 600 C (1,112 F), the steel may experience
another stage of embrittlement, called "temper embrittlement" (TE), which occurs if the steel is held
within the TE temperature range for too long. When heating above this temperature, the steel will
usually not be held for any amount of time, and quickly cooled to avoid temper embrittlement.
[3]

Normalized steel[edit]
Steel that has been heated above its upper critical temperature and then cooled in standing air is
called normalized steel. Normalized steel consists of pearlite, bainite and sometimes martensite
grains, mixed together within the microstructure. This produces steel that is much stronger than full-
annealed steel, and much tougher than tempered quenched-steel. However, added toughness is
sometimes needed at a reduction in strength. Tempering provides a way to carefully decrease the
hardness of the steel, thereby increasing the toughness to a more desirable point. Cast-steel is often
normalized rather than annealed, to decrease the amount of distortion that can occur. Tempering
can further decrease the hardness, increasing the ductility to a point more like annealed
steel.
[7]
Tempering is often used on carbon steels, producing much the same results. The process,
called "normalize and temper", is used frequently on steels such as 1045 carbon steel, or most other
steels containing 0.35 to 0.55% carbon. These steels are usually tempered after normalizing, to
increase the toughness and relieve internal stresses. This can make the metal more suitable for its
intended use and easier to machine.
[8]

Welded steel[edit]
Steel that has been arc welded, gas welded, or welded in any other manner besides forge welded, is
affected in a localized area by the heat from the welding process. This localized area, called
the heat-affected zone (HAZ), consists of steel that varies considerably in hardness, from normalized
steel to steel nearly as hard as quenched steel near the edge of this heat-affected zone. The uneven
heating and cooling also creates internal stresses in the metal, both within and surrounding the weld.
Tempering is sometimes used in place of stress relieving (even heating and cooling of the entire
object to just below the A
1
temperature) to both reduce the internal stresses and to decrease the
brittleness around the weld. Localized tempering is often used on welds when the construction is too
large, intricate, or otherwise too inconvenient to heat the entire object evenly. Tempering
temperatures for this purpose are generally around 205 C (401 F) and 343 C (649 F).
[9]

Quench and self-temper[edit]
Modern reinforcing bar of 500 MPa strength can be made from expensive microalloyed steel or by a
quench and self-temper (QST) process. After the bar exits the final rolling pass, where the final
shape of the bar is applied, the bar is then sprayed with water which quenches the outer surface of
the bar. The bar speed and the amount of water are carefully controlled in order to leave the core of
the bar unquenched. The hot core then tempers the already quenched outer part, leaving a bar with
high strength but with a certain degree of ductility too.
Blacksmithing[edit]
Main article: Blacksmith
Tempering was originally a process used and developed by blacksmiths (forgers of iron). The
process was most likely developed by the Hittites of Anatolia (modern-day Turkey), in the twelfth or
eleventh century BC. Without knowledge of metallurgy, tempering was originally devised through a
trial-and-error method.
Because few methods of precisely measuring temperature existed until modern times, temperature
was usually judged by watching the tempering colors of the metal. Tempering often consisted of
heating above a charcoal or coal forge, or by fire, so holding the work at exactly the right
temperature for the correct amount of time was usually not possible. Tempering was usually
performed by slowly, evenly overheating the metal, as judged by the color, and then immediately
cooling, either in open air or by immersing in water. This produced much the same effect as heating
at the proper temperature for the right amount of time, and avoided embrittlement by tempering
within a short time period. However, although tempering-color guides exist, this method of tempering
usually requires a good amount of practice to perfect, because the final outcome depends on many
factors, including the composition of the steel, the speed at which it was heated, the type of heat
source (oxidizing or carburizing), the cooling rate, oil films or impurities on the surface, and many
other circumstances which vary from smith to smith or even from job to job. The thickness of the
steel also plays a role. With thicker items, it becomes easier to heat only the surface to the right
temperature, before the heat can penetrate through. However, very thick items may not be able to
harden all the way through during quenching.
[10]

Tempering colors[edit]


Pieces of through-tempered steel flatbar. The first one, on the left, is normalized steel. The second is quenched,
untempered martensite. The remaining pieces have been tempered in an oven to their corresponding temperature,
for an hour each. "Tempering standards" like these are sometimes used by blacksmiths for comparison, ensuring that
the work is tempered to the proper color.
If steel has been freshly ground, sanded, or polished, it will form an oxide layer on its surface when
heated. As the temperature of the steel is increased, the thickness of the iron oxide will also
increase. Although iron oxide is not normally transparent, such thin layers do allow light to pass
through, reflecting off both the upper and lower surfaces of the layer. This causes a phenomenon
called thin-film interference, which produces colors on the surface. As the thickness of this layer
increases with temperature, it causes the colors to change from a very light yellow, to brown, then
purple, then blue. These colors appear at very precise temperatures, and provide the blacksmith
with a very accurate gauge for measuring the temperature. The various colors, their corresponding
temperatures, and some of their uses are:
Faint-yellow 176 C (349 F) engravers, razors, scrapers
Light-straw 205 C (401 F) rock drills, reamers, metal-cutting saws
Dark-straw 226 C (439 F) scribers, planer blades
Brown 260 C (500 F) taps, dies, drill bits, hammers, cold chisels
Purple 282 C (540 F) surgical tools, punches, stone carving tools
Dark blue 310 C (590 F) screwdrivers, wrenches
Light blue 337 C (639 F) springs, wood-cutting saws
Grey-blue 371 C (700 F) and higher structural steel
Beyond the grey-blue color, the iron oxide loses its transparency, and the temperature can no longer
be judged in this way. The layer will also increase in thickness as time passes, which is another
reason overheating and immediate cooling is used. Steel in a tempering oven, held at 205 C
(401 F) for a long time, will begin to turn brown, purple or blue, even though the temperature did not
exceed that needed to produce a light-straw color. Oxidizing or carburizing heat sources may also
affect the final result. The iron oxide layer, unlike rust, also protects the steel from corrosion
through passivation.
[11]

Differential tempering[edit]
Main article: Differential tempering


A differentially tempered sword. The center is tempered to a springy hardness while the edges are tempered to a very
high hardness.
Differential tempering is a method of providing different amounts of temper to different parts of the
steel. The method was often used in bladesmithing, for making knives and swords, to provide a very
hard edge while softening the spine or center of the blade. This increased the toughness while
maintaining a very hard, sharp, impact-resistant edge, helping to prevent breakage. This technique
was more often found in Europe, as opposed to the differential hardening techniques more common
in Asia, such as in Japanese swordsmithing.
Differential tempering consists of applying heat to only a portion of the blade, usually the spine, or
the center of double-edged blades. For single-edged blades, the heat, often in the form of a flame or
a red-hot bar, is applied to the spine of the blade only. The blade is then carefully watched as the
tempering colors form and slowly creep toward the edge. The heat is then removed before the light-
straw color reaches the edge. The colors will continue to move toward the edge for a short time after
the heat is removed, so the smith typically removes the heat a little early, so that the pale-yellow just
reaches the edge, and travels no farther. A similar method is used for double-edged blades, but the
heat source is applied to the center of the blade, allowing the colors to creep out toward each
edge.
[12]

Interrupted quenching[edit]
Interrupted quenching methods are often referred to as tempering, although the processes are very
different from traditional tempering. These methods consist of quenching to a specific temperature
that is above the martensite start (M
s
) temperature, and then holding at that temperature for
extended amounts of time. Depending on the temperature and the amount of time, this allows either
pure bainite to form, or holds-off forming the martensite until much of the internal stresses relax.
These methods are known as austempering and martempering.
[13]

Austempering[edit]
Main article: Austempering


Time-temperature transformation (TTT) diagram. The red line shows the cooling curve for austempering.
Austempering is a technique used to form pure bainite, a transitional microstructure found
between pearlite and martensite. In normalizing, both upper and lower bainite are usually found
mixed with pearlite. To avoid the formation of pearlite or martensite, the steel is quenched in a bath
of molten metals or salts. This quickly cools the steel past the point where pearlite can form, and into
the bainite-forming range. The steel is then held at the bainite-forming temperature, beyond the point
where the temperature reaches an equilibrium, until the bainite fully forms. The steel is then
removed from the bath and allowed to air-cool, without the formation of either pearlite or martensite.
Depending on the temperature, austempering can produce either upper or lower bainite. Upper
bainite is a laminate structure formed at temperatures typically above 350 C (662 F) and is a much
tougher microstructure. Lower bainite is a needle-like structure, produced at temperatures below 350
C, and is stronger but much more brittle.
[14]
In either case, austempering produces greater strength
and toughness for a given hardness, and reduced internal stresses which could lead to breakage.
This produces steel with superior impact resistance. Modern punches and chisels are often
austempered. Because austempering does not produce martensite, the steel does not require
further tempering.
[13]

Martempering[edit]
Main article: Martempering
Martempering is similar to austempering, in that the steel is quenched in a bath of molten metal or
salts to quickly cool it past the pearlite-forming range. However, in martempering, the goal is to
create martensite rather than bainite. The steel is quenched to a much lower temperature than is
used for austempering; to just above the martensite start temperature. The metal is then held at this
temperature until the temperature of the steel reaches an equilibrium. The steel is then removed
from the bath before any bainite can form, and then is allowed to air-cool, turning it into martensite.
The interruption in cooling allows much of the internal stresses to relax before the martensite forms,
decreasing the brittleness of the steel. However, the martempered steel will usually need to undergo
further tempering to adjust the hardness and toughness.
[13]

Physical processes[edit]
Tempering involves a three-step process in which unstable martensite decomposes into ferrite and
unstable carbides, and finally into stable cementite, forming various stages of a microstructure called
tempered martensite. The martensite typically consists of laths (strips) or plates, sometimes
appearing acicular (needle-like) or lenticular (lens-shaped). Depending on the carbon content, it also
contains a certain amount of "retained austenite." Retained austenite are crystals which are unable
to transform into martensite, even after quenching below the martensite finish (M
f
) temperature. An
increase in alloying agents or carbon content causes an increase in retained austenite. Austenite
has much higher stacking-fault energy than martensite, lowering the wear resistance of the steel,
although some or most of the retained austenite can be transformed into martensite by cold and
cryogenic treatments prior to tempering.
The martensite forms during a diffusionless transformation, in which the transformation occurs due
to shear-stresses created in the crystal lattices rather than by chemical changes that occur during
precipitation. The shear-stresses create many defects, or "dislocations," between the crystals,
providing less-stressful areas for the carbon atoms to relocate. Upon heating, the carbon atoms first
migrate to these defects, and then begin forming unstable carbides. This reduces the amount of total
martensite by changing some of it to ferrite. Further heating reduces the martensite even more,
transforming the unstable carbides into stable cementite.
The first stage of tempering occurs between room-temperature and 200 C (392 F). In the first
stage, carbon precipitates into -carbon (Fe
24
C). In the second stage, occurring between 150 C
(302 F) and 300 C (572 F), the retained austenite transforms into a form of lower-bainite
containing -carbon rather than cementite. The third stage occurs at 200 C (392 F) and higher. In
the third stage, -carbon precipitates into cementite, and the carbon content in the martensite
decreases. If tempered at higher temperatures, between 650 C (1,202 F) and 700 C (1,292 F), or
for longer amounts of time, the martensite may become fully ferritic and the cementite may become
coarser or spheroidize. In spheroidized steel, the cementite network breaks apart and recedes into
rods or spherical shaped globules, and the steel becomes softer than annealed steel; nearly as soft
as pure iron, making it very easy to form or machine.
[15]

Embrittlement[edit]
Embrittlement occurs during tempering when, through a specific temperature range, the steel
experiences an increase in hardness and a reduction in ductility, as opposed to the normal decrease
in hardness that occurs to either side of this range. The first type is called tempered martensite
embrittlement (TME) or one-step embrittlement. The second is referred to as temper embrittlement
(TE) or two-step embrittlement.
One-step embrittlement usually occurs in carbon steel at temperatures between 230 C (446 F) and
290 C (554 F), and was historically referred to as "500 F embrittlement." This embritttlement
occurs due to the precipitation of Widmanstatten needles or plates, made of cementite, in the
interlath boundaries of the martensite. Impurities such asphosphorus, or alloying agents
like manganese, may increase the embrittlement, or alter the temperature at which it occurs. This
type of embrittlement is permanent, and can only be relieved by heating above the upper critical
temperature and then quenching again. However, these microstructures usually require an hour or
more to form, so are usually not a problem in the blacksmith-method of tempering.
Two-step embrittlement typically occurs by aging the metal within a critical temperature range, or by
slowly cooling it through that range, For carbon steel, this is typically between 370 C (698 F) and
560 C (1,040 F), although impurities like phosphorus and sulfur increase the effect dramatically.
This generally occurs because the impurities are able to migrate to the grain boundaries, creating
weak spots in the structure. The embrittlement can often be avoided by quickly cooling the metal
after tempering. Two-step embrittlement, however, is reversible. The embrittlement can be
eliminated by heating the steel above 600 C (1,112 F) and then quickly cooling.
[16]

In alloy steels[edit]
Many elements are often alloyed with steel. The main purpose for alloying most elements with steel
is to increase its hardenability and to decrease softening under temperature. Tool steels, for
example, may have elements like chromium or vanadium added to increase both toughness and
strength, which is necessary for things like wrenches andscrewdrivers. On the other hand, drill
bits and rotary files need to retain their hardness at high temperatures.
Adding cobalt or molybdenum can cause the steel to retain its hardness, even at red-hot
temperatures, forming high-speed steels. Often, small amounts of many different elements are
added to the steel to give the desired properties, rather than just adding one or two.
Most alloying elements (solutes) have the benefit of not only increasing hardness, but also lowering
both the martensite start temperature and the temperature at which austenite transforms into ferrite
and cementite. During quenching, this allows a slower cooling rate, which allows items with thicker
cross-sections to be hardened to greater depths than is possible in plain carbon-steel, producing
more uniformity in strength.
Tempering methods for alloy steels may vary considerably, depending on the type and amount of
elements added. In general, elements like manganese, nickel, silicon, andaluminum will remain
dissolved in the ferrite during tempering while the carbon precipitates. When quenched, these
solutes will usually produce an increase in hardness over plain carbon-steel of the same carbon
content. When hardened alloy-steels, containing moderate amounts of these elements, are
tempered, the alloy will usually soften somewhat proportionately to carbon steel.
However, during tempering, elements like chromium, vanadium, and molybdenum precipitate with
the carbon. If the steel contains fairly low concentrations of these elements, the softening of the steel
can be retarded until much higher temperatures are reached, when compared to those needed for
tempering carbon steel. This allows the steel to maintain its hardness in high temperature or high
friction applications. However, this also requires very high temperatures during tempering, to achieve
a reduction in hardness. If the steel contains large amounts of these elements, tempering may
produce an increase in hardness until a specific temperature is reached, at which point the hardness
will begin to decrease.
[17][18]
For instance, molybdenum steels will typically reach their highest
hardness around 315 C (599 F) whereas vanadium steels will harden fully when tempered to
around 371 C (700 F). When very large amounts of solutes are added, alloy steels may behave
like precipitation hardening alloys, which do not soften at all during tempering.
[19]

In cast-iron[edit]
Cast-iron comes in many types, depending on the carbon-content. However, they are usually divided
into grey and white cast-iron, depending on the form that the carbides take. In grey cast iron, the
carbon is mainly in the form of graphite but, in white cast-iron, the carbon is usually in the form
of cementite. Grey cast-iron consists mainly of the microstructure called pearlite, mixed with graphite
and sometimes ferrite. Grey cast-iron is usually used as-cast, with its properties being determined by
its composition.
White cast-iron is composed mostly of a microstructure called ledeburite mixed with pearlite.
Ledeburite is very hard, making the cast-iron very brittle. If the white cast-iron has ahypoeutectic
composition, it is usually tempered to produce malleable cast-iron. Two methods of tempering are
used, called "white tempering" and "black tempering." The purposes of both tempering methods is to
cause the cementite to decompose from the ledeburite, increasing the ductility.
[20]

White tempering[edit]
White tempering is used to burn off excess carbon, by heating it for extended amounts of time in an
oxidizing environment. The cast iron will usually be held at temperatures as high as 1,000 C
(1,830 F) for as long as 60 hours. The heating is followed by a slow cooling rate of around 10 C
(18 F) per hour. The entire process may last 160 hours or more. This causes the cementite to
decompose from the ledeburite, and then the carbon burns out through the surface of the metal,
increasing the malleability of the cast-iron.
[20]

Black tempering[edit]
Unlike white tempering, black tempering is done in an inert gas environment, so that the
decomposing carbon does not burn off. Instead, the decomposing carbon turns into a type of
graphite called "temper graphite" or "flaky graphite," increasing the malleability of the metal.
Tempering is usually performed at temperatures as high as 950 C (1,740 F) for up to 20 hours.
The tempering is followed by slow-cooling through the lower critical temperature, over a period that
may last from 50 to over 100 hours.
[20]

Precipitation hardening alloys[edit]
Main article: Precipitation hardening
Precipitation hardening alloys first came into use during the early 1900s. Most heat-treatable alloys
fall into the category of precipitation hardening alloys, including alloys
ofaluminum, magnesium, titanium and nickel. Several high-alloy steels are also precipitation
hardening alloys. These alloys become softer than normal when quenched, and then harden over
time. For this reason, precipitation hardening is often referred to as "aging."
Although most precipitation hardening alloys will harden at room temperature, some will only harden
at elevated temperatures and, in others, the process can be sped up by aging at elevated
temperatures. Aging at temperatures higher than room-temperature is called "artificial aging".
Although the method is similar to tempering, the term "tempering" is usually not used to describe
artificial aging, because the physical processes, (i.e.: precipitation of intermetallic phases from
a supersaturated alloy) the desired results, (i.e.: strengthening rather than softening), and the
amount of time held at a certain temperature are very different from tempering as used in carbon-
steel.



Carbon steel
From Wikipedia, the free encyclopedia
Steels and other ironcarbon alloy
phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E
Carbon steel is steel in which the main interstitial alloying constituent is carbon in the range of 0.12
2.0%. The American Iron and Steel Institute (AISI) defines carbon steel as the following: "Steel is
considered to be carbon steel when no minimum content is specified or required
for chromium, cobalt, molybdenum, nickel, niobium, titanium, tungsten, vanadium or zirconium, or
any other element to be added to obtain a desired alloying effect; when the specified minimum for
copper does not exceed 0.40 percent; or when the maximum content specified for any of the
following elements does not exceed the percentages
noted: manganese 1.65,silicon 0.60, copper 0.60."
[1]

The term "carbon steel" may also be used in reference to steel which is not stainless steel; in this
use carbon steel may include alloy steels.
As the carbon percentage content rises, steel has the ability to
become harder and stronger through heat treating; however it becomes less ductile. Regardless of
the heat treatment, a higher carbon content reduces weldability. In carbon steels, the higher carbon
content lowers the melting point.
[2]

Contents
[hide]
1 Types
o 1.1 Mild and low-carbon steel
o 1.2 Higher carbon steels
2 Heat treatment
3 Case hardening
4 Forging temperature of steel
5 See also
6 References
7 Bibliography
Types[edit]
See also: SAE steel grades
Carbon steel is broken down into four classes based on carbon content:
Mild and low-carbon steel[edit]
Mild steel, also known as plain-carbon steel, is the most common form of steel because its price is
relatively low while it provides material properties that are acceptable for many applications, more so
than iron. Low-carbon steel contains approximately 0.050.320% carbon
[1]
making it malleable and
ductile. Mild steel has a relatively low tensile strength, but it is cheap and malleable; surface
hardness can be increased through carburizing.
[3]

It is often used when large quantities of steel are needed, for example as structural steel. The
density of mild steel is approximately 7.85 g/cm
3
(7850 kg/m
3
or 0.284 lb/in
3
)
[4]
and the Young's
modulus is 210 GPa (30,000,000 psi).
[5]

Low-carbon steels suffer from yield-point runout where the material has two yield points. The first
yield point (or upper yield point) is higher than the second and the yield drops dramatically after the
upper yield point. If a low-carbon steel is only stressed to some point between the upper and lower
yield point then the surface may develop Lder bands.
[6]
Low-carbon steels contain less carbon than
other steels and are easier to cold-form, making them easier to handle.
[7]

Higher carbon steels[edit]
Carbon steels which can successfully undergo heat-treatment have a carbon content in the range of
0.301.70% by weight. Trace impurities of various other elements can have a significant effect on
the quality of the resulting steel. Trace amounts of sulfur in particular make the steel red-short, that
is, brittle and crumbly at working temperatures. Low-alloy carbon steel, such as A36 grade, contains
about 0.05% sulfur and melts around 1,4261,538 C (2,5992,800 F).
[8]
Manganese is often added
to improve the hardenability of low-carbon steels. These additions turn the material into a low-alloy
steel by some definitions, but AISI's definition of carbon steel allows up to 1.65% manganese by
weight.
Low carbon steel
<0.3% carbon content, see above.
Medium carbon steel
Approximately 0.300.59% carbon content.
[1]
Balances ductility and strength and has good wear
resistance; used for large parts, forging and automotive components.
[9][10]

High-carbon steel (ASTM 304)
Approximately 0.60.99% carbon content.
[1]
Very strong, used for springs and high-strength wires.
[11]

Ultra-high-carbon steel
Approximately 1.02.0% carbon content.
[1]
Steels that can be tempered to great hardness. Used for
special purposes like (non-industrial-purpose) knives, axles or punches. Most steels with more than
1.2% carbon content are made using powder metallurgy. Note that steel with a carbon content
above 2.14% is considered cast iron.
Heat treatment[edit]


Iron-carbon phase diagram, showing the temperature and carbon ranges for certain types of heat treatments.
Main article: Heat treatment
The purpose of heat treating carbon steel is to change the mechanical properties of steel, usually
ductility, hardness, yield strength, or impact resistance. Note that the electrical and thermal
conductivity are only slightly altered. As with most strengthening techniques for steel,Young's
modulus (elasticity) is unaffected. All treatments of steel trade ductility for increased strength and
vice versa. Iron has a higher solubility for carbon in the austenite phase; therefore all heat
treatments, except spheroidizing and process annealing, start by heating the steel to a temperature
at which the austenitic phase can exist. The steel is then quenched (heat drawn out) at a high rate
causing cementite to precipitate and finally the remaining pure iron to solidify. The rate at which the
steel is cooled through the eutectoidtemperature affects the rate at which carbon diffuses out of
austenite and forms cementite. Generally speaking, cooling swiftly will leave iron carbide finely
dispersed and produce a fine grained pearlite (until the martensite critical temperature is reached)
and cooling slowly will give a coarser pearlite. Cooling a hypoeutectoid steel (less than 0.77 wt% C)
results in a lamellar-pearlitic structure of iron carbide layers with -ferrite (pure iron) between. If it is
hypereutectoid steel (more than 0.77 wt% C) then the structure is full pearlite with small grains
(larger than the pearlite lamella) of cementite scattered throughout. The relative amounts of
constituents are found using the lever rule. The following is a list of the types of heat treatments
possible:
Spheroidizing: Spheroidite forms when carbon steel is heated to approximately 700 C for over
30 hours. Spheroidite can form at lower temperatures but the time needed drastically increases,
as this is a diffusion-controlled process. The result is a structure of rods or spheres of cementite
within primary structure (ferrite or pearlite, depending on which side of the eutectoid you are on).
The purpose is to soften higher carbon steels and allow more formability. This is the softest and
most ductile form of steel. The image to the right shows where spheroidizing usually occurs.
[12]

Full annealing: Carbon steel is heated to approximately 40 C above Ac3 or Ac1 for 1 hour; this
ensures all the ferrite transforms into austenite (although cementite might still exist if the carbon
content is greater than the eutectoid). The steel must then be cooled slowly, in the realm of 20C
(36F) per hour. Usually it is just furnace cooled, where the furnace is turned off with the steel
still inside. This results in a coarse pearlitic structure, which means the "bands" of pearlite are
thick. Fully annealed steel is soft and ductile, with no internal stresses, which is often necessary
for cost-effective forming. Only spheroidized steel is softer and more ductile.
[13]

Process annealing: A process used to relieve stress in a cold-worked carbon steel with less
than 0.3 wt% C. The steel is usually heated up to 550650 C for 1 hour, but sometimes
temperatures as high as 700 C. The image rightward shows the area where process annealing
occurs.
Isothermal annealing: It is a process in which hypoeutectoid steel is heated above the upper
critical temperature and this temperature is maintained for a time and then the temperature is
brought down below lower critical temperature and is again maintained. Then finally it is cooled
at room temperature. This method rids any temperature gradient.
Normalizing: Carbon steel is heated to approximately 55 C above Ac3 or Acm for 1 hour; this
ensures the steel completely transforms to austenite. The steel is then air-cooled, which is a
cooling rate of approximately 38 C (100 F) per minute. This results in a fine pearlitic structure,
and a more-uniform structure. Normalized steel has a higher strength than annealed steel; it has
a relatively high strength and ductility.
[14]

Quenching: Carbon steel with at least 0.4 wt% C is heated to normalizing temperatures and
then rapidly cooled (quenched) in water, brine, or oil to the critical temperature. The critical
temperature is dependent on the carbon content, but as a general rule is lower as the carbon
content increases. This results in a martensitic structure; a form of steel that possesses a super-
saturated carbon content in a deformed body-centered cubic (BCC) crystalline structure,
properly termed body-centered tetragonal (BCT), with much internal stress. Thus quenched steel
is extremely hard but brittle, usually too brittle for practical purposes. These internal stresses
cause stress cracks on the surface. Quenched steel is approximately three to four (with more
carbon) fold harder than normalized steel.
[15]

Martempering (Marquenching): Martempering is not actually a tempering procedure, hence the
term "marquenching". It is a form of isothermal heat treatment applied after an initial quench of
typically in a molten salt bath at a temperature right above the "martensite start temperature". At
this temperature, residual stresses within the material are relieved and some bainite may be
formed from the retained austenite which did not have time to transform into anything else. In
industry, this is a process used to control the ductility and hardness of a material. With longer
marquenching, the ductility increases with a minimal loss in strength; the steel is held in this
solution until the inner and outer temperatures equalize. Then the steel is cooled at a moderate
speed to keep the temperature gradient minimal. Not only does this process reduce internal
stresses and stress cracks, but it also increases the impact resistance.
[16]

Quench and tempering: This is the most common heat treatment encountered, because the
final properties can be precisely determined by the temperature and time of the tempering.
Tempering involves reheating quenched steel to a temperature below the eutectoid temperature
then cooling. The elevated temperature allows very small amounts of spheroidite to form, which
restores ductility, but reduces hardness. Actual temperatures and times are carefully chosen for
each composition.
[17]

Austempering: The austempering process is the same as martempering, except the steel is
held in the molten salt bath through the bainite transformation temperatures, and then
moderately cooled. The resulting bainite steel has a greater ductility, higher impact resistance,
and less distortion. The disadvantage of austempering is it can only be used on a few steels,
and it requires a special salt bath.
[18]

Case hardening[edit]
Main article: Case hardening
Case hardening processes harden only the exterior of the steel part, creating a hard, wear resistant
skin (the "case") but preserving a tough and ductile interior. Carbon steels are not very hardenable;
therefore wide pieces cannot be through-hardened. Alloy steels have a better hardenability, so they
can through-harden and do not require case hardening. This property of carbon steel can be
beneficial, because it gives the surface good wear characteristics but leaves the core tough.
Forging temperature of steel[edit]
[19]

Steel Type
Maximum forging
temperature (F / C)
Burning
temperature (F /
C)
1.5% carbon 1920 / 1049 2080 / 1138
1.1% carbon 1980 / 1082 2140 / 1171
0.9% carbon 2050 / 1121 2230 / 1221
0.5% carbon 2280 / 1249 2460 / 1349
0.2% carbon 2410 / 1321 2680 / 1471
3.0% nickel steel 2280 / 1249 2500 / 1371
3.0% nickel
chromium steel
2280 / 1249 2500 / 1371
5.0% nickel (case-
hardening) steel
2320 / 1271 2640 / 1449
Chromium
vanadium steel
2280 /
1249
2460 /
1349
High-speed steel 2370 / 1299 2520 / 1382

Stainless steel 2340 / 1282 2520 / 1382

Austenitic
chromiumnickel
steel
2370 / 1299 2590 / 1421

Silico-manganese
spring steel
2280 / 1249 2460 / 1349

See also[edit]

Stainless steel
From Wikipedia, the free encyclopedia

This article may require copy editing for The Maintenance section which needs to be cleaned up
or rewritten.. You can assist by editing it. (February 2014)
Steels and other ironcarbon alloy phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E
In metallurgy, stainless steel, also known as inox steel or inox from French "inoxydable", is
a steel alloy with a minimum of 10.5%
[1]
chromium content by mass.
Stainless steel does not readily corrode, rust or stain with water as ordinary steel does, but despite
the name it is not fully stain-proof, most notably under low-oxygen, high-salinity, or poor-circulation
environments.
[2]
There are different grades and surface finishes of stainless steel to suit the
environment the alloy must endure. Stainless steel is used where both the properties of steel and
resistance to corrosion are required.
Stainless steel differs from carbon steel by the amount of chromium present. Unprotected carbon
steel rusts readily when exposed to air and moisture. This iron oxide film (the rust) is active and
accelerates corrosion by forming more iron oxide, and due to the greater volume of the iron oxide
this tends to flake and fall away. Stainless steels contain sufficient chromium to form a passive film
of chromium oxide, which prevents further surface corrosion by blocking oxygen diffusion to the steel
surface and blocks corrosion from spreading into the metal's internal structure, and due to the similar
size of the steel and oxide ions they bond very strongly and remain attached to the surface.
[3]

Passivation only occurs if the proportion of chromium is high enough and oxygen is present.



Cast iron
From Wikipedia, the free encyclopedia
For cookware, see Cast-iron cookware.

This article needs additional citations for verification. Please help improve this article by adding
citations to reliable sources. Unsourced material may be challenged and removed. (October 2008)


A cast iron pan.
Steels and other ironcarbon alloy phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E
Cast iron is iron or a ferrous alloy which has been heated until it liquefies, and is then poured into a
mould to solidify. It is usually made from pig iron. The alloy constituents affect its colour when
fractured: white cast iron has carbide impurities which allow cracks to pass straight through. Grey
cast iron has graphite flakes which deflect a passing crack and initiate countless new cracks as the
material breaks.
Carbon (C) and silicon (Si) are the main alloying elements, with the amount ranging from 2.14 wt%
and 13 wt%, respectively. Iron alloys with less carbon content are known as steel. While this
technically makes these base alloys ternary FeCSi alloys, the principle of cast iron solidification is
understood from the binary ironcarbon phase diagram. Since the compositions of most cast irons
are around the eutectic point of the ironcarbon system, the melting temperatures closely correlate,
usually ranging from 1,150 to 1,200 C (2,100 to 2,190 F), which is about 300 C (572 F) lower
than the melting point of pure iron.
Cast iron tends to be brittle, except for malleable cast irons. With its relatively low melting point,
good fluidity, castability, excellent machinability, resistance to deformation and wear resistance, cast
irons have become an engineering material with a wide range of applications and are used in pipes,
machines and automotive industry parts, such as cylinder heads (declining usage), cylinder
blocks and gearbox cases (declining usage). It is resistant to destruction and weakening
by oxidation (rust).
The earliest cast iron artefacts date to the 5th century BC, and were discovered by archaeologists in
what is now modern Luhe County, Jiangsu in China. Cast iron was used in ancient China for
warfare, agriculture, and architecture.
[1]
During the 15th century, cast iron became utilized for artillery
in Burgundy, France, and in England during the Reformation.
[2]
The first cast iron bridge was built
during the 1770s by Abraham Darby III, and is known as The Iron Bridge. Cast iron is also used in
the construction of buildings.
Contents
[hide]
1 Production
2 Types
o 2.1 Alloying elements
o 2.2 Grey cast iron
o 2.3 White cast iron
o 2.4 Malleable cast iron
o 2.5 Ductile cast iron
o 2.6 Table of comparative qualities of cast irons
3 History
o 3.1 Cast-iron bridges
o 3.2 Buildings
o 3.3 Textile mills
4 See also
5 References
6 Further reading
7 External links
Production[edit]
Cast iron is made by re-melting pig iron, often along with substantial quantities of scrap iron, scrap
steel, lime stone, carbon (coke) and taking various steps to remove undesirable
contaminants. Phosphorus and sulfur may be burnt out of the molten iron, but this also burns out the
carbon, which must be replaced. Depending on the application, carbon and silicon content are
adjusted to the desired levels, which may be anywhere from 23.5% and 13%, respectively. Other
elements are then added to the melt before the final form is produced by casting.
[citation needed]

Iron is sometimes melted in a special type of blast furnace known as a cupola, but more often melted
in electric induction furnaces or electric arc furnaces.
[citation needed]
After melting is complete, the
molten iron is poured into a holding furnace or ladle.
Types[edit]
Alloying elements[edit]


Iron-cementite meta-stable diagram.
Cast iron's properties are changed by adding various alloying elements, or alloyants. Next
to carbon, silicon is the most important alloyant because it forces carbon out of solution. Instead the
carbon forms graphite which results in a softer iron, reduces shrinkage, lowers strength, and
decreases density. Sulfur, when present, forms iron sulfide, which prevents the formation of graphite
and increaseshardness. The problem with sulfur is that it makes molten cast iron sluggish, which
causes short run defects. To counter the effects of sulfur, manganese is added because the two
form into manganese sulfide instead of iron sulfide. The manganese sulfide is lighter than the melt
so it tends to float out of the melt and into the slag. The amount of manganese required to neutralize
sulfur is 1.7 sulfur content + 0.3%. If more than this amount of manganese is added,
then manganese carbide forms, which increases hardness and chilling, except in grey iron, where up
to 1% of manganese increases strength and density.
[3]

Nickel is one of the most common alloying elements because it refines the pearlite and graphite
structure, improves toughness, and evens out hardness differences between section
thicknesses. Chromium is added in small amounts to the ladle to reduce free graphite, produce chill,
and because it is a powerful carbide stabilizer; nickel is often added in conjunction. A small amount
of tin can be added as a substitute for 0.5% chromium. Copper is added in the ladle or in the
furnace, on the order of 0.52.5%, to decrease chill, refine graphite, and increase
fluidity. Molybdenum is added on the order of 0.31% to increase chill and refine the graphite and
pearlite structure; it is often added in conjunction with nickel, copper, and chromium to form high
strength irons. Titanium is added as a degasser and deoxidizer, but it also increases fluidity. 0.15
0.5% vanadium is added to cast iron to stabilize cementite, increase hardness, and increase
resistance to wear and heat. 0.10.3% zirconium helps to form graphite, deoxidize, and increase
fluidity.
[3]

In malleable iron melts, bismuth is added, on the scale of 0.0020.01%, to increase how much
silicon can be added. In white iron, boron is added to aid in the production of malleable iron; it also
reduces the coarsening effect of bismuth.
[3]

Grey cast iron[edit]
Main article: Grey iron
Grey cast iron is characterised by its graphitic microstructure, which causes fractures of the material
to have a grey appearance. It is the most commonly used cast iron and the most widely used cast
material based on weight. Most cast irons have a chemical composition of 2.54.0% carbon, 13%
silicon, and the remainder is iron. Grey cast iron has lesstensile strength and shock resistance than
steel, but its compressive strength is comparable to low and medium carbon steel.
White cast iron[edit]
It is the cast iron that displays white fractured surface due to the presence of cementite. With a lower
silicon content (graphitizing agent) and faster cooling rate, the carbon in white cast iron precipitates
out of the melt as the metastable phase cementite, Fe
3
C, rather than graphite. The cementite which
precipitates from the melt forms as relatively large particles, usually in a eutectic mixture, where the
other phase is austenite (which on cooling might transform to martensite). These eutectic carbides
are much too large to provide precipitation hardening (as in some steels, where cementite
precipitates might inhibit plastic deformation by impeding the movement of dislocations through the
ferrite matrix). Rather, they increase the bulk hardness of the cast iron simply by virtue of their own
very high hardness and their substantial volume fraction, such that the bulk hardness can be
approximated by a rule of mixtures. In any case, they offer hardness at the expense of toughness.
Since carbide makes up a large fraction of the material, white cast iron could reasonably be
classified as a cermet. White iron is too brittle for use in many structural components, but with good
hardness and abrasion resistance and relatively low cost, it finds use in such applications as the
wear surfaces (impeller and volute) of slurry pumps, shell liners and lifter bars in ball
mills and autogenous grinding mills, balls and rings incoal pulverisers, and the teeth of a backhoe's
digging bucket (although cast medium-carbon martensitic steel is more common for this application).
It is difficult to cool thick castings fast enough to solidify the melt as white cast iron all the way
through. However, rapid cooling can be used to solidify a shell of white cast iron, after which the
remainder cools more slowly to form a core of grey cast iron. The resulting casting, called a chilled
casting, has the benefits of a hard surface and a somewhat tougher interior.
High-chromium white iron alloys allow massive castings (for example, a 10-tonne impeller) to be
sand cast, i.e., a high cooling rate is not required, as well as providing impressive abrasion
resistance.
[citation needed]
These high-chromium alloys attribute their superior hardness to the presence
of chromium carbides. The main form of these carbides are the eutectic or primary M
7
C
3
carbides,
where "M" represents iron or chromium and can vary depending on the alloy's composition. The
eutectic carbides form as bundles of hollow hexagonal rods and grow perpendicular to the to the
hexagonal basal plane. The hardness of these carbides are within the range of 1500-1800HV
[4]

Malleable cast iron[edit]
Main article: Malleable iron
Malleable iron starts as a white iron casting that is then heat treated at about 900 C (1,650 F).
Graphite separates out much more slowly in this case, so that surface tension has time to form it into
spheroidal particles rather than flakes. Due to their lower aspect ratio, spheroids are relatively short
and far from one another, and have a lower cross sectionvis-a-vis a propagating crack or phonon.
They also have blunt boundaries, as opposed to flakes, which alleviates the stress concentration
problems faced by grey cast iron. In general, the properties of malleable cast iron are more like mild
steel. There is a limit to how large a part can be cast in malleable iron, since it is made from white
cast iron.
Ductile cast iron[edit]
Main article: Ductile cast iron
A more recent development is nodular or ductile cast iron. Tiny amounts
of magnesium or cerium added to these alloys slow down the growth of graphite precipitates by
bonding to the edges of the graphite planes. Along with careful control of other elements and timing,
this allows the carbon to separate as spheroidal particles as the material solidifies. The properties
are similar to malleable iron, but parts can be cast with larger sections.
Table of comparative qualities of cast irons[edit]
Comparative qualities of cast irons
[5]

Name
Nominal
composition
[% by
weight]
Form
and
condition
Yield
strength
[ksi(0.2
%
offset)]
Tensile
strengt
h [ksi]
Elongatio
n [% (in
2 inches)]
Hardnes
s [Brinell
scale]
Uses
Grey cast
iron
(ASTMA48
C 3.4,
Si 1.8, Mn 0.5
Cast 50 0.5 260
Engine cylinder
blocks, flywheels, gearbo
x cases, machine-tool
) bases
White cast
iron
C 3.4, Si 0.7,
Mn 0.6
Cast (as
cast)
25 0 450 Bearing surfaces
Malleable
iron
(ASTM
A47)
C 2.5, Si 1.0,
Mn 0.55
Cast
(annealed
)
33 52 12 130
Axle bearings, track
wheels,
automotivecrankshafts
Ductile or
nodular
iron
C 3.4, P 0.1,
Mn 0.4,Ni 1.0
, Mg 0.06
Cast 53 70 18 170
Gears, camshafts,
crankshafts
Ductile or
nodular
iron
(ASTM
A339)

cast
(quench
tempered)
108 135 5 310
Ni-hard
type 2
C 2.7, Si 0.6,
Mn 0.5,
Ni 4.5, Cr 2.0
Sand-cast 55 550
High strength
applications
Ni-resist
type 2
C 3.0, Si 2.0,
Mn 1.0,
Ni 20.0,
Cr 2.5
Cast 27 2 140
Resistance to heat and
corrosion
History[edit]


Cast iron artifact dated from 5th century BC found in Jiangsu, China.


Cast iron drain, waste and vent piping


Cast iron plate on grand piano
The earliest cast iron artifacts date to the 5th century BC, and were discovered by archaeologists in
what is now modern Luhe County, Jiangsu in China. This is based on an analysis of the artifact's
microstructures.
[1]
Because cast iron is comparatively brittle, it is not suitable for purposes where a
sharp edge or flexibility is required. It is strong under compression, but not under tension. Cast iron
was invented in China in the 5th century BC and poured into moulds to make ploughshares and pots
as well as weapons and pagodas.
[6]
Although steel was more desirable, cast iron was cheaper and
thus was more commonly used for implements in ancient China, while wrought iron or steel was
used for weapons.
[1]

In the west, where it did not become available till the 15th century, its earliest uses included cannon
and shot. Henry VIII initiated the casting of cannon in England. Soon, English iron workers
using blast furnaces developed the technique of producing cast iron cannons, which, while heavier
than the prevailing bronze cannons, were much cheaper and enabled England to arm her navy
better. The ironmasters of the Weald continued producing cast irons until the 1760s and armament
was one of the main uses of irons after the Restoration.
Cast iron pots were made at many English blast furnaces at the time. In 1707, Abraham
Darby patented a method of making pots (and kettles) thinner and hence cheaper than his rivals
could. This meant that his Coalbrookdale furnaces became dominant as suppliers of pots, an activity
in which they were joined in the 1720s and 1730s by a small number of other coke-fired blast
furnaces.
The development of the steam engine by Thomas Newcomen provided further market for cast iron,
since cast iron was considerably cheaper than the brass of which the engine cylinders were
originally made. John Wilkinson was a great exponent of cast iron, who, amongst other things, cast
the cylinders for many of James Watt's improved steam engines until the establishment of the Soho
Foundry in 1795.
Cast-iron bridges[edit]
See also: The Iron Bridge
The use of cast iron for structural purposes began in the late 1770s, when Abraham Darby
III built the Iron Bridge, although short beams had already been used, such as in the blast furnaces
at Coalbrookdale. Other inventions followed, including one patented by Thomas Paine. Cast iron
bridges became commonplace as the Industrial Revolution gathered pace. Thomas Telford adopted
the material for his bridge upstream at Buildwas, and then for a canal trough aqueduct at Longdon-
on-Tern on the Shrewsbury Canal.
It was followed by the Chirk Aqueduct and the Pontcysyllte Aqueduct, both of which remain in use
following the recent restorations. Cast iron beam bridges were used widely by the early railways,
such as the Water Street Bridge at the Manchester terminus of the Liverpool and Manchester
Railway. Problems arose when a new bridge carrying the Chester and Holyhead Railway across
the River Dee in Chestercollapsed in May 1847, less than a year after it was opened. The Dee
bridge disaster was caused by excessive loading at the centre of the beam by a passing train, and
many similar bridges had to be demolished and rebuilt, often in wrought iron. The bridge had been
erroneously designed, being trussed with wrought iron straps, which were wrongly thought to
reinforce the structure. The centres of the beams were put into bending, with the lower edge in
tension, where cast iron, like masonry, is very weak.
The best way of using cast iron for bridge construction was by using arches, so that all the material
is in compression. Cast iron, again like masonry, is very strong in compression. Wrought iron, like
most other kinds of iron and indeed like most metals in general, is strong in tension, and
also tough resistant to fracturing. The relationship between wrought iron and cast iron, for
structural purposes, may be thought of as analogous to the relationship between wood and stone.
Nevertheless, cast iron continued to be used in inappropriate structural ways, until the Tay Rail
Bridge disaster of 1879 cast serious doubt on the use of the material. Crucial lugs for holding tie bars
and struts in the Tay Bridge had been cast integral with the columns and they failed in the early
stages of the accident. In addition, the bolt holes were also cast and not drilled, so that all the
tension from the tie bars was placed on a corner, rather than being spread over the length of the
hole. The replacement bridge was built in wrought iron and steel.
Further bridge collapses occurred, however, culminating in the Norwood Junction rail accident of
1891. Thousands of cast iron rail underbridges were eventually replaced by steel equivalents.


Gray iron
From Wikipedia, the free encyclopedia
Gray iron, or grey iron, is a type of cast iron that has a graphitic microstructure. It is named after
the gray color of the fracture it forms, which is due to the presence of graphite.
[1]
It is the most
common cast iron and the most widely used cast material based on weight.
[2]

It is used for housings where tensile strength is non-critical, such as internal combustion
engine cylinder blocks, pump housings, valve bodies, electrical boxes, and decorativecastings. Grey
cast iron's high thermal conductivity and specific heat capacity are often exploited to make cast iron
cookware and disc brake rotors.
[3]

Contents
[hide]
1 Structure
2 Classifications
3 Advantages and disadvantages
4 See also
5 Notes
6 References
7 Further reading
Structure[edit]
A typical chemical composition to obtain a graphitic microstructure is 2.5 to 4.0% carbon and 1 to
3% silicon. Silicon is important to making grey iron as opposed to white cast iron, because silicon is
a graphite stabilizing element in cast iron, which means it helps the alloy produce graphite instead
of iron carbides. Another factor affecting graphitization is the solidification rate; the slower the rate,
the greater the tendency for graphite to form. A moderate cooling rate forms a more pearlitic matrix,
while a fast cooling rate forms a moreferritic matrix. To achieve a fully ferritic matrix the alloy must
be annealed.
[1][4]
Rapid cooling partly or completely suppresses graphitization and leads to formation
of cementite, which is called white iron.
[5]

The graphite takes on the shape of a three-dimensional flake. In two dimensions, as a polished
surface will appear under a microscope, the graphite flakes appear as fine lines. The graphite has no
appreciable strength, so they can be treated as voids. The tips of the flakes act as preexisting
notches; therefore, it is brittle.
[5][6]
The presence of graphite flakes makes the Grey Iron easily
machinable as they tend to crack easily across the graphite flakes.Grey iron also has very good
damping capacity and hence it is mostly used as the base for machine tool mountings.
Classifications[edit]
In the United States, the most commonly used classification for gray iron is ASTM
International standard A48.
[2]
This classifies gray iron into classes which corresponds with its
minimum tensile strength in thousands of pounds per square inch (ksi); e.g. class 20 gray iron has a
minimum tensile strength of 20,000 psi (140 MPa). Class 20 has a highcarbon equivalent and a
ferrite matrix. Higher strength gray irons, up to class 40, have lower carbon equivalents and
a pearlite matrix. Gray iron above class 40 requires alloying to provide solid solution strengthening,
and heat treating is used to modify the matrix. Class 80 is the highest class available, but it is
extremely brittle.
[5]
ASTM A247 is also commonly used to describe the graphite structure. Other
ASTM standards that deal with gray iron include ASTM A126, ASTM A278, and ASTM A319.
[2]

In the automotive industry the SAE International (SAE) standard SAE J431 is used to
designate grades instead of classes. These grades are a measure of the tensile strength-to-Brinell
hardness ratio

Wrought iron
From Wikipedia, the free encyclopedia


The Eiffel tower is constructed frompuddled iron, a form of wrought iron
Steels and other ironcarbon alloy phases
Ferrite
Austenite
Cementite
Graphite
Martensite
Microstructures
Spheroidite
Pearlite
Bainite
Ledeburite
Tempered martensite
Widmanstatten structures
Classes
Crucible steel
Carbon steel
Spring steel
Alloy steel
Maraging steel
Stainless steel
Weathering steel
Tool steel
Other iron-based materials
Cast iron
Gray iron
White iron
Ductile iron
Malleable iron
Wrought iron
V
T
E


Iron pillar at Delhi, India, containing 98% wrought iron
Wrought iron is an iron alloy with a very low carbon (0.1 to 0.25%) content in contrast to cast
iron (2.1% to 4%), and has fibrousinclusions, known as slag up to 2% by weight. It is a semi-fused
mass of iron with slag inclusions which gives it a "grain" resembling wood, that is visible when it is
etched or bent to the point of failure. Wrought iron is tough, malleable, ductile and easily welded.
Historically, it was known as commercially pure iron;
[1][2]
however, it no longer qualifies because
current standards for commercially pure iron require a carbon content of less than 0.008 wt%.
[3][4]

Before the development of effective methods of steelmaking and the availability of large quantities of
steel, wrought iron was the most common form of malleable iron. A modest amount of wrought iron
was used as a raw material for refining into steel, which was used mainly to produce swords, cutlery,
chisels, axes and other edged tools as well as springs and files. The demand for wrought iron
reached its peak in the 1860s with the adaptation of ironclad warships and railways, but then
declined as mild steel quality problems such as brittleness were solved and it became inexpensive
and widely available.
Many items, before they came to be made of mild steel, were produced from wrought iron,
including rivets, nails, wire, chains, rails, railway couplings, water and steam
pipes, nuts, bolts, horseshoes, handrails, straps for timber roof trusses, and ornamental ironwork.
[5]

Wrought iron is no longer produced on a commercial scale. Many products described as wrought
iron, such as guard rails, garden furniture
[6]
and gates, are made of mild steel.
[7]
They retain that
description because in the past they were wrought (worked) by hand.
[8]

Contents
[hide]
1 Terminology
o 1.1 Types and shapes
o 1.2 Origin
o 1.3 Quality
o 1.4 Defects
2 History
o 2.1 Bloomery process
o 2.2 Osmond process
o 2.3 Finery process
o 2.4 Potting and stamping
o 2.5 Puddling process
2.5.1 Shingling
2.5.2 Rolling
o 2.6 Lancashire process
o 2.7 Aston process
o 2.8 Decline
3 Properties
o 3.1 Ductility
o 3.2 Purity
4 Applications
5 See also
6 References
7 Further reading
Terminology[edit]
The word "wrought" is an archaic past participle of the verb "to work," and so "wrought iron" literally
means "worked iron". Wrought iron is a general term for the commodity, but is also used more
specifically for finished iron goods, as manufactured by a blacksmith or other smith. It was used in
that narrower sense in British Customs records, such manufactured iron being subject to a higher
rate of duty than what might be called "unwrought" iron. Cast iron, unlike wrought iron, is brittle and
cannot be worked either hot or cold. Cast iron can break if struck with a hammer.
In the 17th, 18th and 19th centuries, wrought iron went by a wide variety of terms according to its
form, origin, or quality.
While the bloomery process produced wrought iron directly from ore, cast iron or pig iron were the
starting materials used in the finery forge and puddling furnace. Pig iron and cast iron have high
carbon content which make them very brittle, but they have a lower melting point than iron or steel.
Cast and especially pig iron have excess slag which must be at least partially removed to produce
quality wrought iron. At foundries it was common to blend scrap wrought iron with cast iron to
improve the physical properties of castings.
For several years after the introduction of Bessemer and open hearth steel, there were different
opinions as to what differentiated iron from steel, some believed it was the chemical composition and
others that it was whether the iron heated sufficiently to melt and "fuse". Fusion eventually became
generally accepted as relatively more important than composition below a given low carbon
concentration.
[9]
Another difference is that steel can be hardened by heat treating

You might also like