0% found this document useful (0 votes)
11 views45 pages

Iron, Fe: This Article Is About The Metallic Element. For Other Uses, See

Uploaded by

Abdi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
11 views45 pages

Iron, Fe: This Article Is About The Metallic Element. For Other Uses, See

Uploaded by

Abdi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 45

This article is about the metallic element. For other uses, see Iron (disambiguation).

Iron, 26Fe

Iron

Pronunciation /ˈaɪərn/

Allotropes see Allotropes of iron

Appearance lustrous metallic with a grayish


tinge

Standard atomic weight Ar°(Fe)

55.845±0.002[1]
55.845±0.002 (abridged)[2]

Iron in the periodic table


H
yd
ro
ge
n
Lit Be
hi ryl
u liu
m m
So M
di ag
u ne
m siu
m

Po Ca Sc
tas lci an
siuu diu
m m m

RuStr Ytt
bi on riu
di tiu m
u m
m
Ca Ba La C Pra Ne Pr Sa Eu Ga Te Dy Ho Er Th Yt Lu
esi riu nt er seo od om m ro dolrbi spr lm bi uli ter teti
u m ha iu dy ym eth ari pi ini u osi iu u um bi um
m nu m miuiu iu u u umm um m m u
m m m m mm m

Fr Ra Ac T Pro Ur Ne Pl A Cu Be Ca Ei FeMe N La
an di tin h tact ani ptu ut meriu rk lif nst r nde ob wr
ci u iu or iniuum niu on ric m eli or ein m lev eli en
u m m iu m m iu iu u niu iu iu iu u ciu
m m mm m m m mm m m
manganese ←

Atomic number (Z) 26

Group group 8

Period period 4

Block d-block

Electron [Ar] 3d6 4s2


configuration

Electrons per shell 2, 8, 14, 2

Physical properties

Phase at STP solid

Melting point 1811 K (1538 °C, 2800 °F)

Boiling point 3134 K (2861 °C, 5182 °F)

Density (at 20° C) 7.874 g/cm3 [3]

when liquid (at m.p.) 6.98 g/cm3

Heat of fusion 13.81 kJ/mol

Heat of vaporization 340 kJ/mol

Molar heat capacity 25.10 J/(mol·K)


Vapor pressure
P (Pa) 1 10 100 1k 10 k 100 k
at T (K 1728 1890 2091 2346 2679 3132
)

Atomic properties

Oxidation states common: +2, +3


−4,? −2,[4] −1,[4] 0,? +1,[5] +4,[4] +5,[6]
+6,[4] +7[7]

Electronegativity Pauling scale: 1.83

Ionization energies 1st: 762.5 kJ/mol


2nd: 1561.9 kJ/mol
3rd: 2957 kJ/mol
(more)

Atomic radius empirical: 126 pm

Covalent radius Low spin: 132±3 pm


High spin: 152±6 pm

Van der Waals 194 [1] pm


radius

Spectral
lines of iron

Other properties

Natural occurrence primordial

Crystal structure α-Fe: body-centered cubic (bcc)


(cI2)

Lattice constant

a = 286.65 pm (at 20 °C)[3]

Crystal structure γ-Fe (912–1394 °C): face-


centered cubic (fcc) (cF4)
Lattice constant

a = 364.68 pm (at 916 °C)[8]

Thermal expansion 12.07×10−6/K (at 20 °C)[3]

Thermal 80.4 W/(m⋅K)


conductivity

Electrical resistivity 96.1 nΩ⋅m (at 20 °C)

Curie point 1043 K

Magnetic ordering ferromagnetic

Young's modulus 211 GPa

Shear modulus 82 GPa

Bulk modulus 170 GPa

Speed of 5120 m/s (at r.t.) (electrolytic)


sound thin rod

Poisson ratio 0.29

Mohs hardness 4

Vickers hardness 608 MPa

Brinell hardness 200–1180 MPa

CAS Number 7439-89-6

History

Discovery before 5000 BC

Symbol "Fe": from Latin ferrum

Isotopes of iron
 v
 e
Main isotopes[9] Decay
abundance half-life (t1/2) mode product
54
Fe 5.85% stable
55
Fe synth 2.73 y ε 55
Mn
56
Fe 91.8% stable
57
Fe 2.12% stable
58
Fe 0.28% stable
59
Fe synth 44.6 d β− 59
Co
60
Fe trace 2.6×10 y β−
6 60
Co

Category: Iron
 view
 talk
 edit
| references

Iron is a chemical element; it has the symbol Fe (from Latin ferrum 'iron') and atomic
number 26. It is a metal that belongs to the first transition series and group 8 of
the periodic table. It is, by mass, the most common element on Earth, forming much of
Earth's outer and inner core. It is the fourth most abundant element in the Earth's crust,
being mainly deposited by meteorites in its metallic state.

Extracting usable metal from iron ores requires kilns or furnaces capable of reaching
1,500 °C (2,730 °F), about 500 °C (932 °F) higher than that required to smelt copper.
Humans started to master that process in Eurasia during the 2nd millennium BC and the
use of iron tools and weapons began to displace copper alloys – in some regions, only
around 1200 BC. That event is considered the transition from the Bronze Age to the Iron
Age. In the modern world, iron alloys, such as steel, stainless steel, cast
iron and special steels, are by far the most common industrial metals, due to their
mechanical properties and low cost. The iron and steel industry is thus very important
economically, and iron is the cheapest metal, with a price of a few dollars per kilogram
or pound.

Pristine and smooth pure iron surfaces are a mirror-like silvery-gray. Iron reacts readily
with oxygen and water to produce brown-to-black hydrated iron oxides, commonly
known as rust. Unlike the oxides of some other metals that form passivating layers, rust
occupies more volume than the metal and thus flakes off, exposing more fresh surfaces
for corrosion. Chemically, the most common oxidation states of iron
are iron(II) and iron(III). Iron shares many properties of other transition metals, including
the other group 8 elements, ruthenium and osmium. Iron forms compounds in a wide
range of oxidation states, −4 to +7. Iron also forms many coordination compounds;
some of them, such as ferrocene, ferrioxalate, and Prussian blue have substantial
industrial, medical, or research applications.

The body of an adult human contains about 4 grams (0.005% body weight) of iron,
mostly in hemoglobin and myoglobin. These two proteins play essential roles in oxygen
transport by blood and oxygen storage in muscles. To maintain the necessary
levels, human iron metabolism requires a minimum of iron in the diet. Iron is also the
metal at the active site of many important redox enzymes dealing with cellular
respiration and oxidation and reduction in plants and animals.[10]
Characteristics
Allotropes
Main article: Allotropes of iron

Molar volume vs. pressure for α iron at room


temperature
At least four allotropes of iron (differing atom arrangements in the solid) are known,
conventionally denoted α, γ, δ, and ε.

The first three forms are observed at ordinary pressures. As molten iron cools past its
freezing point of 1538 °C, it crystallizes into its δ allotrope, which has a body-centered
cubic (bcc) crystal structure. As it cools further to 1394 °C, it changes to its γ-iron
allotrope, a face-centered cubic (fcc) crystal structure, or austenite. At 912 °C and
below, the crystal structure again becomes the bcc α-iron allotrope.[11]

The physical properties of iron at very high pressures and temperatures have also been
studied extensively,[12][13] because of their relevance to theories about the cores of the
Earth and other planets. Above approximately 10 GPa and temperatures of a few
hundred kelvin or less, α-iron changes into another hexagonal close-packed (hcp)
structure, which is also known as ε-iron. The higher-temperature γ-phase also changes
into ε-iron, but does so at higher pressure.

Some controversial experimental evidence exists for a stable β phase at pressures


above 50 GPa and temperatures of at least 1500 K. It is supposed to have
an orthorhombic or a double hcp structure.[14] (Confusingly, the term "β-iron" is
sometimes also used to refer to α-iron above its Curie point, when it changes from being
ferromagnetic to paramagnetic, even though its crystal structure has not changed.[11])

The inner core of the Earth is generally presumed to consist of an iron-nickel alloy with ε
(or β) structure.[15]

Melting and boiling points


Low-pressure phase diagram of pure iron
The melting and boiling points of iron, along with its enthalpy of atomization, are lower
than those of the earlier 3d elements from scandium to chromium, showing the
lessened contribution of the 3d electrons to metallic bonding as they are attracted more
and more into the inert core by the nucleus;[16] however, they are higher than the values
for the previous element manganese because that element has a half-filled 3d sub-shell
and consequently its d-electrons are not easily delocalized. This same trend appears
for ruthenium but not osmium.[17]

The melting point of iron is experimentally well defined for pressures less than 50 GPa.
For greater pressures, published data (as of 2007) still varies by tens of gigapascals
and over a thousand kelvin.[18]

Magnetic properties

Magnetization curves of 9 ferromagnetic materials,


showing saturation. 1. Sheet steel, 2. Silicon steel, 3. Cast steel, 4. Tungsten steel, 5. Magnet
steel, 6. Cast iron, 7. Nickel, 8. Cobalt, 9. Magnetite[19]
Below its Curie point of 770 °C (1,420 °F; 1,040 K), α-iron changes
from paramagnetic to ferromagnetic: the spins of the two unpaired electrons in each
atom generally align with the spins of its neighbors, creating an overall magnetic field.
[20]
This happens because the orbitals of those two electrons (dz and dx − y ) do not point
2 2 2

toward neighboring atoms in the lattice, and therefore are not involved in metallic
bonding.[11]

In the absence of an external source of magnetic field, the atoms get spontaneously
partitioned into magnetic domains, about 10 micrometers across,[21] such that the atoms
in each domain have parallel spins, but some domains have other orientations. Thus a
macroscopic piece of iron will have a nearly zero overall magnetic field.

Application of an external magnetic field causes the domains that are magnetized in the
same general direction to grow at the expense of adjacent ones that point in other
directions, reinforcing the external field. This effect is exploited in devices that need to
channel magnetic fields to fulfill design function, such as electrical
transformers, magnetic recording heads, and electric motors. Impurities, lattice defects,
or grain and particle boundaries can "pin" the domains in the new positions, so that the
effect persists even after the external field is removed – thus turning the iron object into
a (permanent) magnet.[20]

Similar behavior is exhibited by some iron compounds, such as the ferrites including the
mineral magnetite, a crystalline form of the mixed iron(II,III) oxide Fe3O4 (although the
atomic-scale mechanism, ferrimagnetism, is somewhat different). Pieces of magnetite
with natural permanent magnetization (lodestones) provided the earliest compasses for
navigation. Particles of magnetite were extensively used in magnetic recording media
such as core memories, magnetic tapes, floppies, and disks, until they were replaced
by cobalt-based materials.

Isotopes
Main article: Isotopes of iron
Iron has four stable isotopes: 54Fe (5.845% of natural iron), 56Fe (91.754%), 57Fe
(2.119%) and 58Fe (0.282%). Twenty-four artificial isotopes have also been created. Of
these stable isotopes, only 57Fe has a nuclear spin (−1⁄2). The nuclide 54Fe theoretically
can undergo double electron capture to 54Cr, but the process has never been observed
and only a lower limit on the half-life of 4.4×1020 years has been established.[22]

Fe is an extinct radionuclide of long half-life (2.6 million years).[23] It is not found on


60

Earth, but its ultimate decay product is its granddaughter, the stable nuclide 60Ni.[9] Much
of the past work on isotopic composition of iron has focused on
the nucleosynthesis of 60Fe through studies of meteorites and ore formation. In the last
decade, advances in mass spectrometry have allowed the detection and quantification
of minute, naturally occurring variations in the ratios of the stable isotopes of iron. Much
of this work is driven by the Earth and planetary science communities, although
applications to biological and industrial systems are emerging.[24]

In phases of the meteorites Semarkona and Chervony Kut, a correlation between the
concentration of 60Ni, the granddaughter of 60Fe, and the abundance of the stable iron
isotopes provided evidence for the existence of 60Fe at the time of formation of the Solar
System. Possibly the energy released by the decay of 60Fe, along with that released
by 26Al, contributed to the remelting and differentiation of asteroids after their formation
4.6 billion years ago. The abundance of 60Ni present in extraterrestrial material may
bring further insight into the origin and early history of the Solar System.[25]

The most abundant iron isotope 56Fe is of particular interest to nuclear scientists
because it represents the most common endpoint of nucleosynthesis.[26] Since 56Ni
(14 alpha particles) is easily produced from lighter nuclei in the alpha process in nuclear
reactions in supernovae (see silicon burning process), it is the endpoint of fusion chains
inside extremely massive stars. Although adding more alpha particles is possible, but
nonetheless the sequence does effectively end at 56Ni because conditions in stellar
interiors cause the competition between photodisintegration and the alpha process to
favor photodisintegration around 56Ni.[27][28] This 56Ni, which has a half-life of about 6 days,
is created in quantity in these stars, but soon decays by two successive positron
emissions within supernova decay products in the supernova remnant gas cloud, first to
radioactive 56Co, and then to stable 56Fe. As such, iron is the most abundant element in
the core of red giants, and is the most abundant metal in iron meteorites and in the
dense metal cores of planets such as Earth.[29] It is also very common in the universe,
relative to other stable metals of approximately the same atomic weight.[29][30] Iron is the
sixth most abundant element in the universe, and the most common refractory element.
[31]

Photon mass attenuation coefficient for iron


Although a further tiny energy gain could be extracted by synthesizing 62Ni, which has a
marginally higher binding energy than 56Fe, conditions in stars are unsuitable for this
process. Element production in supernovas greatly favor iron over nickel, and in any
case, 56Fe still has a lower mass per nucleon than 62Ni due to its higher fraction of lighter
protons.[32] Hence, elements heavier than iron require a supernova for their formation,
involving rapid neutron capture by starting 56Fe nuclei.[29]

In the far future of the universe, assuming that proton decay does not occur,
cold fusion occurring via quantum tunnelling would cause the light nuclei in ordinary
matter to fuse into 56Fe nuclei. Fission and alpha-particle emission would then make
heavy nuclei decay into iron, converting all stellar-mass objects to cold spheres of pure
iron.[33]

Origin and occurrence in nature


Cosmogenesis
Iron's abundance in rocky planets like Earth is due to its abundant production during the
runaway fusion and explosion of type Ia supernovae, which scatters the iron into space.
[34][35]

Metallic iron

A polished and chemically etched piece of an iron


meteorite, believed to be similar in composition to the Earth's metallic core, showing individual
crystals of the iron-nickel alloy (Widmanstatten pattern)
Metallic or native iron is rarely found on the surface of the Earth because it tends to
oxidize. However, both the Earth's inner and outer core, which together account for 35%
of the mass of the whole Earth, are believed to consist largely of an iron alloy, possibly
with nickel. Electric currents in the liquid outer core are believed to be the origin of
the Earth's magnetic field. The other terrestrial planets (Mercury, Venus, and Mars) as
well as the Moon are believed to have a metallic core consisting mostly of iron. The M-
type asteroids are also believed to be partly or mostly made of metallic iron alloy.

The rare iron meteorites are the main form of natural metallic iron on the Earth's
surface. Items made of cold-worked meteoritic iron have been found in various
archaeological sites dating from a time when iron smelting had not yet been developed;
and the Inuit in Greenland have been reported to use iron from the Cape York
meteorite for tools and hunting weapons.[36] About 1 in 20 meteorites consist of the
unique iron-nickel minerals taenite (35–80% iron) and kamacite (90–95% iron).[37] Native
iron is also rarely found in basalts that have formed from magmas that have come into
contact with carbon-rich sedimentary rocks, which have reduced the
oxygen fugacity sufficiently for iron to crystallize. This is known as telluric iron and is
described from a few localities, such as Disko Island in West Greenland, Yakutia in
Russia and Bühl in Germany.[38]

Mantle minerals
Ferropericlase (Mg,Fe)O, a solid solution of periclase (MgO) and wüstite (FeO), makes
up about 20% of the volume of the lower mantle of the Earth, which makes it the second
most abundant mineral phase in that region after silicate perovskite (Mg,Fe)SiO3; it also
is the major host for iron in the lower mantle.[39] At the bottom of the transition zone of
the mantle, the reaction γ-(Mg,Fe)2[SiO4] ↔ (Mg,Fe)[SiO3] + (Mg,Fe)O transforms γ-
olivine into a mixture of silicate perovskite and ferropericlase and vice versa. In the
literature, this mineral phase of the lower mantle is also often called magnesiowüstite.
Silicate perovskite may form up to 93% of the lower mantle,[41] and the magnesium
[40]

iron form, (Mg,Fe)SiO3, is considered to be the most abundant mineral in the Earth,
making up 38% of its volume.[42]

Earth's crust

Ochre path in Roussillon


While iron is the most abundant element on Earth, most of this iron is concentrated in
the inner and outer cores.[43][44] The fraction of iron that is in Earth's crust only amounts to
about 5% of the overall mass of the crust and is thus only the fourth most abundant
element in that layer (after oxygen, silicon, and aluminium).[45]

Most of the iron in the crust is combined with various other elements to form many iron
minerals. An important class is the iron oxide minerals such
as hematite (Fe2O3), magnetite (Fe3O4), and siderite (FeCO3), which are the major ores
of iron. Many igneous rocks also contain the sulfide minerals pyrrhotite and pentlandite.
[46][47]
During weathering, iron tends to leach from sulfide deposits as the sulfate and from
silicate deposits as the bicarbonate. Both of these are oxidized in aqueous solution and
precipitate in even mildly elevated pH as iron(III) oxide.[48]

Banded iron formation in McKinley Park, Minnesota


Large deposits of iron are banded iron formations, a type of rock consisting of repeated
thin layers of iron oxides alternating with bands of iron-poor shale and chert. The
banded iron formations were laid down in the time between 3,700 million years
ago and 1,800 million years ago.[49][50]

Materials containing finely ground iron(III) oxides or oxide-hydroxides, such as ochre,


have been used as yellow, red, and brown pigments since pre-historical times. They
contribute as well to the color of various rocks and clays, including entire geological
formations like the Painted Hills in Oregon and the Buntsandstein ("colored sandstone",
British Bunter).[51] Through Eisensandstein (a jurassic 'iron sandstone', e.g.
from Donzdorf in Germany)[52] and Bath stone in the UK, iron compounds are
responsible for the yellowish color of many historical buildings and sculptures.[53] The
proverbial red color of the surface of Mars is derived from an iron oxide-rich regolith.[54]

Significant amounts of iron occur in the iron sulfide mineral pyrite (FeS2), but it is difficult
to extract iron from it and it is therefore not exploited.[55] In fact, iron is so common that
production generally focuses only on ores with very high quantities of it.[56]

According to the International Resource Panel's Metal Stocks in Society report, the
global stock of iron in use in society is 2,200 kg per capita. More-developed countries
differ in this respect from less-developed countries (7,000–14,000 vs 2,000 kg per
capita).[57]

Oceans
Ocean science demonstrated the role of the iron in the ancient seas in both marine
biota and climate.[58]

Chemistry and compounds


Main article: Iron compounds
Oxidation
Representative compound
state

−4 (d10s2) [FeIn6−xSnx][59]

Disodium tetracarbonylferrate (Collman's


−2 (d10)
reagent)

Fe
−1 (d9)
2(CO)
2−

0 (d8) Iron pentacarbonyl


1 (d )
7
Cyclopentadienyliron dicarbonyl dimer ("Fp2")
2 (d6) Ferrous sulfate, Ferrocene
3 (d5) Ferric chloride, Ferrocenium tetrafluoroborate
Fe(diars)
4 (d )
4
2Cl
2+

2, FeO(BF4)2

5 (d3) FeO3−
4
6 (d2) Potassium ferrate
7 (d1) [FeO4]– (matrix isolation, 4K)
Iron shows the characteristic chemical properties of the transition metals, namely the
ability to form variable oxidation states differing by steps of one and a very large
coordination and organometallic chemistry: indeed, it was the discovery of an iron
compound, ferrocene, that revolutionalized the latter field in the 1950s.[60] Iron is
sometimes considered as a prototype for the entire block of transition metals, due to its
abundance and the immense role it has played in the technological progress of
humanity.[61] Its 26 electrons are arranged in the configuration [Ar]3d64s2, of which the 3d
and 4s electrons are relatively close in energy, and thus a number of electrons can be
ionized.[17]

Iron forms compounds mainly in the oxidation states +2 (iron(II), "ferrous") and +3
(iron(III), "ferric"). Iron also occurs in higher oxidation states, e.g., the purple potassium
ferrate (K2FeO4), which contains iron in its +6 oxidation state. The anion [FeO4]– with iron
in its +7 oxidation state, along with an iron(V)-peroxo isomer, has been detected by
infrared spectroscopy at 4 K after cocondensation of laser-ablated Fe atoms with a
mixture of O2/Ar.[62] Iron(IV) is a common intermediate in many biochemical oxidation
reactions.[63][64] Numerous organoiron compounds contain formal oxidation states of +1, 0,
−1, or even −2. The oxidation states and other bonding properties are often assessed
using the technique of Mössbauer spectroscopy.[65] Many mixed valence
compounds contain both iron(II) and iron(III) centers, such as magnetite and Prussian
blue (Fe4(Fe[CN]6)3).[64] The latter is used as the traditional "blue" in blueprints.[66]

Iron is the first of the transition metals that cannot reach its group oxidation state of +8,
although its heavier congeners ruthenium and osmium can, with ruthenium having more
difficulty than osmium.[11] Ruthenium exhibits an aqueous cationic chemistry in its low
oxidation states similar to that of iron, but osmium does not, favoring high oxidation
states in which it forms anionic complexes.[11] In the second half of the 3d transition
series, vertical similarities down the groups compete with the horizontal similarities of
iron with its neighbors cobalt and nickel in the periodic table, which are also
ferromagnetic at room temperature and share similar chemistry. As such, iron, cobalt,
and nickel are sometimes grouped together as the iron triad.[61]

Unlike many other metals, iron does not form amalgams with mercury. As a result,
mercury is traded in standardized 76 pound flasks (34 kg) made of iron.[67]

Iron is by far the most reactive element in its group; it is pyrophoric when finely divided
and dissolves easily in dilute acids, giving Fe2+. However, it does not react with
concentrated nitric acid and other oxidizing acids due to the formation of an impervious
oxide layer, which can nevertheless react with hydrochloric acid.[11] High-purity iron,
called electrolytic iron, is considered to be resistant to rust, due to its oxide layer.

Binary compounds
Oxides and sulfides
Ferrous or iron(II) oxide, FeO

Ferric or iron(III) oxide Fe2O3

Ferrosoferric or iron(II,III) oxide Fe3O4

Iron forms various oxide and hydroxide compounds; the most common are iron(II,III)
oxide (Fe3O4), and iron(III) oxide (Fe2O3). Iron(II) oxide also exists, though it is unstable
at room temperature. Despite their names, they are actually all non-stoichiometric
compounds whose compositions may vary.[68] These oxides are the principal ores for the
production of iron (see bloomery and blast furnace). They are also used in the
production of ferrites, useful magnetic storage media in computers, and pigments. The
best known sulfide is iron pyrite (FeS2), also known as fool's gold owing to its golden
luster.[64] It is not an iron(IV) compound, but is actually an iron(II) polysulfide containing
Fe2+ and S2−
2 ions in a distorted sodium chloride structure.
[68]

Pourbaix diagram of iron


Halides

Hydrated iron(III) chloride (ferric chloride)


The binary ferrous and ferric halides are well-known. The ferrous halides typically arise
from treating iron metal with the corresponding hydrohalic acid to give the
corresponding hydrated salts.[64]

Fe + 2 HX → FeX2 + H2 (X = F, Cl, Br, I)


Iron reacts with fluorine, chlorine, and bromine to give the corresponding ferric
halides, ferric chloride being the most common.[69]

2 Fe + 3 X2 → 2 FeX3 (X = F, Cl, Br)


Ferric iodide is an exception, being thermodynamically unstable due to the
oxidizing power of Fe3+ and the high reducing power of I−:[69]

2 I− + 2 Fe3+ → I2 + 2 Fe2+ (E0 = +0.23 V)


Ferric iodide, a black solid, is not stable in ordinary conditions, but can be
prepared through the reaction of iron pentacarbonyl with iodine and carbon
monoxide in the presence of hexane and light at the temperature of −20 °C,
with oxygen and water excluded.[69] Complexes of ferric iodide with some soft
bases are known to be stable compounds.[70][71]

Solution chemistry

Comparison of colors of solutions of ferrate (left)


and permanganate (right)
The standard reduction potentials in acidic aqueous solution for some
common iron ions are given below:[11]

[Fe(H2O)6]2+ + 2 e− ⇌ Fe E0 = −0.447 V
[Fe(H2O)6]3+ + e− ⇌ [Fe(H2O)6]2+ E0 = +0.77 V
⇌ [Fe(H2O)6]3+ + 6 H2O E0 = +2.20 V
FeO2−
4 + 8 H3O + 3 e
+ −

The red-purple tetrahedral ferrate(VI) anion is such a strong oxidizing


agent that it oxidizes ammonia to nitrogen (N2) and water to oxygen:[69]

4 FeO2−
4 + 34 H

2O → 4 [Fe(H2O)6] + 20 OH−
3+

+ 3 O2
The pale-violet hexaquo complex [Fe(H2O)6]3+ is an acid such that
above pH 0 it is fully hydrolyzed:[72]

⇌ [Fe(H2O)5(OH)]2+ + H+
[Fe(H2O)5(OH)]2+ ⇌ [Fe(H2O)4(OH)2]+ + H+
[Fe(H2O)6]3+ K = 10−3.05 mol dm−3

⇌ [Fe(H2O)4(OH)]4+2 + 2H+ + 2H2O K = 10−2.91 mol dm−3


K = 10−3.26 mol dm−3
2[Fe(H2O)6]3+

Blue-green iron(II) sulfate heptahydrate


As pH rises above 0 the above yellow hydrolyzed species form
and as it rises above 2–3, reddish-brown hydrous iron(III)
oxide precipitates out of solution. Although Fe3+ has a
d5 configuration, its absorption spectrum is not like that of
Mn2+ with its weak, spin-forbidden d–d bands, because Fe3+ has
higher positive charge and is more polarizing, lowering the energy
of its ligand-to-metal charge transfer absorptions. Thus, all the
above complexes are rather strongly colored, with the single
exception of the hexaquo ion – and even that has a spectrum
dominated by charge transfer in the near ultraviolet region.[72] On
the other hand, the pale green iron(II) hexaquo
ion [Fe(H2O)6]2+ does not undergo appreciable hydrolysis. Carbon
dioxide is not evolved when carbonate anions are added, which
instead results in white iron(II) carbonate being precipitated out. In
excess carbon dioxide this forms the slightly soluble bicarbonate,
which occurs commonly in groundwater, but it oxidises quickly in
air to form iron(III) oxide that accounts for the brown deposits
present in a sizeable number of streams.[73]

Coordination compounds
Due to its electronic structure, iron has a very large coordination
and organometallic chemistry.
The two enantiomorphs of the
ferrioxalate ion
Many coordination compounds of iron are known. A typical six-
coordinate anion is hexachloroferrate(III), [FeCl6]3−, found in the
mixed salt tetrakis(methylammonium) hexachloroferrate(III)
chloride.[74][75] Complexes with multiple bidentate ligands
have geometric isomers. For example, the trans-
chlorohydridobis(bis-1,2-(diphenylphosphino)ethane)iron(II) compl
ex is used as a starting material for compounds with
the Fe(dppe)2 moiety.[76][77] The ferrioxalate ion with
three oxalate ligands displays helical chirality with its two non-
superposable geometries labelled Λ (lambda) for the left-handed
screw axis and Δ (delta) for the right-handed screw axis, in line
with IUPAC conventions.[72] Potassium ferrioxalate is used in
chemical actinometry and along with its sodium
salt undergoes photoreduction applied in old-style photographic
processes. The dihydrate of iron(II) oxalate has
a polymeric structure with co-planar oxalate ions bridging between
iron centres with the water of crystallisation located forming the
caps of each octahedron, as illustrated below.[78]

Crystal structure of iron(II) oxalate dihydrate, showing iron (gray),


oxygen (red), carbon (black), and hydrogen (white) atoms.
Blood-red positive thiocyanate test for
iron(III)
Iron(III) complexes are quite similar to those of chromium(III) with
the exception of iron(III)'s preference for O-donor instead of N-
donor ligands. The latter tend to be rather more unstable than
iron(II) complexes and often dissociate in water. Many Fe–O
complexes show intense colors and are used as tests
for phenols or enols. For example, in the ferric chloride test, used
to determine the presence of phenols, iron(III) chloride reacts with
a phenol to form a deep violet complex:[72]

3 ArOH + FeCl3 → Fe(OAr)3 + 3 HCl (Ar = aryl)


Among the halide and pseudohalide complexes, fluoro
complexes of iron(III) are the most stable, with the colorless
[FeF5(H2O)]2− being the most stable in aqueous solution. Chloro
complexes are less stable and favor tetrahedral coordination
as in [FeCl4]−; [FeBr4]− and [FeI4]− are reduced easily to
iron(II). Thiocyanate is a common test for the presence of
iron(III) as it forms the blood-red [Fe(SCN)(H2O)5]2+. Like
manganese(II), most iron(III) complexes are high-spin, the
exceptions being those with ligands that are high in
the spectrochemical series such as cyanide. An example of a
low-spin iron(III) complex is [Fe(CN)6]3−. Iron shows a great
variety of electronic spin states, including every possible spin
quantum number value for a d-block element from 0
(diamagnetic) to 5⁄2 (5 unpaired electrons). This value is always
half the number of unpaired electrons. Complexes with zero to
two unpaired electrons are considered low-spin and those with
four or five are considered high-spin.[68]

Iron(II) complexes are less stable than iron(III) complexes but


the preference for O-donor ligands is less marked, so that for
example [Fe(NH3)6]2+ is known while [Fe(NH3)6]3+ is not. They
have a tendency to be oxidized to iron(III) but this can be
moderated by low pH and the specific ligands used.[73]
Organometallic compounds

Iron penta-
carbonyl
Organoiron chemistry is the study of organometallic
compounds of iron, where carbon atoms are covalently bound
to the metal atom. They are many and varied,
including cyanide complexes, carbonyl
complexes, sandwich and half-sandwich compounds.

Prussian blue
Prussian blue or "ferric ferrocyanide", Fe4[Fe(CN)6]3, is an old
and well-known iron-cyanide complex, extensively used as
pigment and in several other applications. Its formation can be
used as a simple wet chemistry test to distinguish between
aqueous solutions of Fe2+ and Fe3+ as they react (respectively)
with potassium ferricyanide and potassium ferrocyanide to
form Prussian blue.[64]

Another old example of an organoiron compound is iron


pentacarbonyl, Fe(CO)5, in which a neutral iron atom is bound
to the carbon atoms of five carbon monoxide molecules. The
compound can be used to make carbonyl iron powder, a highly
reactive form of metallic iron. Thermolysis of iron
pentacarbonyl gives triiron dodecacarbonyl, Fe3(CO)12, a
complex with a cluster of three iron atoms at its core.
Collman's reagent, disodium tetracarbonylferrate, is a useful
reagent for organic chemistry; it contains iron in the −2
oxidation state. Cyclopentadienyliron dicarbonyl
dimer contains iron in the rare +1 oxidation state.[79]
Structural formula of ferrocene and a powdered sample

A landmark in this field was the discovery in 1951 of the


remarkably stable sandwich compound ferrocene Fe(C5H5)2, by
Pauson and Kealy[80] and independently by Miller and
colleagues,[81] whose surprising molecular structure was
determined only a year later
by Woodward and Wilkinson[82] and Fischer.[83] Ferrocene is still
one of the most important tools and models in this class.[84]

Iron-centered organometallic species are used as catalysts.


The Knölker complex, for example, is a transfer
hydrogenation catalyst for ketones.[85]

Industrial uses
The iron compounds produced on the largest scale in industry
are iron(II) sulfate (FeSO4·7H2O) and iron(III) chloride (FeCl3).
The former is one of the most readily available sources of
iron(II), but is less stable to aerial oxidation than Mohr's
salt ((NH4)2Fe(SO4)2·6H2O). Iron(II) compounds tend to be
oxidized to iron(III) compounds in the air.[64]

History
Main article: History of ferrous metallurgy
Development of iron metallurgy
Iron is one of the elements undoubtedly known to the ancient
world.[86] It has been worked, or wrought, for millennia.
However, iron artefacts of great age are much rarer than
objects made of gold or silver due to the ease with which iron
corrodes.[87] The technology developed slowly, and even after
the discovery of smelting it took many centuries for iron to
replace bronze as the metal of choice for tools and weapons.
Meteoritic iron

Iron harpoon head


from Greenland. The iron edge covers a narwhal tusk harpoon using
meteorite iron from the Cape York meteorite, one of the largest iron
meteorites known.
Beads made from meteoric iron in 3500 BC or earlier were
found in Gerzeh, Egypt by G. A. Wainwright.[88] The beads
contain 7.5% nickel, which is a signature of meteoric origin
since iron found in the Earth's crust generally has only
minuscule nickel impurities.

Meteoric iron was highly regarded due to its origin in the


heavens and was often used to forge weapons and tools.
[88]
For example, a dagger made of meteoric iron was found in
the tomb of Tutankhamun, containing similar proportions of
iron, cobalt, and nickel to a meteorite discovered in the area,
deposited by an ancient meteor shower.[89][90][91] Items that were
likely made of iron by Egyptians date from 3000 to 2500 BC.[87]

Meteoritic iron is comparably soft and ductile and easily cold


forged but may get brittle when heated because of
the nickel content.[92]

Wrought iron
Main article: Wrought iron
Further information: Ancient iron production
The symbol for Mars has been used since

antiquity to represent iron.


The iron pillar of Delhi is an example of the iron extraction and
processing methodologies of early India.
The first iron production started in the Middle Bronze Age, but
it took several centuries before iron displaced bronze.
Samples of smelted iron from Asmar, Mesopotamia and Tall
Chagar Bazaar in northern Syria were made sometime
between 3000 and 2700 BC.[93] The Hittites established an
empire in north-central Anatolia around 1600 BC. They appear
to be the first to understand the production of iron from its ores
and regard it highly in their society.[94] The Hittites began to
smelt iron between 1500 and 1200 BC and the practice spread
to the rest of the Near East after their empire fell in 1180 BC.
[93]
The subsequent period is called the Iron Age.

Artifacts of smelted iron are found in India dating from 1800 to


1200 BC,[95] and in the Levant from about 1500 BC (suggesting
smelting in Anatolia or the Caucasus).[96][97] Alleged references
(compare history of metallurgy in South Asia) to iron in the
Indian Vedas have been used for claims of a very early usage
of iron in India respectively to date the texts as such.
The rigveda term ayas (metal) refers to copper, while iron
which is called as śyāma ayas, literally "black copper", first is
mentioned in the post-rigvedic Atharvaveda.[98]
Some archaeological evidence suggests iron was smelted
in Zimbabwe and southeast Africa as early as the eighth
century BC.[99] Iron working was introduced to Greece in the
late 11th century BC, from which it spread quickly throughout
Europe.[100]

Iron sickle from Ancient Greece


The spread of ironworking in Central and Western Europe is
associated with Celtic expansion. According to Pliny the Elder,
iron use was common in the Roman era.[88] In the lands of what
is now considered China, iron appears approximately 700–
500 BC.[101] Iron smelting may have been introduced into China
through Central Asia.[102] The earliest evidence of the use of
a blast furnace in China dates to the 1st century AD,[103] and
cupola furnaces were used as early as the Warring States
period (403–221 BC).[104] Usage of the blast and cupola furnace
remained widespread during the Tang and Song dynasties.[105]

During the Industrial Revolution in Britain, Henry Cort began


refining iron from pig iron to wrought iron (or bar iron) using
innovative production systems. In 1783 he patented
the puddling process for refining iron ore. It was later improved
by others, including Joseph Hall.[106]

Cast iron
Main article: Cast iron
Cast iron was first produced in China during 5th century BC,
[107]
but was hardly in Europe until the medieval period.[108][109] The
earliest cast iron artifacts 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.[110] During the medieval period, means were found
in Europe of producing wrought iron from cast iron (in this
context known as pig iron) using finery forges. For all these
processes, charcoal was required as fuel.[111]

Coalbrookdale by Night, 1801.


Blast furnaces light the iron making town of Coalbrookdale.
Medieval blast furnaces were about 10 feet (3.0 m) tall and
made of fireproof brick; forced air was usually provided by
hand-operated bellows.[109] Modern blast furnaces have grown
much bigger, with hearths fourteen meters in diameter that
allow them to produce thousands of tons of iron each day, but
essentially operate in much the same way as they did during
medieval times.[111]

In 1709, Abraham Darby I established a coke-fired blast


furnace to produce cast iron, replacing charcoal, although
continuing to use blast furnaces. The ensuing availability of
inexpensive iron was one of the factors leading to
the Industrial Revolution. Toward the end of the 18th century,
cast iron began to replace wrought iron for certain purposes,
because it was cheaper. Carbon content in iron was not
implicated as the reason for the differences in properties of
wrought iron, cast iron, and steel until the 18th century.[93]

Since iron was becoming cheaper and more plentiful, it also


became a major structural material following the building of the
innovative first iron bridge in 1778. This bridge still stands
today as a monument to the role iron played in the Industrial
Revolution. Following this, iron was used in rails, boats, ships,
aqueducts, and buildings, as well as in iron cylinders in steam
engines.[111] Railways have been central to the formation of
modernity and ideas of progress[112] and various languages
refer to railways as iron road (e.g. French chemin de fer,
German Eisenbahn, Turkish demiryolu, Russian железная
дорога, Chinese, Japanese, and Korean 鐵道,
Vietnamese đường sắt).

Steel
Main article: Steel
See also: Steelmaking
Steel (with smaller carbon content than pig iron but more than
wrought iron) was first produced in antiquity by using
a bloomery. Blacksmiths in Luristan in western Persia were
making good steel by 1000 BC.[93] Then improved
versions, Wootz steel by India and Damascus steel were
developed around 300 BC and AD 500 respectively. These
methods were specialized, and so steel did not become a
major commodity until the 1850s.[113]

New methods of producing it by carburizing bars of iron in


the cementation process were devised in the 17th century. In
the Industrial Revolution, new methods of producing bar iron
without charcoal were devised and these were later applied to
produce steel. In the late 1850s, Henry Bessemer invented a
new steelmaking process, involving blowing air through molten
pig iron, to produce mild steel. This made steel much more
economical, thereby leading to wrought iron no longer being
produced in large quantities.[114]

Foundations of modern chemistry


In 1774, Antoine Lavoisier used the reaction of water steam
with metallic iron inside an incandescent iron tube to
produce hydrogen in his experiments leading to the
demonstration of the conservation of mass, which was
instrumental in changing chemistry from a qualitative science
to a quantitative one.[115]

Symbolic role

"Ich gab Gold für Eisen" – "I gave


gold for iron". German-American brooch from WWI.
Iron plays a certain role in mythology and has found various
usage as a metaphor and in folklore.
The Greek poet Hesiod's Works and Days (lines 109–201)
lists different ages of man named after metals like gold, silver,
bronze and iron to account for successive ages of humanity.
[116]
The Iron Age was closely related with Rome, and in
Ovid's Metamorphoses

The Virtues, in despair, quit the earth; and the depravity of


man becomes universal and complete. Hard steel succeeded
then.

— Ovid, Metamorphoses, Book I, Iron age, line 160 ff


An example of the importance of iron's symbolic role may be
found in the German Campaign of 1813. Frederick William
III commissioned then the first Iron Cross as military
decoration. Berlin iron jewellery reached its peak production
between 1813 and 1815, when the Prussian royal family urged
citizens to donate gold and silver jewellery for military funding.
The inscription Ich gab Gold für Eisen (I gave gold for iron)
was used as well in later war efforts.[117]

Production of metallic iron

Iron powder
Laboratory routes
For a few limited purposes when it is needed, pure iron is
produced in the laboratory in small quantities by reducing the
pure oxide or hydroxide with hydrogen, or forming iron
pentacarbonyl and heating it to 250 °C so that it decomposes
to form pure iron powder.[48] Another method is electrolysis of
ferrous chloride onto an iron cathode.[118]

Main industrial route


See also: Iron ore
Iron production 2009 (million tonnes)[119][dubious – discuss]

Country Iron ore Pig iron Direct iron Steel

China 1,114.9 549.4 573.6

Australia 393.9 4.4 5.2

Brazil 305.0 25.1 0.011 26.5

Japan 66.9 87.5

India 257.4 38.2 23.4 63.5

Russia 92.1 43.9 4.7 60.0

Ukraine 65.8 25.7 29.9


South 0.1 27.3 48.6
Korea

German 0.4 20.1 0.38 32.7


y

World 1,594.9 914.0 64.5 1,232.4

Nowadays, the industrial production of iron or steel consists of


two main stages. In the first stage, iron ore
is reduced with coke in a blast furnace, and the molten metal
is separated from gross impurities such as silicate minerals.
This stage yields an alloy – pig iron – that contains relatively
large amounts of carbon. In the second stage, the amount of
carbon in the pig iron is lowered by oxidation to yield wrought
iron, steel, or cast iron.[120] Other metals can be added at this
stage to form alloy steels.

Blast furnace processing


Main article: Blast furnace
The blast furnace is loaded with iron ores,
usually hematite Fe2O3 or magnetite Fe3O4, along with coke
(coal that has been separately baked to remove volatile
components) and flux (limestone or dolomite). "Blasts" of air
pre-heated to 900 °C (sometimes with oxygen enrichment) is
blown through the mixture, in sufficient amount to turn the
carbon into carbon monoxide:[120]

2 C + O2 → 2 CO
This reaction raises the temperature to about 2000 °C. The
carbon monoxide reduces the iron ore to metallic iron:[120]

Fe2O3 + 3 CO → 2 Fe + 3 CO2
Some iron in the high-temperature lower region of the
furnace reacts directly with the coke:[120]

2 Fe2O3 + 3 C → 4 Fe + 3 CO2
The flux removes silicaceous minerals in the ore,
which would otherwise clog the furnace: The heat
of the furnace decomposes the carbonates
to calcium oxide, which reacts with any
excess silica to form a slag composed of calcium
silicate CaSiO3 or other products. At the furnace's
temperature, the metal and the slag are both
molten. They collect at the bottom as two
immiscible liquid layers (with the slag on top), that
are then easily separated.[120] The slag can be used
as a material in road construction or to improve
mineral-poor soils for agriculture.[109]

Steelmaking thus remains one of the largest


industrial contributors of CO2 emissions in the
world.[121]

17th century Chinese illustration of workers at a blast


[122]
furnace, making wrought iron from pig iron

How iron was extracted in the 19th century

Iron furnace in Columbus, Ohio, 1922


Steelmaking
Main articles: Steelmaking and Ironworks
The pig iron produced by the blast furnace process
contains up to 4–5% carbon (by mass), with small
amounts of other impurities like sulfur, magnesium,
phosphorus, and manganese. This high level of
carbon makes it relatively weak and brittle.
Reducing the amount of carbon to 0.002–2.1%
produces steel, which may be up to 1000 times
harder than pure iron. A great variety of steel
articles can then be made by cold working, hot
rolling, forging, machining, etc. Removing the
impurities from pig iron, but leaving 2–4% carbon,
results in cast iron, which is cast by foundries into
articles such as stoves, pipes, radiators, lamp-
posts, and rails.[120]

Steel products often undergo various heat


treatments after they are forged to
shape. Annealing consists of heating them to 700–
800 °C for several hours and then gradual cooling.
It makes the steel softer and more workable.[123]

This heap of iron ore pellets will be used in steel


production.

A pot of molten iron being used to make steel


Direct iron reduction
Owing to environmental concerns, alternative
methods of processing iron have been developed.
"Direct iron reduction" reduces iron ore to a ferrous
lump called "sponge" iron or "direct" iron that is
suitable for steelmaking.[109] Two main reactions
comprise the direct reduction process:

Natural gas is partially oxidized (with heat and a


catalyst):[109]

2 CH4 + O2 → 2 CO + 4 H2
Iron ore is then treated with these gases in a
furnace, producing solid sponge iron:[109]

Fe2O3 + CO + 2 H2 → 2 Fe + CO2 + 2 H2O


Silica is removed by adding a limestone flux
as described above.[109]

Thermite process
Main article: Thermite
Ignition of a mixture of aluminium powder
and iron oxide yields metallic iron via
the thermite reaction:

Fe2O3 + 2 Al → 2 Fe + Al2O3
Alternatively pig iron may be made into
steel (with up to about 2% carbon) or
wrought iron (commercially pure iron).
Various processes have been used for
this, including finery
forges, puddling furnaces, Bessemer
converters, open hearth furnaces, basic
oxygen furnaces, and electric arc
furnaces. In all cases, the objective is to
oxidize some or all of the carbon,
together with other impurities. On the
other hand, other metals may be added
to make alloy steels.[111]

Molten oxide electrolysis


Molten oxide electrolysis (MOE)
uses electrolysis of molten iron oxide to
yield metallic iron. It is studied in
laboratory-scale experiments and is
proposed as a method for industrial iron
production that has no direct emissions
of carbon dioxide. It uses a liquid iron
cathode, an anode formed from an alloy
of chromium, aluminium and iron,[124] and
the electrolyte is a mixture of molten
metal oxides into which iron ore is
dissolved. The current keeps the
electrolyte molten and reduces the iron
oxide. Oxygen gas is produced in
addition to liquid iron. The only carbon
dioxide emissions come from any fossil
fuel-generated electricity used to heat
and reduce the metal.[125][126][127]

Applications
Characteristic values of tensile
strength (TS) and Brinell
hardness (BH) of various forms of
iron.[128][129]

TS BH
Material
(MPa) (Brinell)

Iron whiskers 11000

Ausformed
(hardened) 2930 850–1200
steel

Martensitic steel 2070 600

Bainitic steel 1380 400

Pearlitic steel 1200 350

Cold-worked iron 690 200

Small-grain iron 340 100

Carbon-
140 40
containing iron

Pure, single-
10 3
crystal iron
As structural material
Iron is the most widely used of all the
metals, accounting for over 90% of
worldwide metal production. Its low cost
and high strength often make it the
material of choice to withstand stress or
transmit forces, such as the construction
of machinery and machine
tools, rails, automobiles, ship
hulls, concrete reinforcing bars, and the
load-carrying framework of buildings.
Since pure iron is quite soft, it is most
commonly combined with alloying
elements to make steel.[130]

Mechanical properties
The mechanical properties of iron and its
alloys are extremely relevant to their
structural applications. Those properties
can be evaluated in various ways,
including the Brinell test, the Rockwell
test and the Vickers hardness test.

The properties of pure iron are often


used to calibrate measurements or to
compare tests.[129][131] However, the
mechanical properties of iron are
significantly affected by the sample's
purity: pure, single crystals of iron are
actually softer than aluminium,[128] and the
purest industrially produced iron
(99.99%) has a hardness of 20–
30 Brinell.[132] The pure iron (99.9%~
99.999%), especially called electrolytic
iron, is industrially produced
by electrolytic refining.

An increase in the carbon content will


cause a significant increase in the
hardness and tensile strength of iron.
Maximum hardness of 65 Rc is achieved
with a 0.6% carbon content, although the
alloy has low tensile strength.
[133]
Because of the softness of iron, it is
much easier to work with than its
heavier congeners ruthenium and osmiu
m.[17]

Types of steels and alloys


See also: Steel
Iron-carbon phase diagram
α-Iron is a fairly soft metal that can
dissolve only a small concentration of
carbon (no more than 0.021% by mass
at 910 °C).[134] Austenite (γ-iron) is
similarly soft and metallic but can
dissolve considerably more carbon (as
much as 2.04% by mass at 1146 °C).
This form of iron is used in the type
of stainless steel used for making
cutlery, and hospital and food-service
equipment.[21]

Commercially available iron is classified


based on purity and the abundance of
additives. Pig iron has 3.5–4.5%
carbon[135] and contains varying amounts
of contaminants such as sulfur, silicon
and phosphorus. Pig iron is not a
saleable product, but rather an
intermediate step in the production of
cast iron and steel. The reduction of
contaminants in pig iron that negatively
affect material properties, such as sulfur
and phosphorus, yields cast iron
containing 2–4% carbon, 1–6% silicon,
and small amounts of manganese.[120] Pig
iron has a melting point in the range of
1420–1470 K, which is lower than either
of its two main components, and makes
it the first product to be melted when
carbon and iron are heated together.
[11]
Its mechanical properties vary greatly
and depend on the form the carbon
takes in the alloy.[17]

"White" cast irons contain their carbon in


the form of cementite, or iron carbide
(Fe3C).[17] This hard, brittle compound
dominates the mechanical properties of
white cast irons, rendering them hard,
but unresistant to shock. The broken
surface of a white cast iron is full of fine
facets of the broken iron carbide, a very
pale, silvery, shiny material, hence the
appellation. Cooling a mixture of iron
with 0.8% carbon slowly below 723 °C to
room temperature results in separate,
alternating layers of cementite and α-
iron, which is soft and malleable and is
called pearlite for its appearance. Rapid
cooling, on the other hand, does not
allow time for this separation and creates
hard and brittle martensite. The steel can
then be tempered by reheating to a
temperature in between, changing the
proportions of pearlite and martensite.
The end product below 0.8% carbon
content is a pearlite-αFe mixture, and
that above 0.8% carbon content is a
pearlite-cementite mixture.[17]

In gray iron the carbon exists as


separate, fine flakes of graphite, and
also renders the material brittle due to
the sharp edged flakes of graphite that
produce stress concentration sites within
the material.[136] A newer variant of gray
iron, referred to as ductile iron, is
specially treated with trace amounts
of magnesium to alter the shape of
graphite to spheroids, or nodules,
reducing the stress concentrations and
vastly increasing the toughness and
strength of the material.[136]

Wrought iron contains less than 0.25%


carbon but large amounts of slag that
give it a fibrous characteristic.
[135]
Wrought iron is more corrosion
resistant than steel. It has been almost
completely replaced by mild steel, which
corrodes more readily than wrought iron,
but is cheaper and more widely
available. Carbon steel contains 2.0%
carbon or less,[137] with small amounts
of manganese, sulfur, phosphorus, and
silicon. Alloy steels contain varying
amounts of carbon as well as other
metals, such
as chromium, vanadium, molybdenum,
nickel, tungsten, etc. Their alloy content
raises their cost, and so they are usually
only employed for specialist uses. One
common alloy steel, though, is stainless
steel. Recent developments in ferrous
metallurgy have produced a growing
range of microalloyed steels, also
termed 'HSLA' or high-strength, low alloy
steels, containing tiny additions to
produce high strengths and often
spectacular toughness at minimal cost.
[137][138][139]

Alloys with high purity elemental


makeups (such as alloys of electrolytic
iron) have specifically enhanced
properties such as ductility, tensile
strength, toughness, fatigue strength,
heat resistance, and corrosion
resistance.

Apart from traditional applications, iron is


also used for protection from ionizing
radiation. Although it is lighter than
another traditional protection
material, lead, it is much stronger
mechanically.[140]

The main disadvantage of iron and steel


is that pure iron, and most of its alloys,
suffer badly from rust if not protected in
some way, a cost amounting to over 1%
of the world's economy.
[141]
Painting, galvanization, passivation,
plastic coating and bluing are all used to
protect iron from rust by
excluding water and oxygen or
by cathodic protection. The mechanism
of the rusting of iron is as follows:[141]

Cathode: 3 O2 + 6 H2O + 12 e− → 12 OH−


Anode: 4 Fe → 4 Fe2+ + 8 e−; 4 Fe2+ → 4 Fe3+ + 4 e−
Overall: 4 Fe + 3 O2 + 6 H2O → 4 Fe3+ + 12 OH− → 4 Fe(OH)3 or 4 FeO(OH) + 4
H2O
The electrolyte is
usually iron(II) sulfate in
urban areas (formed when
atmospheric sulfur
dioxide attacks iron), and salt
particles in the atmosphere in
seaside areas.[141]

Catalysts and
reagents
Because Fe is inexpensive
and nontoxic, much effort has
been devoted to the
development of Fe-based
catalysts and reagents. Iron is
however less common as a
catalyst in commercial
processes than more
expensive metals.[142] In
biology, Fe-containing
enzymes are pervasive.[143]

Iron catalysts are traditionally


used in the Haber–Bosch
process for the production of
ammonia and the Fischer–
Tropsch process for
conversion of carbon
monoxide to hydrocarbons for
fuels and lubricants.
[144]
Powdered iron in an acidic
medium is used in
the Bechamp reduction, the
conversion
of nitrobenzene to aniline.[145]
Iron compounds
Iron(III) oxide mixed
with aluminium powder can
be ignited to create a thermite
reaction, used in welding
large iron parts (like rails) and
purifying ores. Iron(III) oxide
and oxyhydroxide are used as
reddish and ocher pigments.

Iron(III) chloride finds use in


water purification and sewage
treatment, in the dyeing of
cloth, as a coloring agent in
paints, as an additive in
animal feed, and as
an etchant for copper in the
manufacture of printed circuit
boards.[146] It can also be
dissolved in alcohol to form
tincture of iron, which is used
as a medicine to stop
bleeding in canaries.[147]

Iron(II) sulfate is used as a


precursor to other iron
compounds. It is also used
to reduce chromate in
cement. It is used to fortify
foods and treat iron deficiency
anemia. Iron(III) sulfate is
used in settling minute
sewage particles in tank
water. Iron(II) chloride is used
as a reducing flocculating
agent, in the formation of iron
complexes and magnetic iron
oxides, and as a reducing
agent in organic synthesis.[146]

Sodium nitroprusside is a
drug used as a vasodilator. It
is on the World Health
Organization's List of
Essential Medicines.[148]
Biological and
pathological
role
Main article: Iron in biology
Iron is required for life.[10][149]
[150]
The iron–sulfur
clusters are pervasive and
include nitrogenase, the
enzymes responsible for
biological nitrogen fixation.
Iron-containing proteins
participate in transport,
storage and use of oxygen.
[10]
Iron proteins are involved
in electron transfer.[151]

Simplified structure of Heme B;


in the protein additional ligand(s)
are attached to Fe.
Examples of iron-containing
proteins in higher organisms
include
hemoglobin, cytochrome (see
high-valent iron),
and catalase.[10][152] The
average adult human contains
about 0.005% body weight of
iron, or about four grams, of
which three quarters is in
hemoglobin—a level that
remains constant despite only
about one milligram of iron
being absorbed each day,
[151]
because the human body
recycles its hemoglobin for
the iron content.[153]

Microbial growth may be


assisted by oxidation of
iron(II) or by reduction of
iron(III).[154]

Biochemistry
Iron acquisition poses a
problem for aerobic
organisms because ferric iron
is poorly soluble near neutral
pH. Thus, these organisms
have developed means to
absorb iron as complexes,
sometimes taking up ferrous
iron before oxidising it back to
ferric iron.[10] In particular,
bacteria have evolved very
high-affinity sequestering age
nts called siderophores.[155][156]
[157]

After uptake in human cells,


iron storage is precisely
regulated.[10][158] A major
component of this regulation
is the protein transferrin,
which binds iron ions
absorbed from
the duodenum and carries it
in the blood to cells.[10]
[159]
Transferrin contains Fe3+ in
the middle of a distorted
octahedron, bonded to one
nitrogen, three oxygens and a
chelating carbonate anion
that traps the Fe3+ ion: it has
such a high stability
constant that it is very
effective at taking up Fe3+ ions
even from the most stable
complexes. At the bone
marrow, transferrin is reduced
from Fe3+ to Fe2+ and stored
as ferritin to be incorporated
into hemoglobin.[151]

The most commonly known


and studied bioinorganic iron
compounds (biological iron
molecules) are the heme
proteins: examples
are hemoglobin, myoglobin,
and cytochrome P450.
[10]
These compounds
participate in transporting
gases, building enzymes, and
transferring electrons.[151] Meta
lloproteins are a group of
proteins with metal
ion cofactors. Some examples
of iron metalloproteins
are ferritin and rubredoxin.[151]
Many enzymes vital to life
contain iron, such
as catalase,[160] lipoxygenases,
[161]
and IRE-BP.[162]

Hemoglobin is an oxygen
carrier that occurs in red
blood cells and contributes
their color, transporting
oxygen in the arteries from
the lungs to the muscles
where it is transferred
to myoglobin, which stores it
until it is needed for the
metabolic oxidation
of glucose, generating
energy.[10] Here the
hemoglobin binds to carbon
dioxide, produced when
glucose is oxidized, which is
transported through the veins
by hemoglobin (predominantly
as bicarbonate anions) back
to the lungs where it is
exhaled.[151] In hemoglobin, the
iron is in one of
four heme groups and has six
possible coordination sites;
four are occupied by nitrogen
atoms in a porphyrin ring, the
fifth by an imidazole nitrogen
in a histidine residue of one of
the protein chains attached to
the heme group, and the sixth
is reserved for the oxygen
molecule it can reversibly bind
to.[151] When hemoglobin is not
attached to oxygen (and is
then called
deoxyhemoglobin), the
Fe2+ ion at the center of
the heme group (in the
hydrophobic protein interior)
is in a high-spin configuration.
It is thus too large to fit inside
the porphyrin ring, which
bends instead into a dome
with the Fe2+ ion about
55 picometers above it. In this
configuration, the sixth
coordination site reserved for
the oxygen is blocked by
another histidine residue.[151]

When deoxyhemoglobin picks


up an oxygen molecule, this
histidine residue moves away
and returns once the oxygen
is securely attached to form
a hydrogen bond with it. This
results in the Fe2+ ion
switching to a low-spin
configuration, resulting in a
20% decrease in ionic radius
so that now it can fit into the
porphyrin ring, which
becomes planar.
[151]
Additionally, this hydrogen
bonding results in the tilting of
the oxygen molecule,
resulting in a Fe–O–O bond
angle of around 120° that
avoids the formation of Fe–
O–Fe or Fe–O2–Fe bridges
that would lead to electron
transfer, the oxidation of
Fe2+ to Fe3+, and the
destruction of hemoglobin.
This results in a movement of
all the protein chains that
leads to the other subunits of
hemoglobin changing shape
to a form with larger oxygen
affinity. Thus, when
deoxyhemoglobin takes up
oxygen, its affinity for more
oxygen increases, and vice
versa.[151] Myoglobin, on the
other hand, contains only one
heme group and hence this
cooperative effect cannot
occur. Thus, while
hemoglobin is almost
saturated with oxygen in the
high partial pressures of
oxygen found in the lungs, its
affinity for oxygen is much
lower than that of myoglobin,
which oxygenates even at low
partial pressures of oxygen
found in muscle tissue.[151] As
described by the Bohr
effect (named after Christian
Bohr, the father of Niels
Bohr), the oxygen affinity of
hemoglobin diminishes in the
presence of carbon dioxide.[151]
A heme unit of
human carboxyhemoglobin,
showing the carbonyl ligand at
the apical position, trans to the
histidine residue[163]
Carbon
monoxide and phosphorus
trifluoride are poisonous to
humans because they bind to
hemoglobin similarly to
oxygen, but with much more
strength, so that oxygen can
no longer be transported
throughout the body.
Hemoglobin bound to carbon
monoxide is known
as carboxyhemoglobin. This
effect also plays a minor role
in the toxicity of cyanide, but
there the major effect is by far
its interference with the
proper functioning of the
electron transport
protein cytochrome a.[151] The
cytochrome proteins also
involve heme groups and are
involved in the metabolic
oxidation of glucose by
oxygen. The sixth
coordination site is then
occupied by either another
imidazole nitrogen or
a methionine sulfur, so that
these proteins are largely
inert to oxygen—with the
exception of cytochrome a,
which bonds directly to
oxygen and thus is very easily
poisoned by cyanide.[151] Here,
the electron transfer takes
place as the iron remains in
low spin but changes between
the +2 and +3 oxidation
states. Since the reduction
potential of each step is
slightly greater than the
previous one, the energy is
released step-by-step and
can thus be stored
in adenosine triphosphate.
Cytochrome a is slightly
distinct, as it occurs at the
mitochondrial membrane,
binds directly to oxygen, and
transports protons as well as
electrons, as follows:[151]

4 Cytc2+ + O2 + 8H+
inside → 4 Cytc + 2 H2O + 4H+
3+

outside

Although the heme


proteins are the most
important class of iron-
containing proteins,
the iron–sulfur
proteins are also very
important, being involved
in electron transfer, which
is possible since iron can
exist stably in either the
+2 or +3 oxidation states.
These have one, two,
four, or eight iron atoms
that are each
approximately
tetrahedrally coordinated
to four sulfur atoms;
because of this tetrahedral
coordination, they always
have high-spin iron. The
simplest of such
compounds is rubredoxin,
which has only one iron
atom coordinated to four
sulfur atoms
from cysteine residues in
the surrounding peptide
chains. Another important
class of iron–sulfur
proteins is the ferredoxins,
which have multiple iron
atoms. Transferrin does
not belong to either of
these classes.[151]

The ability of
sea mussels to maintain
their grip on rocks in the
ocean is facilitated by their
use of organometallic iron-
based bonds in their
protein-rich cuticles.
Based on synthetic
replicas, the presence of
iron in these structures
increased elastic
modulus 770
times, tensile strength 58
times, and toughness 92
times. The amount of
stress required to
permanently damage
them increased 76 times.
[164]

You might also like