You are on page 1of 93

Polonium

Polonium is a chemical element with the


symbol Po and atomic number 84. A rare
and highly radioactive metal with no stable
isotopes, polonium is chemically similar to
selenium and tellurium, though its metallic
character resembles that of its horizontal
neighbors in the periodic table: thallium,
lead, and bismuth. Due to the short half-
life of all its isotopes, its natural
occurrence is limited to tiny traces of the
fleeting polonium-210 (with a half-life of
138 days) in uranium ores, as it is the
penultimate daughter of natural uranium-
238. Though slightly longer-lived isotopes
exist, they are much more difficult to
produce. Today, polonium is usually
produced in milligram quantities by the
neutron irradiation of bismuth. Due to its
intense radioactivity, which results in the
radiolysis of chemical bonds and
radioactive self-heating, its chemistry has
mostly been investigated on the trace
scale only.
Polonium,  84Po

Polonium

Pronunciation /pəˈloʊniəm/ (pə-LOH-


nee-əm)

Allotropes α, β

Appearance silvery

Mass number 209 (most stable


isotope)

Polonium in the periodic table


Te

Po

Lv
bismuth ← polonium → astatine

Atomic number (Z) 84


Group group 16 (chalcogens)

Period period 6

Block p-block

Element category Post-transition metal,


 
but this status is
disputed

Electron configuration [Xe] 4f14 5d10 6s2 6p4

Electrons per shell 2, 8, 18, 32, 18, 6

Physical properties

Phase at STP solid

Melting point 527 K (254 °C, 489 °F)

Boiling point 1235 K (962 °C,


1764 °F)

Density (near r.t.) alpha: 9.196 g/cm3


beta: 9.398 g/cm3
Heat of fusion ca. 13 kJ/mol
Heat of vaporization 102.91 kJ/mol

Molar heat capacity 26.4 J/(mol·K)

Vapor pressure
P (Pa) 1 10 100 1 k 10 k 100 k

at T (K) (846) 1003 1236

Atomic properties

Oxidation states −2, +2, +4, +5,[1] +6


(an amphoteric oxide)

Electronegativity Pauling scale: 2.0

Ionization energies 1st: 812.1 kJ/mol

Atomic radius empirical: 168 pm

Covalent radius 140±4 pm

Van der Waals radius 197 pm


Spectral lines of polonium
Other properties

Natural occurrence from decay

Crystal structure cubic


α-Po

Crystal structure rhombohedral


β-Po

Thermal expansion 23.5 µm/(m·K)


(at 25 °C)

Thermal conductivity 20 W/(m·K) (?)

Electrical resistivity α: 0.40 µΩ·m (at 0 °C)

Magnetic ordering nonmagnetic

CAS Number 7440-08-6

History

Naming after Polonia, Latin for


Poland, homeland of
Marie Curie
Discovery Pierre and Marie Curie
(1898)

First isolation Willy Marckwald (1902)

Main isotopes of polonium


Isotope Abundance Half-life (t1/2) Decay mode Product

204
α Pb
208
Po syn 2.898 y
β+ 208
Bi

α 205Pb
209 [2]
Po syn 125.2 y
β+ 209
Bi

210 206
Po trace 138.376 d α Pb

Polonium was discovered in 1898 by Marie


and Pierre Curie, when it was extracted
from the uranium ore pitchblende and
identified solely by its strong radioactivity:
it was the first element to be so
discovered. Polonium was named after
Marie Curie's homeland of Poland.
Polonium has few applications, and those
are related to its radioactivity: heaters in
space probes, antistatic devices, sources
of neutrons and alpha particles, and
poison. It is a radioactive element, and
extremely dangerous to humans.

Characteristics
210Po is an alpha emitter that has a half-
life of 138.4 days; it decays directly to its
stable daughter isotope, 206Pb. A milligram
(5 curies) of 210Po emits about as many
alpha particles per second as 5 grams of
226Ra.[3] A few curies (1 curie equals
37 gigabecquerels, 1 Ci = 37 GBq) of 210Po
emit a blue glow which is caused by
ionisation of the surrounding air.

About one in 100,000 alpha emissions


causes an excitation in the nucleus which
then results in the emission of a gamma
ray with a maximum energy of 803 keV.[4][5]

Solid state form

The alpha form of solid polonium.


Polonium is a radioactive element that
exists in two metallic allotropes. The alpha
form is the only known example of a
simple cubic crystal structure in a single
atom basis at STP, with an edge length of
335.2 picometers; the beta form is
rhombohedral.[6][7][8] The structure of
polonium has been characterized by X-ray
diffraction[9][10] and electron diffraction.[11]

210Po (in common with 238Pu) has the


ability to become airborne with ease: if a
sample is heated in air to 55 °C (131 °F),
50% of it is vaporized in 45 hours to form
diatomic Po2 molecules, even though the
melting point of polonium is 254 °C
(489 °F) and its boiling point is 962 °C
(1,764 °F).[12][13][1] More than one
hypothesis exists for how polonium does
this; one suggestion is that small clusters
of polonium atoms are spalled off by the
alpha decay.

Chemistry

The chemistry of polonium is similar to


that of tellurium, although it also shows
some similarities to its neighbor bismuth
due to its metallic character. Polonium
dissolves readily in dilute acids but is only
slightly soluble in alkalis. Polonium
solutions are first colored in pink by the
Po2+ ions, but then rapidly become yellow
because alpha radiation from polonium
ionizes the solvent and converts Po2+ into
Po4+. This process is accompanied by
bubbling and emission of heat and light by
glassware due to the absorbed alpha
particles; as a result, polonium solutions
are volatile and will evaporate within days
unless sealed.[14][15] At pH about 1,
polonium ions are readily hydrolyzed and
complexed by acids such as oxalic acid,
citric acid, and tartaric acid.[16]

Compounds
Polonium has no common compounds,
and almost all of its compounds are
synthetically created; more than 50 of
those are known.[17] The most stable class
of polonium compounds are polonides,
which are prepared by direct reaction of
two elements. Na2Po has the antifluorite
structure, the polonides of Ca, Ba, Hg, Pb
and lanthanides form a NaCl lattice, BePo
and CdPo have the wurtzite and MgPo the
nickel arsenide structure. Most polonides
decompose upon heating to about 600 °C,
except for HgPo that decomposes at
~300 °C and the lanthanide polonides,
which do not decompose but melt at
temperatures above 1000 °C. For example,
PrPo melts at 1250 °C and TmPo at
2200 °C.[18] PbPo is one of the very few
naturally occurring polonium compounds,
as polonium alpha decays to form lead.[19]

Polonium hydride (PoH2) is a volatile liquid


at room temperature prone to dissociation;
it is thermally unstable.[18] Water is the
only other known hydrogen chalcogenide
which is a liquid at room temperature;
however, this is due to hydrogen bonding.
The two oxides PoO2 and PoO3 are the
products of oxidation of polonium.[20]

Halides of the structure PoX2, PoX4 and


PoF6 are known. They are soluble in the
corresponding hydrogen halides, i.e., PoClX
in HCl, PoBrX in HBr and PoI4 in HI.[21]
Polonium dihalides are formed by direct
reaction of the elements or by reduction of
PoCl4 with SO2 and with PoBr4 with H2S at
room temperature. Tetrahalides can be
obtained by reacting polonium dioxide
with HCl, HBr or HI.[22]

Other polonium compounds include


potassium polonite as a polonite,
polonate, acetate, bromate, carbonate,
citrate, chromate, cyanide, formate, (II) and
(IV) hydroxides, nitrate, selenate, selenite,
monosulfide, sulfate, disulfate and
sulfite.[21][23]
Polonium compounds[22][24]
m.p. Sublimation Pearson Space a
Formula Color Symmetry No b(pm) c(pm)
(°C) temp. (°C) symbol group (pm)

pale 500
PoO2 885 fcc cF12 Fm3m 225 563.7 563.7 563.7
yellow (dec.)

dark
PoCl2 ruby 355 130 orthorhombic oP3 Pmmm 47 367 435 450
red

purple- 270
PoBr2
brown (dec.)

PoCl4 yellow 300 200 monoclinic

330
PoBr4 red fcc cF100 Fm3m 225 560 560 560
(dec.)

PoI4 black
Oxides

PoO
PoO2
PoO3

Hydrides

PoH2

Halides

PoX2 (except PoF2)


PoX4
PoF6
PoBr2Cl2 (salmon pink)
Isotopes

Polonium has 42 known isotopes, all of


which are radioactive. They have atomic
masses that range from 186 to 227 u.
210Po (half-life 138.376 days) is the most
widely available and is made via neutron
capture by natural bismuth. The longer-
lived 209Po (half-life 125.2 ± 3.3 years,
longest-lived of all polonium isotopes)[2]
and 208Po (half-life 2.9 years) can be made
through the alpha, proton, or deuteron
bombardment of lead or bismuth in a
cyclotron.[29]
History
Tentatively called "radium F", polonium
was discovered by Marie and Pierre Curie
in 1898,[30][31] and was named after Marie
Curie's native land of Poland (Latin:
Polonia).[32][33] Poland at the time was
under Russian, German, and Austro-
Hungarian partition, and did not exist as
an independent country. It was Curie's
hope that naming the element after her
native land would publicize its lack of
independence.[34] Polonium may be the
first element named to highlight a political
controversy.[34]
This element was the first one discovered
by the Curies while they were investigating
the cause of pitchblende radioactivity.
Pitchblende, after removal of the
radioactive elements uranium and thorium,
was more radioactive than the uranium
and thorium combined. This spurred the
Curies to search for additional radioactive
elements. They first separated out
polonium from pitchblende in July 1898,
and five months later, also isolated
radium.[14][30][35] German scientist Willy
Marckwald successfully isolated 3
milligrams of polonium in 1902, though at
the time he believed it was a new element,
which he dubbed "radio-tellurium", and it
was not until 1905 that it was
demonstrated to be the same as
polonium.[36][37]

In the United States, polonium was


produced as part of the Manhattan
Project's Dayton Project during World War
II. It was a critical part of the implosion-
type nuclear weapon design used in the
Fat Man bomb on Nagasaki in 1945.
Polonium and beryllium were the key
ingredients of the 'urchin' detonator at the
center of the bomb's spherical plutonium
pit.[38] The urchin ignited the nuclear chain
reaction at the moment of prompt-
criticality to ensure the bomb did not
fizzle.[38]

Much of the basic physics of polonium


was classified until after the war. The fact
that it was used as an initiator was
classified until the 1960s.[39]

The Atomic Energy Commission and the


Manhattan Project funded human
experiments using polonium on five
people at the University of Rochester
between 1943 and 1947. The people were
administered between 9 and 22
microcuries (330 and 810 kBq) of
polonium to study its excretion.[40][41][42]
Occurrence and production
Polonium is a very rare element in nature
because of the short half-life of all its
isotopes. 210Po, 214Po, and 218Po appear in
the decay chain of 238U; thus polonium can
be found in uranium ores at about 0.1 mg
per metric ton (1 part in 1010),[43][44] which
is approximately 0.2% of the abundance of
radium. The amounts in the Earth's crust
are not harmful. Polonium has been found
in tobacco smoke from tobacco leaves
grown with phosphate fertilizers.[45][46][47]

Because it is present in small


concentrations, isolation of polonium from
natural sources is a tedious process. The
largest batch of the element ever
extracted, performed in the first half of the
20th century, contained only 40 Ci
(1.5 TBq) (9 mg) of polonium-210 and was
obtained by processing 37 tonnes of
residues from radium production.[48]
Polonium is now usually obtained by
irradiating bismuth with high-energy
neutrons or protons.[14][49]

In 1934, an experiment showed that when


natural 209Bi is bombarded with neutrons,
210Bi is created, which then decays to
210Po via beta-minus decay. The final
purification is done pyrochemically
followed by liquid-liquid extraction
techniques.[50] Polonium may now be
made in milligram amounts in this
procedure which uses high neutron fluxes
found in nuclear reactors.[49] Only about
100 grams are produced each year,
practically all of it in Russia, making
polonium exceedingly rare.[51][52]

This process can cause problems in lead-


bismuth based liquid metal cooled nuclear
reactors such as those used in the Soviet
Navy's K-27. Measures must be taken in
these reactors to deal with the unwanted
possibility of 210Po being released from
the coolant.[53][54]
The longer-lived isotopes of polonium,
208Po and 209Po, can be formed by proton
or deuteron bombardment of bismuth
using a cyclotron. Other more neutron-
deficient and more unstable isotopes can
be formed by the irradiation of platinum
with carbon nuclei.[55]

Applications
Polonium-based sources of alpha particles
were produced in the former Soviet
Union.[56] Such sources were applied for
measuring the thickness of industrial
coatings via attenuation of alpha
radiation.[57]
Because of intense alpha radiation, a one-
gram sample of 210Po will spontaneously
heat up to above 500 °C (932 °F)
generating about 140 watts of power.
Therefore, 210Po is used as an atomic heat
source to power radioisotope
thermoelectric generators via
thermoelectric materials.[3][14][58][59] For
example, 210Po heat sources were used in
the Lunokhod 1 (1970) and Lunokhod 2
(1973) Moon rovers to keep their internal
components warm during the lunar nights,
as well as the Kosmos 84 and 90 satellites
(1965).[56][60]
The alpha particles emitted by polonium
can be converted to neutrons using
beryllium oxide, at a rate of 93 neutrons
per million alpha particles.[58] Thus Po-BeO
mixtures or alloys are used as a neutron
source, for example, in a neutron trigger or
initiator for nuclear weapons[14][61] and for
inspections of oil wells. About 1500
sources of this type, with an individual
activity of 1,850 Ci (68 TBq), have been
used annually in the Soviet Union.[62]

Polonium was also part of brushes or


more complex tools that eliminate static
charges in photographic plates, textile
mills, paper rolls, sheet plastics, and on
substrates (such as automotive) prior to
the application of coatings.[63] Alpha
particles emitted by polonium ionize air
molecules that neutralize charges on the
nearby surfaces.[64][65] Some anti-static
brushes contain up to 500 microcuries
(20 MBq) of 210Po as a source of charged
particles for neutralizing static
electricity.[66] In the US, the devices with no
more than 500 μCi (19 MBq) of (sealed)
210Po per unit can be bought in any
amount under a "general license",[67] which
means that a buyer need not be registered
by any authorities. Polonium needs to be
replaced in these devices nearly every year
because of its short half-life; it is also
highly radioactive and therefore has been
mostly replaced by less dangerous beta
particle sources.[3]

Tiny amounts of 210Po are sometimes


used in the laboratory and for teaching
purposes—typically of the order of 4–
40 kBq (0.11–1.08 μCi), in the form of
sealed sources, with the polonium
deposited on a substrate or in a resin or
polymer matrix—are often exempt from
licensing by the NRC and similar
authorities as they are not considered
hazardous. Small amounts of 210Po are
manufactured for sale to the public in the
United States as 'needle sources' for
laboratory experimentation, and they are
retailed by scientific supply companies.
The polonium is a layer of plating which in
turn is plated with a material such as gold,
which allows the alpha radiation (used in
experiments such as cloud chambers) to
pass while preventing the polonium from
being released and presenting a toxic
hazard. According to United Nuclear, they
typically sell between four and eight such
sources per year.[68][69]

Polonium spark plugs were marketed by


Firestone from 1940 to 1953. While the
amount of radiation from the plugs was
minuscule and not a threat to the
consumer, the benefits of such plugs
quickly diminished after approximately a
month because of polonium's short half-
life and because buildup on the
conductors would block the radiation that
improved engine performance. (The
premise behind the polonium spark plug,
as well as Alfred Matthew Hubbard's
prototype radium plug that preceded it,
was that the radiation would improve
ionization of the fuel in the cylinder and
thus allow the motor to fire more quickly
and efficiently.)[70][71]

Biology and toxicity


Overview

Polonium is highly dangerous and has no


biological role.[14] By mass, polonium-210
is around 250,000 times more toxic than
hydrogen cyanide (the LD50 for 210Po is
less than 1 microgram for an average
adult (see below) compared with about
250 milligrams for hydrogen cyanide[72]).
The main hazard is its intense radioactivity
(as an alpha emitter), which makes it
difficult to handle safely. Even in
microgram amounts, handling 210Po is
extremely dangerous, requiring specialized
equipment (a negative pressure alpha
glove box equipped with high-performance
filters), adequate monitoring, and strict
handling procedures to avoid any
contamination. Alpha particles emitted by
polonium will damage organic tissue
easily if polonium is ingested, inhaled, or
absorbed, although they do not penetrate
the epidermis and hence are not
hazardous as long as the alpha particles
remain outside the body. Wearing
chemically resistant and intact gloves is a
mandatory precaution to avoid
transcutaneous diffusion of polonium
directly through the skin. Polonium
delivered in concentrated nitric acid can
easily diffuse through inadequate gloves
(e.g., latex gloves) or the acid may damage
the gloves.[73]

It has been reported that some microbes


can methylate polonium by the action of
methylcobalamin.[74][75] This is similar to
the way in which mercury, selenium, and
tellurium are methylated in living things to
create organometallic compounds.
Studies investigating the metabolism of
polonium-210 in rats have shown that only
0.002 to 0.009% of polonium-210 ingested
is excreted as volatile polonium-210.[76]

Acute effects
The median lethal dose (LD50) for acute
radiation exposure is about 4.5 Sv.[77] The
committed effective dose equivalent 210Po
is 0.51 µSv/Bq if ingested, and 2.5 µSv/Bq
if inhaled.[78] So a fatal 4.5 Sv dose can be
caused by ingesting 8.8 MBq (240 μCi),
about 50 nanograms (ng), or inhaling
1.8 MBq (49 μCi), about 10 ng. One gram
of 210Po could thus in theory poison 20
million people of whom 10 million would
die. The actual toxicity of 210Po is lower
than these estimates because radiation
exposure that is spread out over several
weeks (the biological half-life of polonium
in humans is 30 to 50 days[79]) is
somewhat less damaging than an
instantaneous dose. It has been estimated
that a median lethal dose of 210Po is 15
megabecquerels (0.41 mCi), or 0.089
micrograms (μg), still an extremely small
amount.[80][81] For comparison, one grain
of table salt is about 0.06 mg = 60 μg.[82]

Long term (chronic) effects

In addition to the acute effects, radiation


exposure (both internal and external)
carries a long-term risk of death from
cancer of 5–10% per Sv.[77] The general
population is exposed to small amounts of
polonium as a radon daughter in indoor air;
the isotopes 214Po and 218Po are thought
to cause the majority[83] of the estimated
15,000–22,000 lung cancer deaths in the
US every year that have been attributed to
indoor radon.[84] Tobacco smoking causes
additional exposure to polonium.[85]

Regulatory exposure limits and


handling

The maximum allowable body burden for


ingested 210Po is only 1.1 kBq (30 nCi),
which is equivalent to a particle massing
only 6.8 picograms. The maximum
permissible workplace concentration of
airborne 210Po is about 10 Bq/m3
(3 × 10−10 µCi/cm3).[86] The target organs
for polonium in humans are the spleen and
liver.[87] As the spleen (150 g) and the liver
(1.3 to 3 kg) are much smaller than the
rest of the body, if the polonium is
concentrated in these vital organs, it is a
greater threat to life than the dose which
would be suffered (on average) by the
whole body if it were spread evenly
throughout the body, in the same way as
caesium or tritium (as T2O).

210Po is widely used in industry, and


readily available with little regulation or
restriction.[88] In the US, a tracking system
run by the Nuclear Regulatory Commission
was implemented in 2007 to register
purchases of more than 16 curies
(590 GBq) of polonium-210 (enough to
make up 5,000 lethal doses). The IAEA "is
said to be considering tighter
regulations ... There is talk that it might
tighten the polonium reporting
requirement by a factor of 10, to 1.6 curies
(59 GBq)."[89] As of 2013, this is still the
only alpha emitting byproduct material
available, as a NRC Exempt Quantity,
which may be held without a radioactive
material license.

Polonium and its compounds must be


handled in a glove box, which is further
enclosed in another box, maintained at a
slightly higher pressure than the glove box
to prevent the radioactive materials from
leaking out. Gloves made of natural rubber
do not provide sufficient protection
against the radiation from polonium;
surgical gloves are necessary. Neoprene
gloves shield radiation from polonium
better than natural rubber.[90]

Cases of poisoning

20th century

Polonium was administered to humans for


experimental purposes from 1943 to 1947;
it was injected into four hospitalised
patients, and orally given to a fifth. Studies
such as this were funded by the
Manhattan Project and the AEC and
conducted at the University of Rochester.
The objective was to obtain data on
human excretion of polonium to correlate
with more extensive data from rats.
Patients selected as subjects were chosen
because experimenters wanted persons
who had not been exposed to polonium
either through work or accident. All
subjects had incurable diseases. Excretion
of polonium was followed, and an autopsy
was conducted at that time on the
deceased patient to determine which
organs absorbed the polonium. Patients'
ages ranged from "early thirties" to "early
forties". The experiments were described
in Chapter 3 of Biological Studies with
Polonium, Radium, and Plutonium,
National Nuclear Energy Series, Volume VI-
3, McGraw-Hill, New York, 1950. Not
specified is the isotope under study, but at
the time polonium-210 was the most
readily available polonium isotope. The
DoE factsheet submitted for this
experiment reported no follow up on these
subjects.[91]

It has also been suggested that Irène


Joliot-Curie was the first person to die
from the radiation effects of polonium.
She was accidentally exposed to polonium
in 1946 when a sealed capsule of the
element exploded on her laboratory bench.
In 1956, she died from leukemia.[92]

According to the 2008 book The Bomb in


the Basement, several deaths in Israel
during 1957–1969 were caused by
210Po.[93] A leak was discovered at a
Weizmann Institute laboratory in 1957.
Traces of 210Po were found on the hands
of Professor Dror Sadeh, a physicist who
researched radioactive materials. Medical
tests indicated no harm, but the tests did
not include bone marrow. Sadeh died from
cancer. One of his students died of
leukemia, and two colleagues died after a
few years, both from cancer. The issue
was investigated secretly, and there was
never any formal admission that a
connection between the leak and the
deaths had existed.[94]

21st century

The cause of death in the 2006 homicide


of Alexander Litvinenko, a Russian KGB
agent who had defected to the British MI6
intelligence agency, was determined to be
210Po poisoning.[95][96] According to Prof.
Nick Priest of Middlesex University, an
environmental toxicologist and radiation
expert, speaking on Sky News on
December 3, 2006, Litvinenko was
probably the first person to die of the
acute α-radiation effects of 210Po.[97]

Abnormally high concentrations of 210Po


were detected in July 2012 in clothes and
personal belongings of the Palestinian
leader Yasser Arafat, a heavy smoker, who
died on 11 November 2004 of uncertain
causes. The spokesman for the Institut de
Radiophysique in Lausanne, Switzerland,
where those items were analyzed,
stressed that the "clinical symptoms
described in Arafat's medical reports were
not consistent with polonium-210 and that
conclusions could not be drawn as to
whether the Palestinian leader was
poisoned or not", and that "the only way to
confirm the findings would be to exhume
Arafat's body to test it for polonium-
210."[98] On 27 November 2012 Arafat's
body was exhumed, and samples were
taken for separate analysis by experts
from France, Switzerland and Russia.[99]
On 12 October 2013, The Lancet published
the group's finding that high levels of the
element were found in Arafat's blood,
urine, and in saliva stains on his clothes
and toothbrush.[100] The French tests later
found some polonium but stated it was
from "natural environmental origin".[101]
Following later Russian tests, Vladimir
Uiba, the head of the Russian Federal
Medical and Biological Agency, stated in
December 2013 that Arafat died of natural
causes, and they had no plans to conduct
further tests.[101]

Treatment

It has been suggested that chelation


agents, such as British Anti-Lewisite
(dimercaprol), can be used to
decontaminate humans.[102] In one
experiment, rats were given a fatal dose of
1.45 MBq/kg (8.7 ng/kg) of 210Po; all
untreated rats were dead after 44 days, but
90% of the rats treated with the chelation
agent HOEtTTC remained alive for 5
months.[103]

Detection in biological
specimens

Polonium-210 may be quantified in


biological specimens by alpha particle
spectrometry to confirm a diagnosis of
poisoning in hospitalized patients or to
provide evidence in a medicolegal death
investigation. The baseline urinary
excretion of polonium-210 in healthy
persons due to routine exposure to
environmental sources is normally in a
range of 5–15 mBq/day. Levels in excess
of 30 mBq/day are suggestive of
excessive exposure to the
radionuclide.[104]

Occurrence in humans and the


biosphere

Polonium-210 is widespread in the


biosphere, including in human tissues,
because of its position in the uranium-238
decay chain. Natural uranium-238 in the
Earth's crust decays through a series of
solid radioactive intermediates including
radium-226 to the radioactive noble gas
radon-222, some of which, during its 3.8-
day half-life, diffuses into the atmosphere.
There it decays through several more
steps to polonium-210, much of which,
during its 138-day half-life, is washed back
down to the Earth's surface, thus entering
the biosphere, before finally decaying to
stable lead-206.[105][106][107]

As early as the 1920s Antoine


Lacassagne, using polonium provided by
his colleague Marie Curie, showed that the
element has a specific pattern of uptake in
rabbit tissues, with high concentrations,
particularly in liver, kidney, and testes.[108]
More recent evidence suggests that this
behavior results from polonium
substituting for its congener sulfur, also in
group 16 of the periodic table, in sulfur-
containing amino-acids or related
molecules[109] and that similar patterns of
distribution occur in human tissues.[110]
Polonium is indeed an element naturally
present in all humans, contributing
appreciably to natural background dose,
with wide geographical and cultural
variations, and particularly high levels in
arctic residents, for example.[111]

Tobacco

Polonium-210 in tobacco contributes to


many of the cases of lung cancer
worldwide. Most of this polonium is
derived from lead-210 deposited on
tobacco leaves from the atmosphere; the
lead-210 is a product of radon-222 gas,
much of which appears to originate from
the decay of radium-226 from fertilizers
applied to the tobacco
soils.[47][112][113][114][115]

The presence of polonium in tobacco


smoke has been known since the early
1960s.[116][117] Some of the world's biggest
tobacco firms researched ways to remove
the substance—to no avail—over a 40-year
period. The results were never
published.[47]
Food

Polonium is found in the food chain,


especially in seafood.[118][119]

See also
Polonium halo

References
1. Thayer, John S. (2010). "Relativistic
Effects and the Chemistry of the
Heavier Main Group Elements".
Relativistic Methods for Chemists: 78.
doi:10.1007/978-1-4020-9975-5_2 .
ISBN 978-1-4020-9974-8.
2. Boutin, Chad. "Polonium's Most Stable
Isotope Gets Revised Half-Life
Measurement" . nist.gov. NIST Tech
Beat. Retrieved 9 September 2014.
3. "Polonium" (PDF). Argonne National
Laboratory. Archived from the original
(PDF) on 2007-07-03. Retrieved
2009-05-05.
4. Greenwood, p. 250
5. "210PO α decay" . Nuclear Data
Center, Korea Atomic Energy Research
Institute. 2000. Retrieved 2009-05-05.
6. Greenwood, p. 753
7. Miessler, Gary L.; Tarr, Donald A.
(2004). Inorganic Chemistry (3rd ed.).
Upper Saddle River, N.J.: Pearson
Prentice Hall. p. 285. ISBN 978-0-13-
120198-9.
8. "The beta Po (A_i) Structure" . Naval
Research Laboratory. 2000-11-20.
Archived from the original on 2001-
02-04. Retrieved 2009-05-05.
9. Desando, R. J.; Lange, R. C. (1966).
"The structures of polonium and its
compounds—I α and β polonium
metal". Journal of Inorganic and
Nuclear Chemistry. 28 (9): 1837–
1846. doi:10.1016/0022-
1902(66)80270-1 .
10. Beamer, W. H.; Maxwell, C. R. (1946).
"The Crystal Structure of Polonium".
Journal of Chemical Physics. 14 (9):
569. doi:10.1063/1.1724201 .
11. Rollier, M. A.; Hendricks, S. B.; Maxwell,
L. R. (1936). "The Crystal Structure of
Polonium by Electron Diffraction".
Journal of Chemical Physics. 4 (10):
648. Bibcode:1936JChPh...4..648R .
doi:10.1063/1.1749762 .
12. Wąs, Bogdan; Misiak, Ryszard;
Bartyzel, Mirosław; Petelenz, Barbara
(2006). "Thermochromatographic
Separation of 206,208Po from a Bismuth
Target Bombardet with Protons"
(PDF). Nukleonika. 51 (Suppl. 2): s3–
s5.
13. Lide, D. R., ed. (2005). CRC Handbook
of Chemistry and Physics (86th ed.).
Boca Raton (FL): CRC Press. ISBN 0-
8493-0486-5.
14. Emsley, John (2001). Nature's Building
Blocks. New York: Oxford University
Press. pp. 330–332. ISBN 978-0-19-
850341-5.
15. Bagnall, p. 206
16. Keller, Cornelius; Wolf, Walter; Shani,
Jashovam, "Radionuclides, 2.
Radioactive Elements and Artificial
Radionuclides", Ullmann's
Encyclopedia of Industrial Chemistry,
Weinheim: Wiley-VCH,
doi:10.1002/14356007.o22_o15
17. Bagnall, p. 199
18. Greenwood, p. 766
19. Weigel, F. (1959). "Chemie des
Poloniums". Angewandte Chemie. 71
(9): 289–316.
doi:10.1002/ange.19590710902 .
20. Holleman, A. F.; Wiberg, E. (2001).
Inorganic Chemistry. San Diego:
Academic Press. ISBN 978-0-12-
352651-9.
21. Figgins, P. E. (1961) The
Radiochemistry of Polonium , National
Academy of Sciences, US Atomic
Energy Commission, pp. 13–14
Google Books
22. Greenwood, pp. 765, 771, 775
23. Bagnall, pp. 212–226
24. Wiberg, Egon; Holleman, A. F. and
Wiberg, Nils Inorganic Chemistry ,
Academic Press, 2001, p. 594, ISBN 0-
12-352651-5.
25. Bagnall, K. W.; d'Eye, R. W. M. (1954).
"The Preparation of Polonium Metal
and Polonium Dioxide". J. Chem. Soc.:
4295–4299.
doi:10.1039/JR9540004295 .
26. Bagnall, K. W.; d'Eye, R. W. M.;
Freeman, J. H. (1955). "The polonium
halides. Part I. Polonium chlorides".
Journal of the Chemical Society
(Resumed): 2320.
doi:10.1039/JR9550002320 .
27. Bagnall, K. W.; d'Eye, R. W. M.;
Freeman, J. H. (1955). "The polonium
halides. Part II. Bromides". Journal of
the Chemical Society (Resumed):
3959. doi:10.1039/JR9550003959 .
28. Bagnall, K. W.; d'Eye, R. W. M.;
Freeman, J. H. (1956). "657. The
polonium halides. Part III. Polonium
tetraiodide". Journal of the Chemical
Society (Resumed): 3385.
doi:10.1039/JR9560003385 .
29. Emsley, John (2011). Nature's Building
Blocks: An A-Z Guide to the Elements
(New ed.). New York, NY: Oxford
University Press. p. 415. ISBN 978-0-
19-960563-7.
30. Curie, P.; Curie, M. (1898). "Sur une
substance nouvelle radio-active,
contenue dans la pechblende" [On a
new radioactive substance contained
in pitchblende] (PDF). Comptes
Rendus (in French). 127: 175–178.
Archived from the original on July 23,
2013. English translation.
31. Krogt, Peter van der. "84. Polonium -
Elementymology & Elements
Multidict" . elements.vanderkrogt.net.
Retrieved 2017-04-26.
32. Pfützner, M. (1999). "Borders of the
Nuclear World – 100 Years After
Discovery of Polonium". Acta Physica
Polonica B. 30 (5): 1197.
Bibcode:1999AcPPB..30.1197P .
33. Adloff, J. P. (2003). "The centennial of
the 1903 Nobel Prize for physics".
Radichimica Acta. 91 (12–2003): 681–
688.
doi:10.1524/ract.91.12.681.23428 .
34. Kabzinska, K. (1998). "Chemical and
Polish aspects of polonium and
radium discovery". Przemysł
Chemiczny. 77 (3): 104–107.
35. Curie, P.; Curie, M.; Bémont, G. (1898).
"Sur une nouvelle substance fortement
radio-active contenue dans la
pechblende" [On a new, strongly
radioactive substance contained in
pitchblende] (PDF). Comptes Rendus
(in French). 127: 1215–1217. Archived
from the original on July 22, 2013.
English translation
36. "Polonium and Radio-Tellurium".
Nature. 73 (549): 549. 1906.
Bibcode:1906Natur..73R.549. .
doi:10.1038/073549b0 .
37. Neufeldt, Sieghard (2012). Chronologie
Chemie: Entdecker und
Entdeckungen . John Wiley & Sons.
ISBN 9783527662845.
38. Nuclear Weapons FAQ, Section 4.1,
Version 2.04: 20 February 1999 .
Nuclearweaponarchive.org. Retrieved
on 2013-04-28.
39. RESTRICTED DATA
DECLASSIFICATION DECISIONS, 1946
TO THE PRESENT (RDD-7) , January 1,
2001, U.S. Department of Energy
Office of Declassification, via fas.org
40. American nuclear guinea pigs: three
decades of radiation experiments on
U.S. citizens . United States.
Congress. House. of the Committee
on Energy and Commerce.
Subcommittee on Energy
Conservation and Power, published by
U.S. Government Printing Office, 1986,
Identifier Y 4.En 2/3:99-NN, Electronic
Publication Date 2010, at the
University of Nevada, Reno, unr.edu
41. "Studies of polonium metabolism in
human subjects", Chapter 3 in
Biological Studies with Polonium,
Radium, and Plutonium, National,
Nuclear Energy Series, Volume VI-3,
McGraw-Hill, New York, 1950, cited in
"American Nuclear Guinea Pigs ...",
1986 House Energy and Commerce
committee report
42. Moss, William and Eckhardt, Roger
(1995) "The Human Plutonium
Injection Experiments" , Los Alamos
Science, Number 23.
43. Greenwood, p. 746
44. Bagnall, p. 198
45. Kilthau, Gustave F. (1996). "Cancer risk
in relation to radioactivity in tobacco".
Radiologic Technology. 67 (3): 217–
222. PMID 8850254 .
46. "Alpha Radioactivity (210 Polonium)
and Tobacco Smoke" . Archived from
the original on June 9, 2013.
Retrieved 2009-05-05.
47. Monique, E. Muggli; Ebbert, Jon O.;
Robertson, Channing; Hurt, Richard D.
(2008). "Waking a Sleeping Giant: The
Tobacco Industry's Response to the
Polonium-210 Issue" . American
Journal of Public Health. 98 (9):
1643–50.
doi:10.2105/AJPH.2007.130963 .
PMC 2509609 . PMID 18633078 .
48. Adloff, J. P. & MacCordick, H. J.
(1995). "The Dawn of
Radiochemistry" . Radiochimica Acta.
70/71: 13–22.
doi:10.1524/ract.1995.7071.special-
issue.13 ., reprinted in Adloff, J. P.
(1996). One hundred years after the
discovery of radioactivity . p. 17.
ISBN 978-3-486-64252-0.
49. Greenwood, p. 249
50. Schulz, Wallace W.; Schiefelbein, Gary
F.; Bruns, Lester E. (1969).
"Pyrochemical Extraction of Polonium
from Irradiated Bismuth Metal". Ind.
Eng. Chem. Process Des. Dev. 8 (4):
508–515. doi:10.1021/i260032a013 .
51. "Q&A: Polonium-210" . RSC Chemistry
World. 2006-11-27. Retrieved
2009-01-12.
52. "Most Polonium Made Near the Volga
River" . The St. Petersburg Times –
News. 2001-01-23.
53. Usanov, V. I.; Pankratov, D. V.; Popov, É.
P.; Markelov, P. I.; Ryabaya, L. D.;
Zabrodskaya, S. V. (1999). "Long-lived
radionuclides of sodium, lead-bismuth,
and lead coolants in fast-neutron
reactors". Atomic Energy. 87 (3): 658–
662. doi:10.1007/BF02673579 .
54. Naumov, V. V. (November 2006). За
какими корабельными реакторами
будущее? . Атомная стратегия (in
Russian). 26.
55. Atterling, H.; Forsling, W. (1959). "Light
Polonium Isotopes from Carbon Ion
Bombardments of Platinum". Arkiv för
Fysik. 15 (1): 81–88. OSTI 4238755 .
56. "Радиоизотопные источники
тепла" . Archived from the original on
May 1, 2007. Retrieved June 1, 2016.
(in Russian). npc.sarov.ru
57. Bagnall, p. 225
58. Greenwood, p. 251
59. Hanslmeier, Arnold (2002). The sun
and space weather . Springer. p. 183.
ISBN 978-1-4020-0684-5.
60. Wilson, Andrew (1987). Solar System
Log. London: Jane's Publishing
Company Ltd. p. 64. ISBN 978-0-7106-
0444-6.
61. Rhodes, Richard (2002). Dark Sun: The
Making of the Hydrogen Bomb. New
York: Walker & Company. pp. 187–188.
ISBN 978-0-684-80400-2.
62. Красивая версия "самоубийства"
Литвиненко вследствие
криворукости (in Russian).
stringer.ru (2006-11-26).
63. Boice, John D.; Cohen, Sarah S.; et al.
(2014). "Mortality Among Mound
Workers Exposed to Polonium-210 and
Other Sources of Radiation, 1944–
1979". Radiation Research. 181 (2):
208–28.
Bibcode:2014RadR..181..208B .
doi:10.1667/RR13395.1 . ISSN 0033-
7587 . PMID 24527690 .
64. "Static Control for Electronic Balance
Systems" (PDF). Archived from the
original (PDF) on November 10, 2013.
Retrieved 2009-05-05.
65. "BBC News : College breaches
radioactive regulations" . 2002-03-12.
Retrieved 2009-05-05.
66. "Staticmaster Ionizing Brushes" .
AMSTAT Industries. Retrieved
2009-05-05.
67. "General domestic licenses for
byproduct material" . Retrieved
2009-05-05.
68. Singleton, Don (2006-11-28). "The
Availability of polonium-210" .
Retrieved 2006-11-29.
69. "Radioactive Isotopes" . United
Nuclear. Retrieved 2007-03-19.
70. "Radioactive spark plugs" . Oak Ridge
Associated Universities. January 20,
1999. Retrieved August 23, 2018.
71. Pittman, Cassandra (February 3,
2017). "Polonium" . The
Instrumentation Center. University of
Toledo. Retrieved August 23, 2018.
72. "Safety data for hydrogen cyanide" .
Physical & Theoretical Chemistry Lab,
Oxford University. Archived from the
original on 2002-02-11.
73. Bagnall, pp. 202–6
74. Momoshima, N.; Song, L. X.; Osaki, S.;
Maeda, Y. (2001). "Formation and
emission of volatile polonium
compound by microbial activity and
polonium methylation with
methylcobalamin". Environ Sci
Technol. 35 (15): 2956–2960.
doi:10.1021/es001730 .
75. Momoshima, N.; Song, L. X.; Osaki, S.;
Maeda, Y. (2002). "Biologically induced
Po emission from fresh water". J
Environ Radioact. 63 (2): 187–197.
doi:10.1016/S0265-931X(02)00028-0 .
PMID 12363270 .
76. Li, Chunsheng; Sadi, Baki; Wyatt,
Heather; Bugden, Michelle; et al.
(2010). "Metabolism of 210Po in rats:
volatile 210Po in excreta" . Radiation
Protection Dosimetry. 140 (2): 158–
162. doi:10.1093/rpd/ncq047 .
PMID 20159915 . Retrieved
2013-04-09.
77. "Health Impacts from Acute Radiation
Exposure" (PDF). Pacific Northwest
National Laboratory. Retrieved
2009-05-05.
78. "Nuclide Safety Data Sheet: Polonium–
210" (PDF). hpschapters.org.
Retrieved 2009-05-05.
79. Naimark, D.H. (1949-01-04). "Effective
half-life of polonium in the human".
Technical Report MLM-272/XAB,
Mound Lab., Miamisburg, OH.
OSTI 7162390 .
80. Carey Sublette (2006-12-14).
"Polonium Poisoning" . Retrieved
2009-05-05.
81. Harrison, J.; Leggett, Rich; Lloyd,
David; Phipps, Alan; et al. (2007).
"Polonium-210 as a poison". J. Radiol.
Prot. 27 (1): 17–40.
Bibcode:2007JRP....27...17H .
doi:10.1088/0952-4746/27/1/001 .
PMID 17341802 . "The conclusion is
reached that 0.1–0.3 GBq or more
absorbed to blood of an adult male is
likely to be fatal within 1 month. This
corresponds to ingestion of 1–3 GBq
or more, assuming 10% absorption to
blood"
82. Yasar Safkan. "Approximately how
many atoms are in a grain of salt?" .
PhysLink.com: Physics & Astronomy.
83. Health Risks of Radon and Other
Internally Deposited Alpha-Emitters:
BEIR IV. National Academy Press.
1988. p. 5. ISBN 978-0-309-03789-1.
84. Health Effects Of Exposure To Indoor
Radon . Washington: National
Academy Press. 1999. Archived from
the original on 2006-09-19.
85. "The Straight Dope: Does smoking
organically grown tobacco lower the
chance of lung cancer?" . 2007-09-28.
Retrieved 2009-05-05.
86. "Nuclear Regulatory Commission
limits for 210Po" . U.S. NRC. 2008-12-
12. Retrieved 2009-01-12.
87. "PilgrimWatch – Pilgrim Nuclear –
Health Impact" . Archived from the
original on 2009-01-05. Retrieved
2009-05-05.
88. Bastian, R.K.; Bachmaier, J.T.; Schmidt,
D.W.; Salomon, S.N.; Jones, A.; Chiu,
W.A.; Setlow, L.W.; Wolbarst, A.W.; Yu,
C. (2004-01-01). "Radioactive
Materials in Biosolids: National Survey,
Dose Modeling & POTW Guidance" .
Proceedings of the Water Environment
Federation. 2004 (1): 777–803.
doi:10.2175/193864704784343063 .
ISSN 1938-6478 .
89. Zimmerman, Peter D. (2006-12-19).
"The Smoky Bomb Threat" . The New
York Times. Retrieved 2006-12-19.
90. Bagnall, p. 204
91. "American nuclear guinea pigs : three
decades of radiation experiments on
U.S. citizens" . Retrieved 2015-06-09.
92. Manier, Jeremy (2006-12-04).
"Innocent chemical a killer" . The Daily
Telegraph (Australia). Archived from
the original on January 6, 2009.
Retrieved 2009-05-05.
93. Karpin, Michael (2006). The bomb in
the basement: How Israel went nuclear
and what that means for the world.
Simon and Schuster. ISBN 978-0-7432-
6594-2.
94. Maugh, Thomas; Karen Kaplan (2007-
01-01). "A restless killer radiates
intrigue" . Los Angeles Times.
Retrieved 2008-09-17.
95. Geoghegan, Tom (2006-11-24). "The
mystery of Litvinenko's death" . BBC
News.
96. "UK requests Lugovoi extradition" .
BBC News. 2007-05-28. Retrieved
2009-05-05.
97. Watson, Roland (2006-12-03). "Focus:
Cracking the code of the nuclear
assassin" . The Sunday Times.
London. Archived from the original on
2008-02-10. Retrieved 2010-05-22.
98. Bart, Katharina (2012-07-03). Swiss
institute finds polonium in Arafat's
effects . Reuters.
99. "Experts exhume Arafat, seek evidence
of poison" . Reuters. 2012-11-27.
Retrieved 2012-11-27.
100. Froidevaux, P.; Baechler, S. B.; Bailat, C.
J.; Castella, V.; Augsburger, M.;
Michaud, K.; Mangin, P.; Bochud, F. O.
O. (2013). "Improving forensic
investigation for polonium poisoning".
The Lancet. 382 (9900): 1308.
doi:10.1016/S0140-6736(13)61834-6 .
PMID 24120205 .
101. Isachenkov, Vadim (2013-12-27)
Russia: Arafat's death not caused by
radiation . Associated Press.
102. "Guidance for Industry. Internal
Radioactive Contamination —
Development of Decorporation
Agents" (PDF). US Food and Drug
Administration. Retrieved 2009-07-07.
103. Rencováa J.; Svoboda V.; Holuša R.;
Volf V.; et al. (1997). "Reduction of
subacute lethal radiotoxicity of
polonium-210 in rats by chelating
agents". International Journal of
Radiation Biology. 72 (3): 247.
doi:10.1080/095530097143338 .
104. Baselt, R. Disposition of Toxic Drugs
and Chemicals in Man , 10th edition,
Biomedical Publications, Seal Beach,
CA.
105. Hill, C. R. (1960). "Lead-210 and
Polonium-210 in Grass". Nature. 187
(4733): 211–212.
Bibcode:1960Natur.187..211H .
doi:10.1038/187211a0 .
106. Hill, C. R. (1963). "Natural occurrence
of unsupported radium-F (Po-210) in
tissue". Health Physics. 9: 952–953.
PMID 14061910 .
107. Heyraud, M.; Cherry, R. D. (1979).
"Polonium-210 and lead-210 in marine
food chains". Marine Biology. 52 (3):
227–236. doi:10.1007/BF00398136 .
108. Lacassagne, A. & Lattes, J. (1924)
Bulletin d'Histologie Appliquée à la
Physiologie et à la Pathologie, 1, 279.
109. Vasken Aposhian, H.; Bruce, D. C.
(1991). "Binding of Polonium-210 to
Liver Metallothionein". Radiation
Research. 126 (3): 379–382.
doi:10.2307/3577929 .
JSTOR 3577929 . PMID 2034794 .
110. Hill, C. R. (1965). "Polonium-210 in
man". Nature. 208 (5009): 423–8.
Bibcode:1965Natur.208..423H .
doi:10.1038/208423a0 .
PMID 5867584 .
111. Hill, C. R. (1966). "Polonium-210
Content of Human Tissues in Relation
to Dietary Habit". Science. 152 (3726):
1261–2.
Bibcode:1966Sci...152.1261H .
doi:10.1126/science.152.3726.1261 .
PMID 5949242 .
112. Martell, E. A. (1974). "Radioactivity of
tobacco trichomes and insoluble
cigarette smoke particles" . Nature.
249 (5454): 214–217.
Bibcode:1974Natur.249..215M .
doi:10.1038/249215a0 . Retrieved
20 July 2014.
113. Martell, E. A. (1975). "Tobacco
Radioactivity and Cancer in Smokers:
Alpha interactions with chromosomes
of cells surrounding insoluble
radioactive smoke particles may
cause cancer and contribute to early
atherosclerosis development in
cigarette smokers". American
Scientist. 63 (4): 404–412.
Bibcode:1975AmSci..63..404M .
JSTOR 27845575 . PMID 1137236 .
114. Tidd, M. J. (2008). "The big idea:
polonium, radon and cigarettes" .
Journal of the Royal Society of
Medicine. 101 (3): 156–7.
doi:10.1258/jrsm.2007.070021 .
PMC 2270238 . PMID 18344474 .
115. Birnbauer, William (2008-09-07) "Big
Tobacco covered up radiation
danger" . The Age, Melbourne,
Australia
116. Radford EP Jr; Hunt VR (1964).
"Polonium 210: a volatile radioelement
in cigarettes". Science. 143 (3603):
247–9. Bibcode:1964Sci...143..247R .
doi:10.1126/science.143.3603.247 .
PMID 14078362 .
117. Kelley TF (1965). "Polonium 210
content of mainstream cigarette
smoke". Science. 149 (3683): 537–
538. Bibcode:1965Sci...149..537K .
doi:10.1126/science.149.3683.537 .
PMID 14325152 .
118. Ota, Tomoko; Sanada, Tetsuya;
Kashiwara, Yoko; Morimoto, Takao; et
al. (2009). "Evaluation for Committed
Effective Dose Due to Dietary Foods by
the Intake for Japanese Adults" .
Japanese Journal of Health Physics.
44: 80–88. doi:10.5453/jhps.44.80 .
119. Smith-Briggs, JL; Bradley, EJ (1984).
"Measurement of natural radionuclides
in U.K. diet". Science of the Total
Environment. 35 (3): 431–40.
Bibcode:1984ScTEn..35..431S .
doi:10.1016/0048-9697(84)90015-9 .
PMID 6729447 .

Bibliography
Bagnall, K. W. (1962). "The Chemistry of
Polonium" . Advances in Inorganic
Chemistry and Radiochemistry. 4. New
York: Academic Press. pp. 197–226.
doi:10.1016/S0065-2792(08)60268-X .
ISBN 978-0-12-023604-6. Retrieved
June 14, 2012.
Greenwood, Norman N.; Earnshaw, Alan
(1997). Chemistry of the Elements (2nd
ed.). Butterworth–Heinemann.
ISBN 978-0080379418.

External links

Wikimedia Commons has media


related to Polonium.

Look up Polonium in Wiktionary, the


free dictionary.

Polonium at The Periodic Table of


Videos (University of Nottingham)

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Polonium&oldid=903982403"

Last edited 4 days ago by an anony…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like