Sie sind auf Seite 1von 131

Iodine

Iodine is a chemical element with symbol I


and atomic number 53. The heaviest of the
stable halogens, it exists as a lustrous,
black non-metallic solid at standard
conditions that melts to form a deep violet
almost black liquid at 114 degrees Celsius,
and boils to a violet gas at 184 degrees
Celsius. The element was discovered by
the French chemist Bernard Courtois in
1811. It was named two years later by
Joseph Louis Gay-Lussac from this
property, after the Greek ἰώδης "violet-
coloured".
Iodine,  53I

Iodine

Pronunciation /ˈaɪədaɪn, -dɪn, -diːn/


(EYE-ə-dyn, -din, -deen)

Appearance lustrous metallic gray,


violet as a gas

Standard atomic weight 126.904 47(3)[1]


Ar, std(I)

Iodine in the periodic table


Br

I

At
tellurium ← iodine → xenon

Atomic number (Z) 53

Group group 17 (halogens)


Period period 5
Block p-block

Element category reactive nonmetal


 
Electron configuration [Kr] 4d10 5s2 5p5

Electrons per shell 2, 8, 18, 18, 7

Physical properties

Phase at STP solid

Melting point 386.85 K (113.7 °C,


236.66 °F)

Boiling point 457.4 K (184.3 °C,


363.7 °F)

Density (near r.t.) 4.933 g/cm3

Triple point 386.65 K, 12.1 kPa

Critical point 819 K, 11.7 MPa

Heat of fusion (I2) 15.52 kJ/mol


Heat of vaporisation (I2) 41.57 kJ/mol
Molar heat capacity (I2) 54.44 J/(mol·K)

Vapour pressure (rhombic)
P (Pa) 1 10 100 1 k 10 k 100 k

at T (K) 260 282 309 342 381 457

Atomic properties

Oxidation states −1, +1, +3, +4, +5, +6,


+7 (a strongly acidic
oxide)

Electronegativity Pauling scale: 2.66

Ionisation energies 1st: 1008.4 kJ/mol


2nd: 1845.9 kJ/mol
3rd: 3180 kJ/mol

Atomic radius empirical: 140 pm

Covalent radius 139±3 pm

Van der Waals radius 198 pm


Spectral lines of iodine

Other properties

Natural occurrence primordial

Crystal structure orthorhombic

Thermal conductivity 0.449 W/(m·K)

Electrical resistivity 1.3×107 Ω·m (at 0 °C)

Magnetic ordering diamagnetic[2]

Magnetic susceptibility −88.7·10−6 cm3/mol


(298 K)[3]

Bulk modulus 7.7 GPa

CAS Number 7553-56-2

History

Discovery and first Bernard Courtois (1811)


isolation
Main isotopes of iodine
Isotope Abundance Half-life (t1/2) Decay mode Product

123I syn 13 h ε, γ 123Te

124 124
I syn 4.176 d ε Te

125 125
I syn 59.40 d ε Te

127I 100% stable

129I trace 1.57×107 y β− 129Xe

131
I syn 8.02070 d β−, γ 131
Xe

135
I syn 6.57 h β− 135
Xe

Iodine occurs in many oxidation states,


including iodide (I−), iodate (IO−3), and the
various periodate anions. It is the least
abundant of the stable halogens, being the
sixty-first most abundant element. It is the
heaviest essential mineral nutrient. Iodine
is essential in the synthesis of thyroid
hormones.[4] Iodine deficiency affects
about two billion people and is the leading
preventable cause of intellectual
disabilities.

The dominant producers of iodine today


are Chile and Japan. Iodine and its
compounds are primarily used in nutrition.
Due to its high atomic number and ease of
attachment to organic compounds, it has
also found favour as a non-toxic
radiocontrast material. Because of the
specificity of its uptake by the human
body, radioactive isotopes of iodine can
also be used to treat thyroid cancer. Iodine
is also used as a catalyst in the industrial
production of acetic acid and some
polymers.

History
In 1811, iodine was discovered by French
chemist Bernard Courtois,[5][6] who was
born to a manufacturer of saltpetre (an
essential component of gunpowder). At
the time of the Napoleonic Wars, saltpetre
was in great demand in France. Saltpetre
produced from French nitre beds required
sodium carbonate, which could be isolated
from seaweed collected on the coasts of
Normandy and Brittany. To isolate the
sodium carbonate, seaweed was burned
and the ash washed with water. The
remaining waste was destroyed by adding
sulfuric acid. Courtois once added
excessive sulfuric acid and a cloud of
purple vapour rose. He noted that the
vapour crystallised on cold surfaces,
making dark crystals.[7] Courtois
suspected that this material was a new
element but lacked funding to pursue it
further.[8]

Courtois gave samples to his friends,


Charles Bernard Desormes (1777–1838)
and Nicolas Clément (1779–1841), to
continue research. He also gave some of
the substance to chemist Joseph Louis
Gay-Lussac (1778–1850), and to physicist
André-Marie Ampère (1775–1836). On 29
November 1813, Desormes and Clément
made Courtois' discovery public. They
described the substance to a meeting of
the Imperial Institute of France.[9] On 6
December, Gay-Lussac announced that the
new substance was either an element or a
compound of oxygen.[10][11][12] It was Gay-
Lussac who suggested the name "iode",
from the Greek word ἰοειδής[13] (ioeidēs)
for violet (because of the colour of iodine
vapor).[5][10] Ampère had given some of his
sample to English chemist Humphry Davy
(1778–1829), who experimented on the
substance and noted its similarity to
chlorine.[14] Davy sent a letter dated 10
December to the Royal Society of London
stating that he had identified a new
element.[15] Arguments erupted between
Davy and Gay-Lussac over who identified
iodine first, but both scientists
acknowledged Courtois as the first to
isolate the element.[8]

In early periodic tables, iodine is often


given the symbol J, for jod, its name in
German.[16]

Properties
Violet iodine vapour in a round-bottomed flask

Iodine is the fourth halogen, being a


member of group 17 in the periodic table,
below fluorine, chlorine, and bromine; it is
the heaviest stable member of its group.
(The scarce and fugitive fifth halogen, the
radioactive astatine, is not well-studied
due to its expense and inaccessibility in
large quantities, but appears to show
various unusual properties due to
relativistic effects.) Iodine has an electron
configuration of [Kr]4d105s25p5, with the
seven electrons in the fifth and outermost
shell being its valence electrons. Like the
other halogens, it is one electron short of a
full octet and is hence a strong oxidising
agent, reacting with many elements in
order to complete its outer shell, although
in keeping with periodic trends, it is the
weakest oxidising agent among the stable
halogens: it has the lowest
electronegativity among them, just 2.66 on
the Pauling scale (compare fluorine,
chlorine, and bromine at 3.98, 3.16, and
2.96 respectively; astatine continues the
trend with an electronegativity of 2.2).
Elemental iodine hence forms diatomic
molecules with chemical formula I2, where
two iodine atoms share a pair of electrons
in order to each achieve a stable octet for
themselves; at high temperatures, these
diatomic molecules reversibly dissociate a
pair of iodine atoms. Similarly, the iodide
anion, I−, is the strongest reducing agent
among the stable halogens, being the
most easily oxidised back to diatomic
I2.[17] (Astatine goes further, being indeed
unstable as At− and readily oxidised to At0
or At+, although the existence of At2 is not
settled.)[18]

The halogens darken in colour as the


group is descended: fluorine is a very pale
yellow gas, chlorine is greenish-yellow, and
bromine is a reddish-brown volatile liquid.
Iodine conforms to the prevailing trend,
being a shiny black crystalline solid that
melts at 114 °C and boils at 183 °C to form
a violet gas. This trend occurs because the
wavelengths of visible light absorbed by
the halogens increase down the group
(though astatine may not conform to it,
depending on how metallic it turns out to
be).[17] Specifically, the violet colour of
iodine gas results from the electron
transition between the highest occupied
antibonding πg molecular orbital and the
lowest vacant antibonding σu molecular
orbital.[19]
Elemental iodine is slightly soluble in
water, with one gram dissolving in 3450 ml
at 20 °C and 1280 ml at 50 °C; potassium
iodide may be added to increase solubility
via formation of triiodide ions, among
other polyiodides.[19] Nonpolar solvents
such as hexane and carbon tetrachloride
provide a higher solubility.[20] Polar
solutions, such as aqueous solutions, are
brown, reflecting the role of these solvents
as Lewis bases; on the other hand,
nonpolar solutions are violet, the color of
iodine vapour.[19] Charge-transfer
complexes form when iodine is dissolved
in polar solvents, hence changing the
colour. Iodine is violet when dissolved in
carbon tetrachloride and saturated
hydrocarbons but deep brown in alcohols
and amines, solvents that form charge-
transfer adducts.[21]

I2•PPh3 charge-transfer complexes in CH2Cl2. From left


to right: (1) I2 dissolved in dichloromethane – no CT
complex. (2) A few seconds after excess PPh3 was
added – CT complex is forming. (3) One minute later
after excess PPh3 was added, the CT complex
[Ph3PI]+I− has been formed. (4) Immediately after
excess I2 was added, which contains [Ph3PI]+[I3]−.[22]
The melting and boiling points of iodine
are the highest among the halogens,
conforming to the increasing trend down
the group, since iodine has the largest
electron cloud among them that is the
most easily polarised, resulting in its
molecules having the strongest van der
Waals interactions among the halogens.
Similarly, iodine is the least volatile of the
halogens.[17] Because it has the largest
atomic radius among the halogens, iodine
has the lowest first ionisation energy,
lowest electron affinity, lowest
electronegativity and lowest reactivity of
the halogens.[17]
Structure of solid iodine

The interhalogen bond in diiodine is the


weakest of all the halogens. As such, 1%
of a sample of gaseous iodine at
atmospheric pressure is dissociated into
iodine atoms at 575 °C. Temperatures
greater than 750 °C are required for
fluorine, chlorine, and bromine to
dissociate to a similar extent. Most bonds
to iodine are weaker than the analogous
bonds to the lighter halogens.[17] Gaseous
iodine is composed of I2 molecules with
an I–I bond length of 266.6 pm. The I–I
bond is one of the longest single bonds
known. It is even longer (271.5 pm) in solid
orthorhombic crystalline iodine, which has
the same crystal structure as chlorine and
bromine. (The record is held by iodine's
neighbour xenon: the Xe–Xe bond length
is 308.71 pm.)[23] As such, within the
iodine molecule, significant electronic
interactions occur with the two next-
nearest neighbours of each atom, and
these interactions give rise, in bulk iodine,
to a shiny appearance and
semiconducting properties.[17] Iodine is a
two-dimensional semiconductor with a
band gap of 1.3 eV (125 kJ/mol): it is a
semiconductor in the plane of its
crystalline layers and an insulator in the
perpendicular direction.[17]

Isotopes

Of the thirty-seven known isotopes of


iodine, only one occurs in nature, iodine-
127. The others are radioactive and have
half-lives too short to be primordial. As
such, iodine is monoisotopic and its
atomic weight is known to great precision,
as it is a constant of nature.[17]

The longest-lived of the radioactive


isotopes of iodine is iodine-129, which has
a half-life of 15.7 million years, decaying
via beta decay to stable xenon-129.[24]
Some iodine-129 was formed along with
iodine-127 before the formation of the
Solar System, but it has by now completely
decayed away, making it an extinct
radionuclide that is nevertheless still
useful in dating the history of the early
Solar System or very old groundwaters,
due to its mobility in the environment. Its
former presence may be determined from
an excess of its daughter xenon-
129.[25][26][27][28][29] Traces of iodine-129
still exist today, as it is also a cosmogenic
nuclide, formed from cosmic ray spallation
of atmospheric xenon: these traces make
up 10−14 to 10−10 of all terrestrial iodine. It
also occurs from open-air nuclear testing,
and is not hazardous because of its
incredibly long half-life, the longest of all
fission products. At the peak of
thermonuclear testing in the 1960s and
1970s, iodine-129 still made up only about
10−7 of all terrestrial iodine.[30] Excited
states of iodine-127 and iodine-129 are
often used in Mössbauer spectroscopy.[17]

The other iodine radioisotopes have much


shorter half-lives, no longer than days.[24]
Some of them have medical applications
involving the thyroid gland, where the
iodine that enters the body is stored and
concentrated. Iodine-123 has a half-life of
thirteen hours and decays by electron
capture to tellurium-123, emitting gamma
radiation; it is used in nuclear medicine
imaging, including single photon emission
computed tomography (SPECT) and X-ray
computed tomography (X-Ray CT)
scans.[31] Iodine-125 has a half-life of fifty-
nine days, decaying by electron capture to
tellurium-125 and emitting low-energy
gamma radiation; the second-longest-lived
iodine radioisotope, it has uses in
biological assays, nuclear medicine
imaging and in radiation therapy as
brachytherapy to treat a number of
conditions, including prostate cancer,
uveal melanomas, and brain tumours.[32]
Finally, iodine-131, with a half-life of eight
days, beta decays to an excited state of
stable xenon-131 that then converts to the
ground state by emitting gamma radiation.
It is a common fission product and thus is
present in high levels in radioactive fallout.
It may then be absorbed through
contaminated food, and will also
accumulate in the thyroid. As it decays, it
may cause damage to the thyroid. The
primary risk from exposure to high levels
of iodine-131 is the chance occurrence of
radiogenic thyroid cancer in later life.
Other risks include the possibility of non-
cancerous growths and thyroiditis.[33]
The usual means of protection against the
negative effects of iodine-131 is by
saturating the thyroid gland with stable
iodine-127 in the form of potassium iodide
tablets, taken daily for optimal
prophylaxis.[34] However, iodine-131 may
also be used for medicinal purposes in
radiation therapy for this very reason,
when tissue destruction is desired after
iodine uptake by the tissue.[35] Iodine-131
is also used as a radioactive
tracer.[36][37][38][39]

Chemistry and compounds


Halogen bond energies (kJ/mol)[19]
X XX HX BX3 AlX3 CX4
F 159 574 645 582 456
Cl 243 428 444 427 327
Br 193 363 368 360 272
I 151 294 272 285 239

Though it is the least reactive of the


halogens, iodine is still one of the more
reactive elements. For example, while
chlorine gas will halogenate carbon
monoxide, nitric oxide, and sulfur dioxide
(to phosgene, nitrosyl chloride, and sulfuryl
chloride respectively), iodine will not do
so. Furthermore, iodination of metals
tends to result in lower oxidation states
than chlorination or bromination; for
example, rhenium metal reacts with
chlorine to form rhenium hexachloride, but
with bromine it forms only rhenium
pentabromide and iodine can achieve only
rhenium tetraiodide.[17] By the same token,
however, since iodine has the lowest
ionisation energy among the halogens and
is the most easily oxidised of them, it has
a more significant cationic chemistry and
its higher oxidation states are rather more
stable than those of bromine and chlorine,
for example in iodine heptafluoride.[19]

Hydrogen iodide

The simplest compound of iodine is


hydrogen iodide, HI. It is a colourless gas
that reacts with oxygen to give water and
iodine. Although it is useful in iodination
reactions in the laboratory, it does not
have large-scale industrial uses, unlike the
other hydrogen halides. Commercially, it is
usually made by reacting iodine with
hydrogen sulfide or hydrazine:[40]

H2O
2 I2 + N2H4 ⟶ 4 HI + N2

At room temperature, it is a colourless gas,


like all of the hydrogen halides except
hydrogen fluoride, since hydrogen cannot
form strong hydrogen bonds to the large
and only mildly electronegative iodine
atom. It melts at −51.0 °C and boils at
−35.1 °C. It is an endothermic compound
that can exothermically dissociate at room
temperature, although the process is very
slow unless a catalyst is present: the
reaction between hydrogen and iodine at
room temperature to give hydrogen iodide
does not proceed to completion. The H–I
bond dissociation energy is likewise the
smallest of the hydrogen halides, at
295 kJ/mol.[41]

Aqueous hydrogen iodide is known as


hydroiodic acid, which is a strong acid.
Hydrogen iodide is exceptionally soluble in
water: one litre of water will dissolve 425
litres of hydrogen iodide, and the saturated
solution has only four water molecules per
molecule of hydrogen iodide.[42]
Commercial so-called "concentrated"
hydroiodic acid usually contains 48–57%
HI by mass; the solution forms an
azeotrope with boiling point 126.7 °C at
56.7 g HI per 100 g solution. Hence
hydroiodic acid cannot be concentrated
past this point by evaporation of water.[41]

Unlike hydrogen fluoride, anhydrous liquid


hydrogen iodide is difficult to work with as
a solvent, because its boiling point is low,
it has a small liquid range, its dielectric
constant is low and it does not dissociate
appreciably into H2I+ and HI−2 ions – the
latter, in any case, are much less stable
than the bifluoride ions (HF−2) due to the
very weak hydrogen bonding between
hydrogen and iodine, though its salts with
very large and weakly polarising cations
such as Cs+ and NR+4 (R = Me, Et, Bun) may
still be isolated. Anhydrous hydrogen
iodide is a poor solvent, able to dissolve
only small molecular compounds such as
nitrosyl chloride and phenol, or salts with
very low lattice energies such as
tetraalkylammonium halides.[41]

Other binary iodides

Nearly all elements in the periodic table


form binary iodides. The exceptions are
decidedly in the minority and stem in each
case from one of three causes: extreme
inertness and reluctance to participate in
chemical reactions (the noble gases);
extreme nuclear instability hampering
chemical investigation before decay and
transmutation (many of the heaviest
elements beyond bismuth); and having an
electronegativity higher than iodine's
(oxygen, nitrogen, and the first three
halogens), so that the resultant binary
compounds are formally not iodides but
rather oxides, nitrides, or halides of iodine.
(Nonetheless, nitrogen triiodide is named
as an iodide as it is analogous to the other
nitrogen trihalides.)[43]
Given the large size of the iodide anion
and iodine's weak oxidising power, high
oxidation states are difficult to achieve in
binary iodides, the maximum known being
in the pentaiodides of niobium, tantalum,
and protactinium. Iodides can be made by
reaction of an element or its oxide,
hydroxide, or carbonate with hydroiodic
acid, and then dehydrated by mildly high
temperatures combined with either low
pressure or anhydrous hydrogen iodide
gas. These methods work best when the
iodide product is stable to hydrolysis;
otherwise, the possibilities include high-
temperature oxidative iodination of the
element with iodine or hydrogen iodide,
high-temperature iodination of a metal
oxide or other halide by iodine, a volatile
metal halide, carbon tetraiodide, or an
organic iodide. For example,
molybdenum(IV) oxide reacts with
aluminium(III) iodide at 230 °C to give
molybdenum(II) iodide. An example
involving halogen exchange is given below,
involving the reaction of tantalum(V)
chloride with excess aluminium(III) iodide
at 400 °C to give tantalum(V) iodide:[43]

Lower iodides may be produced either


through thermal decomposition or
disproportionation, or by reducing the
higher iodide with hydrogen or a metal, for
example:[43]

Most of the iodides of the pre-transition


metals (groups 1, 2, and 3, along with the
lanthanides and actinides in the +2 and +3
oxidation states) are mostly ionic, while
nonmetals tend to form covalent
molecular iodides, as do metals in high
oxidation states from +3 and above. Ionic
iodides MIn tend to have the lowest
melting and boiling points among the
halides MXn of the same element, because
the electrostatic forces of attraction
between the cations and anions are
weakest for the large iodide anion. In
contrast, covalent iodides tend to instead
have the highest melting and boiling
points among the halides of the same
element, since iodine is the most
polarisable of the halogens and, having the
most electrons among them, can
contribute the most to van der Waals
forces. Naturally, exceptions abound in
intermediate iodides where one trend
gives way to the other. Similarly,
solubilities in water of predominantly ionic
iodides (e.g. potassium and calcium) are
the greatest among ionic halides of that
element, while those of covalent iodides
(e.g. silver) are the lowest of that element.
In particular, silver iodide is very insoluble
in water and its formation is often used as
a qualitative test for iodine.[43]

Iodine halides

The halogens form many binary,


diamagnetic interhalogen compounds with
stoichiometries XY, XY3, XY5, and XY7
(where X is heavier than Y), and iodine is
no exception. Iodine forms all three
possible diatomic interhalogens, a
trifluoride and trichloride, as well as a
pentafluoride and, exceptionally among
the halogens, a heptafluoride. Numerous
cationic and anionic derivatives are also
characterised, such as the wine-red or
bright orange compounds of ICl+2 and the
dark brown or purplish black compounds
of I2Cl+. Apart from these, some
pseudohalides are also known, such as
cyanogen iodide (ICN), iodine thiocyanate
(ISCN), and iodine azide (IN3).[44]
Iodine monochloride

Iodine monofluoride (IF) is unstable at


room temperature and disproportionates
very readily and irreversibly to iodine and
iodine pentafluoride, and thus cannot be
obtained pure. It can be synthesised from
the reaction of iodine with fluorine gas in
trichlorofluoromethane at −45 °C, with
iodine trifluoride in trichlorofluoromethane
at −78 °C, or with silver(I) fluoride at
0 °C.[44] Iodine monochloride (ICl) and
iodine monobromide (IBr), on the other
hand, are moderately stable. The former, a
volatile red-brown compound, was
discovered independently by Joseph Louis
Gay-Lussac and Humphry Davy in 1813–4
not long after the discoveries of chlorine
and iodine, and it mimics the intermediate
halogen bromine so well that Justus von
Liebig was misled into mistaking bromine
(which he had found) for iodine
monochloride. Iodine monochloride and
iodine monobromide may be prepared
simply by reacting iodine with chlorine or
bromine at room temperature and purified
by fractional crystallisation. Both are quite
reactive and attack even platinum and
gold, though not boron, carbon, cadmium,
lead, zirconium, niobium, molybdenum,
and tungsten. Their reaction with organic
compounds depends on conditions. Iodine
chloride vapour tends to chlorinate phenol
and salicyclic acid, since when iodine
chloride undergoes homolytic
dissociation, chlorine and iodine are
produced and the former is more reactive.
However, iodine chloride in
tetrachloromethane solution results in
iodination being the main reaction, since
now heterolytic fission of the I–Cl bond
occurs and I+ attacks phenol as an
electrophile. However, iodine
monobromide tends to brominate phenol
even in tetrachloromethane solution
because it tends to dissociate into its
elements in solution, and bromine is more
reactive than iodine.[44] When liquid, iodine
monochloride and iodine monobromide
+
dissociate into I2X and IX−2 anions (X = Cl,
Br); thus they are significant conductors of
electricity and can be used as ionising
solvents.[44]

Iodine trifluoride (IF3) is an unstable yellow


solid that decomposes above −28 °C. It is
thus little-known. It is difficult to produce
because fluorine gas would tend to oxidise
iodine all the way to the pentafluoride;
reaction at low temperature with xenon
difluoride is necessary. Iodine trichloride,
which exists in the solid state as the
planar dimer I2Cl6, is a bright yellow solid,
synthesised by reacting iodine with liquid
chlorine at −80 °C; caution is necessary
during purification because it easily
dissociates to iodine monochloride and
chlorine and hence can act as a strong
chlorinating agent. Liquid iodine trichloride
conducts electricity, possibly indicating
dissociation to ICl+2 and ICl−4 ions.[45]

Iodine pentafluoride (IF5), a colourless,


volatile liquid, is the most
thermodynamically stable iodine fluoride,
and can be made by reacting iodine with
fluorine gas at room temperature. It is a
fluorinating agent, but is mild enough to
store in glass apparatus. Again, slight
electrical conductivity is present in the
liquid state because of dissociation to IF+4
and IF−6. The pentagonal bipyramidal iodine
heptafluoride (IF7) is an extremely
powerful fluorinating agent, behind only
chlorine trifluoride, chlorine pentafluoride,
and bromine pentafluoride among the
interhalogens: it reacts with almost all the
elements even at low temperatures,
fluorinates Pyrex glass to form iodine(VII)
oxyfluoride (IOF5), and sets carbon
monoxide on fire.[46]

Iodine oxides and oxoacids


Structure of iodine pentoxide

Iodine oxides are the most stable of all the


halogen oxides, because of the strong I–O
bonds resulting from the large
electronegativity difference between
iodine and oxygen, and they have been
known for the longest time.[21] The stable,
white, hygroscopic iodine pentoxide (I2O5)
has been known since its formation in
1813 by Gay-Lussac and Davy. It is most
easily made by the dehydration of iodic
acid (HIO3), of which it is the anhydride. It
will quickly oxidise carbon monoxide
completely to carbon dioxide at room
temperature, and is thus a useful reagent
in determining carbon monoxide
concentration. It also oxidises nitrogen
oxide, ethylene, and hydrogen sulfide. It
reacts with sulfur trioxide and
peroxydisulfuryl difluoride (S2O6F2) to
form salts of the iodyl cation, [IO2]+, and is
reduced by concentrated sulfuric acids to
iodosyl salts involving [IO]+. It may be
fluorinated by fluorine, bromine trifluoride,
sulfur tetrafluoride, or chloryl fluoride,
resulting iodine pentafluoride, which also
reacts with iodine pentoxide, giving
iodine(V) oxyfluoride, IOF3. A few other
less stable oxides are known, notably I4O9
and I2O4; their structures have not been
determined, but reasonable guesses are
IIII(IVO3)3 and [IO]+[IO3]− respectively.[47]
Standard reduction potentials for aqueous I species[48]
+ −
E°(couple) a(H ) = 1 E°(couple) a(OH ) = 1
(acid) (base)
I2/I− +0.535 I2/I− +0.535
HOI/I− +0.987 IO−/I− +0.48
    IO−3/I− +0.26
HOI/I2 +1.439 IO−/I2 +0.42
IO−3/I2 +1.195    
IO−3/HOI +1.134 IO−3/IO− +0.15
IO−4/IO−3 +1.653    
H5IO6/IO−3 +1.601 H3IO2− −
6 /IO3 +0.65

More important are the four oxoacids:


hypoiodous acid (HIO), iodous acid (HIO2),
iodic acid (HIO3), and periodic acid (HIO4
or H5IO6). When iodine dissolves in
aqueous solution, the following reactions
occur:[48]
I2 + ⇌ HIO + H+ Kac = 2.0 × 10−13
H2O + I− mol2 l−2
I2 + 2 ⇌ IO− + H2O
Kalk = 30 mol−1 l
OH− + I−

Hypoiodous acid is unstable to


disproportionation. The hypoiodite ions
thus formed disproportionate immediately
to give iodide and iodate:[48]

3 IO− ⇌ 2 I− + IO−3 K = 1020

Iodous acid and iodite are even less stable


and exist only as a fleeting intermediate in
the oxidation of iodide to iodate, if at all.[48]
Iodates are by far the most important of
these compounds, which can be made by
oxidising alkali metal iodides with oxygen
at 600 °C and high pressure, or by
oxidising iodine with chlorates. Unlike
chlorates, which disproportionate very
slowly to form chloride and perchlorate,
iodates are stable to disproportionation in
both acidic and alkaline solutions. From
these, salts of most metals can be
obtained. Iodic acid is most easily made
by oxidation of an aqueous iodine
suspension by electrolysis or fuming nitric
acid. Iodate has the weakest oxidising
power of the halates, but reacts the
quickest.[49]
Many periodates are known, including not
only the expected tetrahedral IO−4, but also
square-pyramidal IO3−
5 , octahedral
orthoperiodate IO5−
6 , [IO3(OH)3]2−,

[I2O8(OH2)]4−, and I2O4−


9 . They are usually
made by oxidising alkaline sodium iodate
electrochemically (with lead(IV) oxide as
the anode) or by chlorine gas:[50]

IO−3 + 6 OH− → IO5−


6 + 3 H2O + 2 e−

IO−3 + 6 OH− + Cl2 → IO5−


6 + 2 Cl−+3H O
2

They are thermodymically and kinetically


powerful oxidising agents, quickly
oxidising Mn2+ to MnO−4, and cleaving
glycols, α-diketones, α-ketols, α-
aminoalcohols, and α-diamines.[50]
Orthoperiodate especially stabilises high
oxidation states among metals because of
its very high negative charge of −5.
Orthoperiodic acid, H5IO6, is stable, and
dehydrates at 100 °C in a vacuum to
metaperiodic acid, HIO4. Attempting to go
further does not result in the nonexistent
iodine heptoxide (I2O7), but rather iodine
pentoxide and oxygen. Periodic acid may
be protonated by sulfuric acid to give the
I(OH)+6 cation, isoelectronic to Te(OH)6 and
Sb(OH)−6, and giving salts with bisulfate
and sulfate.[21]

Polyiodine compounds
When iodine dissolves in strong acids,
such as fuming sulfuric acid, a bright blue
paramagnetic solution including I+2 cations
is formed. A solid salt of the diiodine
cation may be obtained by oxidising iodine
with antimony pentafluoride:[21]

SO2
2 I2 + 5 SbF5 ⟶ 2 I2Sb2F11 + SbF3
20°C

The salt I2Sb2F11 is dark blue, and the blue


tantalum analogue I2Ta2F11 is also known.
Whereas the I–I bond length in I2 is
267 pm, that in I+2 is only 256 pm as the
missing electron in the latter has been
removed from an antibonding orbital,
making the bond stronger and hence
shorter. In fluorosulfuric acid solution,
deep-blue I+2 reversibly dimerises below
−60 °C, forming red rectangular
diamagnetic I2+
4 . Other polyiodine cations
are not as well-characterised, including
bent dark-brown or black I+3 and
centrosymmetric C2h green or black I+5,
known in the AsF−6 and AlCl−4 salts among
others.[21][51]

The only important polyiodide anion in


aqueous solution is linear triiodide, I−3. Its
formation explains why the solubility of
iodine in water may be increased by the
addition of potassium iodide solution:[21]
I2 + I− ⇌ I−3 (Keq = ~700 at 20 °C)

Many other polyiodides may be found


when solutions containing iodine and
iodide crystallise, such as I−5, I−9, I2−
4 , and I2−
8 ,
whose salts with large, weakly polarising
cations such as Cs+ may be isolated.[21][52]

Organoiodine compounds

Structure of the oxidising agent 2-iodoxybenzoic acid


Organoiodine compounds have been
fundamental in the development of
organic synthesis, such as in the Hofmann
elimination of amines,[53] the Williamson
ether synthesis,[54] the Wurtz coupling
reaction,[55] and in Grignard reagents.[56]

The carbon–iodine bond is a common


functional group that forms part of core
organic chemistry; formally, these
compounds may be thought of as organic
derivatives of the iodide anion. The
simplest organoiodine compounds, alkyl
iodides, may be synthesised by the
reaction of alcohols with phosphorus
triiodide; these may then be used in
nucleophilic substitution reactions, or for
preparing Grignard reagents. The C–I bond
is the weakest of all the carbon–halogen
bonds due to the minuscule difference in
electronegativity between carbon (2.55)
and iodine (2.66). As such, iodide is the
best leaving group among the halogens, to
such an extent that many organoiodine
compounds turn yellow when stored over
time due to decomposition into elemental
iodine; as such, they are commonly used in
organic synthesis, because of the easy
formation and cleavage of the C–I
bond.[57] They are also significantly denser
than the other organohalogen compounds
thanks to the high atomic weight of
iodine.[58] A few organic oxidising agents
like the iodanes contain iodine in a higher
oxidation state than −1, such as 2-
iodoxybenzoic acid, a common reagent for
the oxidation of alcohols to aldehydes,[59]
and iodobenzene dichloride (PhICl2), used
for the selective chlorination of alkenes
and alkynes.[60] One of the more well-
known uses of organoiodine compounds
is the so-called iodoform test, where
iodoform (CHI3) is produced by the
exhaustive iodination of a methyl ketone
(or another compound capable of being
oxidised to a methyl ketone), as follows:[61]
Some drawbacks of using organoiodine
compounds as compared to
organochlorine or organobromine
compounds is the greater expense and
toxicity of the iodine derivatives, since
iodine is expensive and organoiodine
compounds are stronger alkylating
agents.[62] For example, iodoacetamide
and iodoacetic acid denature proteins by
irreversibly alkylating cysteine residues
and preventing the reformation of disulfide
linkages.[63]
Halogen exchange to produce iodoalkanes
by the Finkelstein reaction is slightly
complicated by the fact that iodide is a
better leaving group than chloride or
bromide. The difference is nevertheless
small enough that the reaction can be
driven to completion by exploiting the
differential solubility of halide salts, or by
using a large excess of the halide salt.[61]
In the classic Finkelstein reaction, an alkyl
chloride or an alkyl bromide is converted
to an alkyl iodide by treatment with a
solution of sodium iodide in acetone.
Sodium iodide is soluble in acetone and
sodium chloride and sodium bromide are
not.[64] The reaction is driven toward
products by mass action due to the
precipitation of the insoluble salt.[65][66]

Occurrence and production


Iodine is the least abundant of the stable
halogens, comprising only 0.46 parts per
million of Earth's crustal rocks (compare:
fluorine 544 ppm, chlorine 126 ppm,
bromine 2.5 ppm). Among the eighty-four
elements which occur in significant
quantities (elements 1–42, 44–60, 62–83,
and 90–92), it ranks sixty-first in
abundance.[67] Iodide minerals are rare,
and most deposits that are concentrated
enough for economical extraction are
iodate minerals instead. Examples include
lautarite, Ca(IO3)2, and dietzeite,
7Ca(IO3)2·8CaCrO4.[67] These are the
minerals that occur as trace impurities in
the caliche, found in Chile, whose main
product is sodium nitrate. In total, they can
contain at least 0.02% and at most 1%
iodine by weight.[68] Sodium iodate is
extracted from the caliche and reduced to
iodide by sodium bisulfite. This solution is
then reacted with freshly extracted iodate,
resulting in comproportionation to iodine,
which may be filtered off.[17]

The caliche was the main source of iodine


in the 19th century and continues to be
important today, replacing kelp (which is
no longer an economically viable
source),[69] but in the late 20th century
brines emerged as a comparable source.
The Japanese Minami Kanto gas field east
of Tokyo and the American Anadarko
Basin gas field in northwest Oklahoma are
the two largest such sources. The brine is
hotter than 60 °C from the depth of the
source. The brine is first purified and
acidified using sulfuric acid, then the
iodide present is oxidised to iodine with
chlorine. An iodine solution is produced,
but is dilute and must be concentrated. Air
is blown into the solution to evaporate the
iodine, which is passed into an absorbing
tower where sulfur dioxide reduces the
iodine. The hydrogen iodide (HI) is reacted
with chlorine to precipitate the iodine.
After filtering and purification the iodine is
packed.[68][70]

2 HI + Cl2 → I2↑ + 2 HCl


I2 + 2 H2O + SO2 → 2 HI + H2SO4
2 HI + Cl2 → I2↓ + 2 HCl

These sources ensure that Chile and


Japan are the largest producers of iodine
today.[67] Alternatively, the brine may be
treated with silver nitrate to precipitate out
iodine as silver iodide, which is then
decomposed by reaction with iron to form
metallic silver and a solution of iron(II)
iodide. The iodine may then be liberated by
displacement with chlorine.[71]

Applications
Unlike chlorine and bromine, which have
one significant main use dwarfing all
others, iodine is used in many applications
of varying importance. About half of all
produced iodine goes into various
organoiodine compounds; another 15%
remains as the pure element, another 15%
is used to form potassium iodide, and
another 15% for other inorganic iodine
compounds. The remaining 5% is for
minor uses. Among the major uses of
iodine compounds are catalysts, animal
feed supplements, stabilisers, dyes,
colourants and pigments, pharmaceutical,
sanitation (from tincture of iodine), and
photography; minor uses include smog
inhibition, cloud seeding, and various uses
in analytical chemistry.[17]

Chemical analysis

Testing a seed for starch with a solution of iodine


Potassium tetraiodomercurate(II), K2HgI4,
is also known as Nessler's reagent. It is
often used as a sensitive spot test for
ammonia. Similarly, Cu2HgI4 is used as a
precipitating reagent to test for alkaloids.
The iodide and iodate anions are often
used for quantitative volumetric analysis,
for example in iodometry and the iodine
clock reaction (in which iodine also serves
as a test for starch, forming a dark blue
complex),[17] and aqueous alkaline iodine
solution is used in the iodoform test for
methyl ketones.[61] The iodine test for
starch is still used to detect counterfeit
banknotes printed on starch-containing
paper.[72]
Spectroscopy

The spectra of the iodine molecule, I2,


consists of (not exclusively) tens of
thousands of sharp spectral lines in the
wavelength range 500-700 nm. It is
therefore a commonly used wavelength
reference (secondary standard). By
measuring with a spectroscopic Doppler-
free technique while focusing on one of
these lines, the hyperfine structure of the
iodine molecule reveals itself. A line is now
resolved such that either 15 components,
(from even rotational quantum numbers,
Jeven), or 21 components (from odd
rotational quantum numbers, Jodd) are
measurable.

Medicine

Elemental iodine

Elemental iodine is used as a disinfectant


either as the element, or as the water-
soluble triiodide anion I3− generated in situ
by adding iodide to poorly water-soluble
elemental iodine (the reverse chemical
reaction makes some free elemental
iodine available for antisepsis). Elemental
iodine may also be used to treat iodine
deficiency.[73]
In the alternative, iodine may be produced
from iodophors, which contain iodine
complexed with a solubilizing agent
(iodide ion may be thought of loosely as
the iodophor in triiodide water solutions).
Examples of such preparations include:[74]

Tincture of iodine: iodine in ethanol, or


iodine and sodium iodide in a mixture of
ethanol and water.
Lugol's iodine: iodine and iodide in water
alone, forming mostly triiodide. Unlike
tincture of iodine, Lugol's iodine has a
minimised amount of the free iodine (I2)
component.
Povidone iodine (an iodophor).
The antimicrobial action of iodine is quick
and works at low concentrations, and thus
it is used in operating theatres.[75] Its
specific mode of action is unknown. It
penetrates into microorganisms and
attacks particular amino acids (such as
cysteine and methionine), nucleotides, and
fatty acids, ultimately resulting in cell
death. It also has an antiviral action, but
nonlipid viruses and parvoviruses are less
sensitive than lipid enveloped viruses.
Iodine probably attacks surface proteins
of enveloped viruses, and it may also
destabilise membrane fatty acids by
reacting with unsaturated carbon
bonds.[76]
Other formulations

In medicine, a saturated solution of


potassium iodide is used to treat acute
thyrotoxicosis. It is also used to block
uptake of iodine-131 in the thyroid gland
(see isotopes section above), when this
isotope is used as part of
radiopharmaceuticals (such as
iobenguane) that are not targeted to the
thyroid or thyroid-type tissues.[77][78]

Iodine-131 (usually as iodide) is a


component of nuclear fallout, and is
particularly dangerous owing to the thyroid
gland's propensity to concentrate ingested
iodine and retain it for periods longer than
this isotope's radiological half-life of eight
days. For this reason, people at risk of
exposure to environmental radioactive
iodine (iodine-131) in fallout may be
instructed to take non-radioactive
potassium iodide tablets. The typical adult
dose is one 130 mg tablet per 24 hours,
supplying 100 mg (100,000 micrograms)
of ionic iodine. (The typical daily dose of
iodine for normal health is of order 100
micrograms; see "Dietary Intake" below.)
Ingestion of this large dose of non-
radioactive iodine minimises the uptake of
radioactive iodine by the thyroid gland.[79]
Diatrizoic acid, an iodine-containing radiocontrast
agent

As an element with high electron density


and atomic number, iodine absorbs X-rays
weaker than 33.3 keV due to the
photoelectric effect of the innermost
electrons.[80] Organoiodine compounds
are used with intravenous injection as X-
ray radiocontrast agents. This application
is often in conjunction with advanced X-ray
techniques such as angiography and CT
scanning. At present, all water-soluble
radiocontrast agents rely on iodine.

Others

The production of ethylenediamine


dihydroiodide, provided as a nutritional
supplement for livestock, consumes a
large portion of available iodine. Another
significant use is a catalyst for the
production of acetic acid by the Monsanto
and Cativa processes. In these
technologies, which support the world's
demand for acetic acid, hydroiodic acid
converts the methanol feedstock into
methyl iodide, which undergoes
carbonylation. Hydrolysis of the resulting
acetyl iodide regenerates hydroiodic acid
and gives acetic acid.[81]

Inorganic iodides find specialised uses.


Titanium, zirconium, hafnium, and thorium
are purified by the van Arkel process,
which involves the reversible formation of
the tetraiodides of these elements. Silver
iodide is a major ingredient to traditional
photographic film. Thousands of
kilograms of silver iodide are used
annually for cloud seeding to induce
rain.[81]
The organoiodine compound erythrosine is
an important food coloring agent.
Perfluoroalkyl iodides are precursors to
important surfactants, such as
perfluorooctanesulfonic acid.[81]

Biological role

The thyroid system of the thyroid hormones T3 and T4


Comparison of the iodine content in urine in France (in
microgramme/day), for some regions and departments
(average levels of urine iodine, measured in
micrograms per liter at the end of the twentieth century
(1980 to 2000)[82]

Iodine is an essential element for life and,


at atomic number Z = 53, is the heaviest
element commonly needed by living
organisms. (Lanthanum and the other
lanthanides, as well as tungsten with Z =
74, are used by a few
microorganisms.)[83][84] It is required for
the synthesis of the growth-regulating
thyroid hormones thyroxine and
triiodothyronine (T4 and T3 respectively,
named after their number of iodine
atoms). A deficiency of iodine leads to
decreased production of T3 and T4 and a
concomitant enlargement of the thyroid
tissue in an attempt to obtain more iodine,
causing the disease known as simple
goitre. The major form of thyroid hormone
in the blood is thyroxine (T4), which has a
longer half-life than T3. In humans, the
ratio of T4 to T3 released into the blood is
between 14:1 and 20:1. T4 is converted to
the active T3 (three to four times more
potent than T4) within cells by deiodinases
(5'-iodinase). These are further processed
by decarboxylation and deiodination to
produce iodothyronamine (T1a) and
thyronamine (T0a'). All three isoforms of
the deiodinases are selenium-containing
enzymes; thus dietary selenium is
essential for T3 production.[85]

Iodine accounts for 65% of the molecular


weight of T4 and 59% of T3. Fifteen to
20 mg of iodine is concentrated in thyroid
tissue and hormones, but 70% of all iodine
in the body is found in other tissues,
including mammary glands, eyes, gastric
mucosa, fetal thymus, cerebro-spinal fluid
and choroid plexus, arterial walls, the
cervix, and salivary glands. In the cells of
those tissues, iodide enters directly by
sodium-iodide symporter (NIS). The action
of iodine in mammary tissue is related to
fetal and neonatal development, but in the
other tissues, it is (at least) partially
unknown.[86]

Dietary intake

Recommendations by the United States


Institute of Medicine are between 110 and
130 µg for infants up to 12 months, 90 µg
for children up to eight years, 130 µg for
children up to 13 years, 150 µg for adults,
220 µg for pregnant women and 290 µg
for lactation.[4][87] The Tolerable Upper
Intake Level (UL) for adults is
1,100 μg/day.[88] This upper limit was
assessed by analyzing the effect of
supplementation on thyroid-stimulating
hormone.[86]

The thyroid gland needs no more than


70 μg/day to synthesise the requisite daily
amounts of T4 and T3.[4] The higher
recommended daily allowance levels of
iodine seem necessary for optimal
function of a number of body systems,
including lactation, gastric mucosa,
salivary glands, brain cells, choroid plexus,
thymus, and arterial walls.[4][89][90][91]

Natural sources of dietary iodine include


seafood, such as fish, seaweeds (such as
kelp) and shellfish, dairy products and
eggs so long as the animals received
enough iodine, and plants grown on iodine-
rich soil.[92][93] Iodised salt is fortified with
iodine in the form of sodium iodide.[93][94]

As of 2000, the median intake of iodine


from food in the United States was 240 to
300 μg/day for men and 190 to
210 μg/day for women.[88] The general US
population has adequate iodine
nutrition,[95][96] with women of childbearing
age and pregnant women having a
possible mild risk of deficiency.[96] In
Japan, consumption was considered
much higher, ranging between
5,280 μg/day to 13,800 μg/day from
dietary seaweed or kombu kelp,[86] often in
the form of Kombu Umami extracts for
soup stock and potato chips. However,
new studies suggest that Japan's
consumption is closer to 1,000–
3,000 μg/day.[97] The adult UL in Japan
was last revised to 3,000 µg/day in
2015.[98]
After iodine fortification programs such as
iodisation of salt have been implemented,
some cases of iodine-induced
hyperthyroidism have been observed (so-
called Jod-Basedow phenomenon). The
condition seems to occur mainly in people
over forty, and the risk appears higher
when iodine deficiency is severe and the
initial rise in iodine intake is high.[99]

Deficiency

In areas where there is little iodine in the


diet,[100] typically remote inland areas and
semi-arid equatorial climates where no
marine foods are eaten, iodine deficiency
gives rise to hypothyroidism, symptoms of
which are extreme fatigue, goitre, mental
slowing, depression, weight gain, and low
basal body temperatures.[101] Iodine
deficiency is the leading cause of
preventable intellectual disability, a result
that occurs primarily when babies or small
children are rendered hypothyroidic by a
lack of the element. The addition of iodine
to table salt has largely eliminated this
problem in wealthier nations, but iodine
deficiency remains a serious public health
problem in the developing world today.[102]
Iodine deficiency is also a problem in
certain areas of Europe. Information
processing, fine motor skills, and visual
problem solving are improved by iodine
repletion in moderately iodine-deficient
children.[103]

Toxicity
Iodine
Hazards

GHS pictograms

GHS signal word Danger

GHS hazard statements H312, H332, H315,


H319, H335, H372,
H400

GHS precautionary P261, P273, P280,


statements
P305, P351, P338,
P314[104]

NFPA 704 0
[105]
3 0

Elemental iodine (I2) is toxic if taken orally


undiluted. The lethal dose for an adult
human is 30 mg/kg, which is about 2.1–
2.4 grams for a human weighing 70 to
80 kg (even if experiments on rats
demonstrated that these animals could
survive after eating a 14000 mg/kg dose).
Excess iodine can be more cytotoxic in the
presence of selenium deficiency.[106]
Iodine supplementation in selenium-
deficient populations is, in theory,
problematic, partly for this reason.[86] The
toxicity derives from its oxidizing
properties, through which it denaturates
proteins (including enzymes).[107]

Elemental iodine is also a skin irritant, and


direct contact with skin can cause damage
and solid iodine crystals should be
handled with care. Solutions with high
elemental iodine concentration, such as
tincture of iodine and Lugol's solution, are
capable of causing tissue damage if used
in prolonged cleaning or antisepsis;
similarly, liquid Povidone-iodine (Betadine)
trapped against the skin resulted in
chemical burns in some reported
cases.[108]

Occupational exposure

People can be exposed to iodine in the


workplace by inhalation, ingestion, skin
contact, and eye contact. The
Occupational Safety and Health
Administration (OSHA) has set the legal
limit (Permissible exposure limit) for
iodine exposure in the workplace at 0.1
ppm (1 mg/m3) during an 8-hour workday.
The National Institute for Occupational
Safety and Health (NIOSH) has set a
Recommended exposure limit (REL) of
0.1 ppm (1 mg/m3) during an 8-hour
workday. At levels of 2 ppm, iodine is
immediately dangerous to life and
health.[109]

Allergic reactions

Some people develop a hypersensitivity to


products and foods containing iodine.
Applications of tincture of iodine or
Betadine can cause rashes, sometimes
severe.[110] Parenteral use of iodine-based
contrast agents (see above) can cause
reactions ranging from a mild rash to fatal
anaphylaxis. Such reactions have led to
the misconception (widely held, even
among physicians) that some people are
allergic to iodine itself; even allergies to
iodine-rich seafood have been so
construed.[111] In fact, there has never
been a confirmed report of a true iodine
allergy, and an allergy to elemental iodine
or simple iodide salts is theoretically
impossible. Hypersensitivity reactions to
products and foods containing iodine are
apparently related to their other molecular
components;[112] thus, a person who has
demonstrated an allergy to one food or
product containing iodine may not have an
allergic reaction to another. Patients with
various food allergies (shellfish, egg, milk,
etc.) or asthma are more likely to suffer
reactions to contrast media containing
iodine.[112] As with all medications, the
patient's allergy history should be
questioned and consulted before any
containing iodine are administered.[113]

References
1. Meija, Juris; et al. (2016). "Atomic
weights of the elements 2013 (IUPAC
Technical Report)". Pure and Applied
Chemistry. 88 (3): 265–91.
doi:10.1515/pac-2015-0305 .
2. Magnetic susceptibility of the elements
and inorganic compounds , in Handbook of
Chemistry and Physics 81st edition, CRC
press.
3. Weast, Robert (1984). CRC, Handbook of
Chemistry and Physics. Boca Raton, Florida:
Chemical Rubber Company Publishing.
pp. E110. ISBN 0-8493-0464-4.
4. "Iodine" . Micronutrient Information
Center, Linus Pauling Institute, Oregon State
University, Corvallis. 2015. Retrieved
20 November 2017.
5. Courtois, Bernard (1813). "Découverte
d'une substance nouvelle dans le Vareck" .
Annales de chimie. 88: 304. In French,
seaweed that had been washed onto the
shore was called "varec", "varech", or
"vareck", whence the English word "wrack".
Later, "varec" also referred to the ashes of
such seaweed: The ashes were used as a
source of iodine and salts of sodium and
potassium.
6. Swain, Patricia A. (2005). "Bernard
Courtois (1777–1838) famed for
discovering iodine (1811), and his life in
Paris from 1798" (PDF). Bulletin for the
History of Chemistry. 30 (2): 103. Archived
from the original (PDF) on 14 July 2010.
Retrieved 2 April 2009.
7. Greenwood and Earnshaw, p. 794
8. "53 Iodine" . Elements.vanderkrogt.net.
Retrieved 23 October 2016.
9. Desormes and Clément made their
announcement at the Institut impérial de
France on 29 November 1813; a summary
of their announcement appeared in the
Gazette nationale ou Le Moniteur Universel
of 2 December 1813. See: F. D. Chattaway
(23 April 1909) "The discovery of iodine,"
The Chemical News... , 99 (2578) : 193–
195.
10. Gay-Lussac, J. (1813). "Sur un nouvel
acide formé avec la substance décourverte
par M. Courtois" . Annales de Chimie. 88:
311.
11. Gay-Lussac, J. (1813). "Sur la
combination de l'iode avec d'oxigène" .
Annales de Chimie. 88: 319.
12. Gay-Lussac, J. (1814). "Mémoire sur
l'iode" . Annales de Chimie. 91: 5.
13. Liddell–Scott–Jones Greek–English
Lexicon at the Perseus Digital Library
14. Davy, H. (1813). "Sur la nouvelle
substance découverte par M. Courtois,
dans le sel de Vareck" . Annales de Chimie.
88: 322.
15. Davy, Humphry (1 January 1814). "Some
Experiments and Observations on a New
Substance Which Becomes a Violet
Coloured Gas by Heat". Phil. Trans. R. Soc.
Lond. 104: 74. doi:10.1098/rstl.1814.0007 .
16. "Mendeleev's First Periodic Table" .
web.lemoyne.edu.
17. Greenwood and Earnshaw, pp. 800–4
18. Kugler, H. K.; Keller, C. (1985). 'At,
Astatine', System No. 8a. Gmelin Handbook
of Inorganic and Organometallic Chemistry.
8 (8th ed.). Springer-Verlag. ISBN 978-3-
540-93516-2.
19. Greenwood and Earnshaw, pp. 804–9
20. Windholz, Martha; Budavari, Susan;
Stroumtsos, Lorraine Y.; Fertig, Margaret
Noether, eds. (1976). Merck Index of
Chemicals and Drugs (9th ed.). J A Majors
Company. ISBN 978-0-911910-26-1.
21. King, R. Bruce (1995). Inorganic
Chemistry of Main Group Elements. Wiley-
VCH. pp. 173–98. ISBN 978-0-471-18602-1.
22. Housecroft, C. E.; Sharpe, A. G. (2008).
Inorganic Chemistry (3rd ed.). Prentice Hall.
p. 541. ISBN 978-0-13-175553-6.
23. Li, Wai-Kee; Zhou, Gong-Du; Mak,
Thomas C. W. (2008). Advanced Structural
Inorganic Chemistry. Oxford University
Press. p. 674. ISBN 978-0-19-921694-9.
24. Audi, Georges; Bersillon, Olivier; Blachot,
Jean; Wapstra, Aaldert Hendrik (2003), "The
NUBASE evaluation of nuclear and decay
properties" , Nuclear Physics A, 729: 3–128,
Bibcode:2003NuPhA.729....3A ,
doi:10.1016/j.nuclphysa.2003.11.001
25. J. Throck Watson; David K. Roe; Herbert
A. Selenkow (1965). "Iodine-129 as a
Nonradioactive Tracer". Radiation Research.
26 (1): 159–163. doi:10.2307/3571805 .
JSTOR 3571805 .
26. https://e-reports-
ext.llnl.gov/pdf/234761.pdf P. Santschi et
al. (1998) "129Iodine: A new tracer for
surface water/groundwater interaction."
Lawrence Livermore National Laboratory
preprint UCRL-JC-132516. Livermore, USA.
27. Snyder, G.; Fabryka-Martin, J. (2007). "I-
129 and Cl-36 in dilute hydrocarbon waters:
Marine-cosmogenic, in situ, and
anthropogenic sources". Applied
Geochemistry. 22 (3): 692–714.
Bibcode:2007ApGC...22..692S .
doi:10.1016/j.apgeochem.2006.12.011 .
28. Clayton, Donald D. (1983). Principles of
Stellar Evolution and Nucleosynthesis (2nd
ed.). University of Chicago Press. p. 75.
ISBN 978-0-226-10953-4.
29. Bolt, B. A.; Packard, R. E.; Price, P. B.
(2007). "John H. Reynolds, Physics:
Berkeley" . The University of California,
Berkeley. Retrieved 2007-10-01.
30. SCOPE 50 - Radioecology after
Chernobyl Archived 13 May 2014 at the
Wayback Machine, the Scientific Committee
on Problems of the Environment (SCOPE),
1993. See table 1.9 in Section 1.4.5.2.
31. Hupf HB, Eldridge JS, Beaver JE (April
1968). "Production of iodine-123 for
medical applications". Int J Appl Radiat Isot.
19 (4): 345–51. doi:10.1016/0020-
708X(68)90178-6 . PMID 5650883 .
32. Harper, P.V. ; Siemens, W.D. ; Lathrop,
K.A. ; Brizel, H.E. ; Harrison, R.W. Iodine-125.
Proc. Japan Conf. Radioisotopes; Vol: 4th
Jan 01, 1961
33. Rivkees, Scott A.; Sklar, Charles;
Freemark, Michael (1998). "The
Management of Graves' Disease in Children,
with Special Emphasis on Radioiodine
Treatment". Journal of Clinical
Endocrinology & Metabolism. 83 (11):
3767–76. doi:10.1210/jc.83.11.3767 .
PMID 9814445 .
34. Zanzonico PB, Becker DV (2000).
"Effects of time of administration and
dietary iodine levels on potassium iodide
(KI) blockade of thyroid irradiation by 131I
from radioactive fallout". Health Phys. 78
(6): 660–7. doi:10.1097/00004032-
200006000-00008 . PMID 10832925 .
35. "Medical isotopes the likely cause of
radiation in Ottawa waste" . CBC News. 4
February 2009. Retrieved 30 September
2015.
36. Moser, H.; Rauert, W. (2007). "Isotopic
Tracers for Obtaining Hydrologic
Parameters" . In Aggarwal, Pradeep K.; Gat,
Joel R.; Froehlich, Klaus F. Isotopes in the
water cycle : past, present and future of a
developing science. Dordrecht: Springer.
p. 11. ISBN 978-1-4020-6671-9. Retrieved
6 May 2012.
37. Rao, S. M. (2006). "Radioisotopes of
hydrological interest" . Practical isotope
hydrology. New Delhi: New India Publishing
Agency. pp. 12–13. ISBN 978-81-89422-33-
2. Retrieved 6 May 2012.
38. "Investigating leaks in Dams &
Reservoirs" (PDF). IAEA.org. Archived from
the original (PDF) on 30 July 2013.
Retrieved 6 May 2012.
39. Araguás, Luis Araguás; Plata Bedmar,
Antonio (2002). "Artificial radioactive
tracers" . Detection and prevention of leaks
from dams. Taylor & Francis. pp. 179–181.
ISBN 978-90-5809-355-4. Retrieved 6 May
2012.
40. Greenwood and Earnshaw, pp. 809–12
41. Greenwood and Earnshaw, pp. 812–9
42. Holleman, A. F.; Wiberg, E. "Inorganic
Chemistry" Academic Press: San Diego,
2001. ISBN 0-12-352651-5.
43. Greenwood and Earnshaw, pp. 821–4
44. Greenwood and Earnshaw, pp. 824–8
45. Greenwood and Earnshaw, pp. 828–31
46. Greenwood and Earnshaw, pp. 832–5
47. Greenwood and Earnshaw, pp. 851–3
48. Greenwood and Earnshaw, pp. 853–9
49. Greenwood and Earnshaw, pp. 863–4
50. Greenwood and Earnshaw, pp. 872–5
51. Greenwood and Earnshaw, pp. 842–4
52. Greenwood and Earnshaw, pp. 835–9
53. Aug. Wilh. von Hofmann (1851).
"Beiträge zur Kenntniss der flüchtigen
organischen Basen". Annalen der Chemie
und Pharmacie. 78 (3): 253–286.
doi:10.1002/jlac.18510780302 .
54. Williamson, Alexander (1850). "Theory
of Aetherification". Philosophical Magazine.
37 (251): 350–356.
doi:10.1080/14786445008646627 . (Link to
excerpt. )
55. Adolphe Wurtz (1855). "Ueber eine neue
Klasse organischer Radicale". Annalen der
Chemie und Pharmacie. 96 (3): 364–375.
doi:10.1002/jlac.18550960310 .
56. Grignard, V. (1900). "Sur quelques
nouvelles combinaisons organométaliques
du magnésium et leur application à des
synthèses d'alcools et d'hydrocabures" .
Compt. Rend. 130: 1322–25.
57. Phyllis A. Lyday, "Iodine and Iodine
Compounds", Ullmann's Encyclopedia of
Industrial Chemistry, Weinheim: Wiley-VCH,
doi:10.1002/14356007.a14_381
58. Blanksby SJ, Ellison GB (April 2003).
"Bond dissociation energies of organic
molecules" (PDF). Acc. Chem. Res. 36 (4):
255–63. CiteSeerX 10.1.1.616.3043 .
doi:10.1021/ar020230d . PMID 12693923 .
Archived from the original (PDF) on 6
February 2009. Retrieved 25 October 2017.
59. Boeckman, R. K. Jr.; Shao, P.; Mullins, J.
J. (2000). "Dess–Martin periodinane: 1,1,1-
Triacetoxy-1,1-dihydro-1,2-benziodoxol-
3(1H)-one" (PDF). Organic Syntheses. 77:
141.; Collective Volume, 10, p. 696
60. Michael E. Jung and Michael H. Parker
(1997). "Synthesis of Several Naturally
Occurring Polyhalogenated Monoterpenes
of the Halomon Class". Journal of Organic
Chemistry. 62 (21): 7094–7095.
doi:10.1021/jo971371 . PMID 11671809 .
61. Smith, Michael B.; March, Jerry (2007),
Advanced Organic Chemistry: Reactions,
Mechanisms, and Structure (6th ed.), New
York: Wiley-Interscience, ISBN 978-0-471-
72091-1
62. "Safety data for iodomethane" . Oxford
University.
63. Polgar, L (1979). "Deuterium isotope
effects on papain acylation. Evidence for
lack of general base catalysis and for
enzyme-leaving group. interaction". Eur. J.
Biochem. 98 (2): 369–374.
doi:10.1111/j.1432-1033.1979.tb13196.x .
PMID 488108 .
64. Ervithayasuporn, V. (2013). "One-pot
synthesis of halogen exchanged
silsesquioxanes: octakis(3-
bromopropyl)octasilsesquioxane and
octakis(3-iodopropyl)octasilsesquioxane".
Dalton Trans. 42 (37): 13747–13753.
doi:10.1039/C3DT51373D .
PMID 23907310 .
65. Streitwieser, A. (1956). "Solvolytic
Displacement Reactions at Saturated
Carbon Atoms". Chem. Rev. 56 (4): 571–
752. doi:10.1021/cr50010a001 .
66. Bordwell, F. G.; Brannen, W. T. (1964).
"The Effect of the Carbonyl and Related
Groups on the Reactivity of Halides in SN2
Reactions". J. Am. Chem. Soc. 86 (21):
4645–4650. doi:10.1021/ja01075a025 .
67. Greenwood and Earnshaw, pp. 795–6
68. Kogel, Jessica Elzea; et al. (2006).
Industrial Minerals & Rocks: Commodities,
Markets, and Uses . SME. pp. 541–552.
ISBN 978-0-87335-233-8.
69. Stanford, Edward C. C. (1862). "On the
Economic Applications of Seaweed" .
Journal of the Society of Arts: 185–189.
70. Maekawa, Tatsuo; Igari, Shun-Ichiro;
Kaneko, Nobuyuki (2006). "Chemical and
isotopic compositions of brines from
dissolved-in-water type natural gas fields in
Chiba, Japan". Geochemical Journal. 40 (5):
475. doi:10.2343/geochemj.40.475 .
71. Greenwood and Earnshaw, p. 799
72. Emsley, John (2001). Nature's Building
Blocks (Hardcover, First ed.). Oxford
University Press. pp. 244–250. ISBN 978-0-
19-850340-8.
73. WHO Model Formulary 2008 (PDF).
World Health Organization. 2009. p. 499.
ISBN 9789241547659. Retrieved 8 January
2017.
74. Block, Seymour Stanton (2001).
Disinfection, sterilization, and preservation.
Hagerstwon, MD: Lippincott Williams &
Wilkins. p. 159. ISBN 978-0-683-30740-5.
75. Patwardhan, Narendra; Kelkar, Uday
(2011). "Disinfection, sterilization and
operation theater guidelines for
dermatosurgical practitioners in India".
Dermatosurgery Specials. 77 (1): 83–93.
doi:10.4103/0378-6323.74965 .
PMID 21220895 .
76. McDonnell G, Russell AD (1999).
"Antiseptics and disinfectants: activity,
action, and resistance" . Clin Microbiol Rev.
12 (1): 147–79.
doi:10.1128/CMR.12.1.147 . PMC 88911 .
PMID 9880479 .
77. "Solubility of KI in water" . Hazard.com.
1998-04-21. Retrieved 2013-01-21.
78.
https://web.archive.org/web/20111007202
042/https://www.eanm.org/scientific_info/
guidelines/gl_radio_ther_benzyl.pdf?
PHPSESSID=46d05b62d235c36a12166bf93
9b656c7
79. U.S. Centers for Disease Control "CDC
Radiation Emergencies" , U.S. Centers for
Disease Control, 11 October 2006, accessed
14 November 2010.
80. Lancaster, Jack L. Chapter 4: Physical
Determinants of Contrast Archived 10
October 2015 at the Wayback Machine, in
Physics of Medical X-Ray Imaging. The
University of Texas Health Science Center.
81. Lyday, Phyllis A.; Tatsuo Kaiho"Iodine
and Iodine Compounds" in Ullmann's
Encyclopedia of Industrial Chemistry, 2015,
Wiley-VCH, Weinheim,
doi:10.1002/14356007.a14_381.pub2 Vol.
A14 pp. 382–390.
82. Mornex,1987 et LeGuenelal., 2000, cités
par Le Guen, B., Hémidy, P. Y., Gonin, M.,
Bailloeuil, C., Van Boxsom, D., Renier, S., &
Garcier, Y. (2001). Arguments et retour
d'expérience sur la distribution d'iode stable
autour des centrales nucléaires françaises.
Radioprotection, 36(4), 417-430.
URL:https://www.researchgate.net/profile/B
ernard_Le_Guen/publication/245276139_Ar
guments_et_retour_d%27experiencesur_la_
distribution_d%27iode_stable_autourdes_ce
ntrales_nucleaires_francaises/links/57cfff4
b08ae057987ae67cc/Arguments-et-retour-
dexperiencesur-la-distribution-diode-stable-
autourdes-centrales-nucleaires-
francaises.pdf
83. Pol, Arjan; Barends, Thomas R. M.; Dietl,
Andreas; Khadem, Ahmad F.; Eygensteyn,
Jelle; Jetten, Mike S. M.; Op Den Camp,
Huub J. M. (2013). "Rare earth metals are
essential for methanotrophic life in volcanic
mudpots". Environmental Microbiology. 16
(1): 255–64. doi:10.1111/1462-
2920.12249 . PMID 24034209 .
84. Koribanics, N. M.; Tuorto, S. J.; Lopez-
Chiaffarelli, N.; McGuinness, L. R.;
Häggblom, M. M.; Williams, K. H.; Long, P.
E.; Kerkhof, L. J. (2015). "Spatial Distribution
of an Uranium-Respiring
Betaproteobacterium at the Rifle, CO Field
Research Site" . PLoS ONE. 10 (4):
e0123378.
doi:10.1371/journal.pone.0123378 .
PMC 4395306 . PMID 25874721 .
85. Irizarry, Lisandro (23 April 2014).
"Thyroid Hormone Toxicity" . Medscape.
WedMD LLC. Retrieved 2 May 2014.
86. Patrick, L. (2008). "Iodine: deficiency
and therapeutic considerations" (PDF).
Altern Med Rev. 13 (2): 116–27.
PMID 18590348 . Archived from the
original (PDF) on 31 May 2013.
87. "Dietary Reference Intakes (DRIs):
Recommended Intakes for Individuals,
Vitamins" . Institute of Medicine. 2004.
Archived from the original on 30 October
2009. Retrieved 9 June 2010.
88. United States National Research
Council (2000). Dietary Reference Intakes
for Vitamin A, Vitamin K, Arsenic, Boron,
Chromium, Copper, Iodine, Iron, Manganese,
Molybdenum, Nickel, Silicon, Vanadium, and
Zinc . National Academies Press. pp. 258–
259.
89. Venturi, S, Venturi. M (2009). "Iodine,
thymus, and immunity". Nutrition. 25 (9):
977–979. doi:10.1016/j.nut.2009.06.002 .
PMID 19647627 .
90. Ullberg, S.; Ewaldsson, B. (1964).
"Distribution of radio-iodine studied by
whole-body autoradiography". Acta
Radiologica Therapy Physics Biology. 41:
24–32. doi:10.3109/02841866409134127 .
91. Venturi, Sebastiano (2014). "Iodine,
PUFAs and Iodolipids in Health and Disease:
An Evolutionary Perspective". Human
Evolution-. 29 (1–3): 185–205. ISSN 0393-
9375 .
92. "Where do we get iodine from?" . Iodine
Global Network. Archived from the original
on 13 August 2015.
93. "Iodine in diet" . MedlinePlus Medical
Encyclopedia.
94. "American Thyroid Association" .
thyroid.org. American Thyroid Association.
Retrieved 4 April 2014.
95. Caldwell KL, Makhmudov A, Ely E, Jones
RL, Wang RY (2011). "Iodine status of the
U.S. population, National Health and
Nutrition Examination Survey, 2005–2006
and 2007–2008". Thyroid. 21 (4): 419–27.
doi:10.1089/thy.2010.0077 .
PMID 21323596 .
96. Leung AM, Braverman LE, Pearce EN
(2012). "History of U.S. iodine fortification
and supplementation" . Nutrients. 4 (11):
1740–6. doi:10.3390/nu4111740 .
PMC 3509517 . PMID 23201844 .
97. Zava, T. T.; Zava, D. T. (2011).
"Assessment of Japanese iodine intake
based on seaweed consumption in Japan:
A literature-based analysis" . Thyroid
Research. 4: 14. doi:10.1186/1756-6614-4-
14 . PMC 3204293 . PMID 21975053 .
98. Overview of Dietary Reference Intakes
for Japanese (2015) Minister of Health,
Labour and Welfare, Japan| url =
http://www.mhlw.go.jp/file/06-
Seisakujouhou-10900000-
Kenkoukyoku/Overview.pdf
99. Wu, T.; Liu, G. J.; Li, P.; Clar, C. (2002).
Wu, Taixiang, ed. "Iodised salt for preventing
iodine deficiency disorders". Cochrane
Database Syst Rev (3): CD003204.
doi:10.1002/14651858.CD003204 .
PMID 12137681 .
100. Dissanayake, C. B.; Chandrajith,
Rohana; Tobschall, H. J. (1999). "The iodine
cycle in the tropical environment —
implications on iodine deficiency disorders".
International Journal of Environmental
Studies. 56 (3): 357.
doi:10.1080/00207239908711210 .
101. Felig, Philip; Frohman, Lawrence A.
(2001). "Endemic Goiter" . Endocrinology &
metabolism. McGraw-Hill Professional.
ISBN 978-0-07-022001-0.
102. "Micronutrient deficiency: iodine
deficiency disorders" . WHO.
103. Zimmermann, Michael B.; Connolly, K.;
et al. (2006). "Iodine supplementation
improves cognition in iodine-deficient
schoolchildren in Albania: a randomized,
controlled, double-blind study" . American
Journal of Clinical Nutrition. 83 (1): 108–
114. doi:10.1093/ajcn/83.1.108 .
PMID 16400058 .
104.
https://www.sigmaaldrich.com/catalog/pro
duct/sigald/207772?lang=en&region=US
105. Technical data for Iodine .
periodictable.com
106. Smyth, P. P. (2003). "Role of iodine in
antioxidant defence in thyroid and breast
disease". BioFactors. 19 (3–4): 121–30.
doi:10.1002/biof.5520190304 .
PMID 14757962 .
107. Yerkes, Christine (2007). "Lecture 29:
Protein Structure and Denaturation" .
chem.uiuc.edu. University of Illinois.
Retrieved 23 October 2016.
108. Lowe, D. O.; Knowles, S. R.; et al.
(2006). "Povidone-iodine-induced burn: case
report and review of the literature".
Pharmacotherapy. 26 (11): 1641–5.
doi:10.1592/phco.26.11.1641 .
PMID 17064209 .
109. "CDC - NIOSH Pocket Guide to
Chemical Hazards - Iodine" . www.cdc.gov.
Retrieved 2015-11-06.
110. DermNet New Zealand Trust, Iodine
111. Boehm, I (2008). "Seafood allergy and
radiocontrast media: Are physicians
propagating a myth?". Am J Med. 121 (8):
E19. doi:10.1016/j.amjmed.2008.03.035 .
PMID 18691465 .
112. UCSF Department of Radiology &
Biomedical Imaging, Iodine Allergy and
Contrast Administration
113. Katelaris, Constance (2009). " 'Iodine
Allergy' label is misleading" . Australian
Prescriber. 32 (5): 125–128.
doi:10.18773/austprescr.2009.061 .
Archived from the original on 3 March
2016.

Bibliography
Greenwood, Norman N.; Earnshaw, Alan
(1997). Chemistry of the Elements (2nd
ed.). Butterworth-Heinemann. ISBN 978-
0-08-037941-8.
Iodine

Period 5 elements
Books
View or order
Halogens
collections of
articles
Chemical elements
(sorted alphabetically)

Chemical elements
(sorted by number)

Portals
Access Chemistry portal
related topics

Find out
more on Media
Wikipedia's from Commons
Sister Definitions
projects from Wiktionary

Textbooks
from Wikibooks

Quotations
from Wikiquote

Source texts
from Wikisource

Learning resources
from Wikiversity
Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Iodine&oldid=892841204"

Last edited 7 hours ago by Sethrc225

Content is available under CC BY-SA 3.0 unless


otherwise noted.

Das könnte Ihnen auch gefallen