This is a good article. Click here for more information.

Lead

From Wikipedia, the free encyclopedia
Jump to: navigation, search
This article is about the metal. For other uses, see Lead (disambiguation).
Lead,  82Pb
Lead electrolytic and 1cm3 cube.jpg
General properties
Name, symbol lead, Pb
Pronunciation /ˈlɛd/
LED
Appearance metallic gray
Lead in the periodic table
Hydrogen (diatomic nonmetal)
Helium (noble gas)
Lithium (alkali metal)
Beryllium (alkaline earth metal)
Boron (metalloid)
Carbon (polyatomic nonmetal)
Nitrogen (diatomic nonmetal)
Oxygen (diatomic nonmetal)
Fluorine (diatomic nonmetal)
Neon (noble gas)
Sodium (alkali metal)
Magnesium (alkaline earth metal)
Aluminium (post-transition metal)
Silicon (metalloid)
Phosphorus (polyatomic nonmetal)
Sulfur (polyatomic nonmetal)
Chlorine (diatomic nonmetal)
Argon (noble gas)
Potassium (alkali metal)
Calcium (alkaline earth metal)
Scandium (transition metal)
Titanium (transition metal)
Vanadium (transition metal)
Chromium (transition metal)
Manganese (transition metal)
Iron (transition metal)
Cobalt (transition metal)
Nickel (transition metal)
Copper (transition metal)
Zinc (transition metal)
Gallium (post-transition metal)
Germanium (metalloid)
Arsenic (metalloid)
Selenium (polyatomic nonmetal)
Bromine (diatomic nonmetal)
Krypton (noble gas)
Rubidium (alkali metal)
Strontium (alkaline earth metal)
Yttrium (transition metal)
Zirconium (transition metal)
Niobium (transition metal)
Molybdenum (transition metal)
Technetium (transition metal)
Ruthenium (transition metal)
Rhodium (transition metal)
Palladium (transition metal)
Silver (transition metal)
Cadmium (transition metal)
Indium (post-transition metal)
Tin (post-transition metal)
Antimony (metalloid)
Tellurium (metalloid)
Iodine (diatomic nonmetal)
Xenon (noble gas)
Caesium (alkali metal)
Barium (alkaline earth metal)
Lanthanum (lanthanide)
Cerium (lanthanide)
Praseodymium (lanthanide)
Neodymium (lanthanide)
Promethium (lanthanide)
Samarium (lanthanide)
Europium (lanthanide)
Gadolinium (lanthanide)
Terbium (lanthanide)
Dysprosium (lanthanide)
Holmium (lanthanide)
Erbium (lanthanide)
Thulium (lanthanide)
Ytterbium (lanthanide)
Lutetium (lanthanide)
Hafnium (transition metal)
Tantalum (transition metal)
Tungsten (transition metal)
Rhenium (transition metal)
Osmium (transition metal)
Iridium (transition metal)
Platinum (transition metal)
Gold (transition metal)
Mercury (transition metal)
Thallium (post-transition metal)
Lead (post-transition metal)
Bismuth (post-transition metal)
Polonium (post-transition metal)
Astatine (metalloid)
Radon (noble gas)
Francium (alkali metal)
Radium (alkaline earth metal)
Actinium (actinide)
Thorium (actinide)
Protactinium (actinide)
Uranium (actinide)
Neptunium (actinide)
Plutonium (actinide)
Americium (actinide)
Curium (actinide)
Berkelium (actinide)
Californium (actinide)
Einsteinium (actinide)
Fermium (actinide)
Mendelevium (actinide)
Nobelium (actinide)
Lawrencium (actinide)
Rutherfordium (transition metal)
Dubnium (transition metal)
Seaborgium (transition metal)
Bohrium (transition metal)
Hassium (transition metal)
Meitnerium (unknown chemical properties)
Darmstadtium (unknown chemical properties)
Roentgenium (unknown chemical properties)
Copernicium (transition metal)
Ununtrium (unknown chemical properties)
Flerovium (post-transition metal)
Ununpentium (unknown chemical properties)
Livermorium (unknown chemical properties)
Ununseptium (unknown chemical properties)
Ununoctium (unknown chemical properties)
Sn

Pb

Fl
thalliumleadbismuth
Atomic number (Z) 82
Group, block group 14 (carbon group), p-block
Period period 6
Element category   post-transition metal
Standard atomic weight (±) (Ar) 207.2(1)[1]
Electron configuration [Xe] 4f14 5d10 6s2 6p2
per shell
2, 8, 18, 32, 18, 4
Physical properties
Phase solid
Melting point 600.61 K ​(327.46 °C, ​621.43 °F)
Boiling point 2022 K ​(1749 °C, ​3180 °F)
Density near r.t. 11.34 g/cm3
when liquid, at m.p. 10.66 g/cm3
Heat of fusion 4.77 kJ/mol
Heat of vaporization 179.5 kJ/mol
Molar heat capacity 26.650 J/(mol·K)
vapor pressure
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 978 1088 1229 1412 1660 2027
Atomic properties
Oxidation states 4, 3, 2, 1, −1, −2, −4 ​(an amphoteric oxide)
Electronegativity Pauling scale: 1.87
Ionization energies 1st: 715.6 kJ/mol
2nd: 1450.5 kJ/mol
3rd: 3081.5 kJ/mol
Atomic radius empirical: 175 pm
Covalent radius 146±5 pm
Van der Waals radius 202 pm
Miscellanea
Crystal structure face-centered cubic (fcc)
Face-centered cubic crystal structure for lead
Speed of sound thin rod 1190 m/s (at r.t.) (annealed)
Thermal expansion 28.9 µm/(m·K) (at 25 °C)
Thermal conductivity 35.3 W/(m·K)
Electrical resistivity 208 nΩ·m (at 20 °C)
Magnetic ordering diamagnetic
Young's modulus 16 GPa
Shear modulus 5.6 GPa
Bulk modulus 46 GPa
Poisson ratio 0.44
Mohs hardness 1.5
Brinell hardness 38–50 MPa
CAS Number 7439-92-1
History
Discovery Middle Easterns (7000 BCE)
Most stable isotopes of lead
iso NA half-life DM DE (MeV) DP
204Pb 1.4% >1.4×1017 y (α) 1.972 200Hg
205Pb syn 1.53×107 y ε 0.051 205Tl
206Pb 24.1% (α) 1.1366 202Hg
207Pb 22.1% (α) 0.3915 203Hg
208Pb 52.4% >2×1019 y (α) 0.5188 204Hg
210Pb trace 22.3 y α 3.792 206Hg
β 0.064 210Bi
Decay modes in parentheses are predicted, but have not yet been observed
| references

Lead (/lɛd/) is a chemical element in the carbon group with symbol Pb (from Latin: plumbum) and atomic number 82. It is a soft, malleable and heavy post-transition metal. Freshly cut, solid lead has a bluish-white color that soon tarnishes to a dull grayish color when exposed to air; the liquid metal has shiny chrome-silver luster. Lead has the highest atomic number[a] of any non-radioactive element (two radioactive elements, namely technetium and promethium, are lighter), although the next higher element, bismuth, has one isotope with a half-life that is long enough (over one billion times the estimated age of the universe) to be considered stable. Lead's four stable isotopes each have 82 protons, a magic number in the nuclear shell model of atomic nuclei. The isotope lead-208 also has 126 neutrons, another magic number, and is hence double magic, a property that grants it enhanced stability: lead-208 is the heaviest known stable nuclide.

Lead is used in building construction, lead-acid batteries, bullets and shot, weights, as part of solders, pewters, fusible alloys, and as a radiation shield.

If ingested or inhaled, lead and its compounds are poisonous to animals and humans. Lead is a neurotoxin that accumulates both in soft tissues and the bones, damaging the nervous system and causing brain disorders. Excessive lead also causes blood disorders in mammals. Lead poisoning has been documented since ancient Rome, ancient Greece, and ancient China.

Physical characteristics[edit]

Bulk properties[edit]

A sample of lead freshly solidified from a molten state

Lead is a bright silvery metal with a very slight shade of blue in a dry atmosphere.[2] It tarnishes on contact with air, forming a complex mixture of compounds whose color and composition depend on conditions, sometimes with significant amounts of carbonates and hydroxycarbonates.[3][4] Lead's characteristic properties include high density, softness, ductility, malleability, poor electrical conductivity compared to other metals, high resistance to corrosion, and ability to react with organic chemicals.[2]

Various traces of other metals significantly change its properties: adding small amounts of antimony or copper increases the lead alloy's hardness and improves resistance to sulfuric acid corrosion.[2] Some other metals, such as cadmium, tin, and tellurium, improve hardness and fight metal fatigue. Sodium and calcium also have this ability, but they reduce the alloy's chemical stability.[2] Finally, zinc and bismuth simply impair the corrosion resistance.[2] Adding small amounts of lead improves the ductility of steel, but lead impurities mostly worsen the quality of industrial materials.[2]

Lead has only one common allotrope, which is face-centered cubic, with the length of an edge of a unit cell being 349 pm.[5] At 327.5 °C (621.5 °F),[6] lead melts; the melting point exceeds that of tin (232 °C, 449.5 °F),[6] but is significantly below that of germanium (938 °C, 1721 °F).[7] The boiling point of lead is 1749 °C (3180 °F),[8] below those of both tin (2602 °C, 4716 °F)[6] and germanium (2833 °C, 5131 °F).[7] Densities increase down the group: the values of germanium and tin (5.23[9] and 7.29 g·cm−3,[10] respectively) are significantly below that of lead: 11.32 g·cm−3.[9]

Isotopes[edit]

Main article: Isotopes of lead

Lead has four observationally stable isotopes, lead-204, lead-206, lead-207, and lead-208.[11] The comparable multitude of its stable isotopes relies on the fact that lead's atomic number of 82 is even.[b] With its high atomic number, lead is the second heaviest element that occurs naturally in the form of isotopes that could be treated as stable for any practical applications: bismuth has a higher atomic number of 83, but its only stable isotope was found in 2003 to be actually very slightly radioactive.[c] The four stable isotopes of lead could theoretically undergo alpha decay with release of energy as well, but this has not been observed for any of them.[12] As such, lead is often quoted as the heaviest stable element.

Three of these isotopes also found in three of the four major decay chains: lead-206, lead-207, and lead-208 are final decay products of uranium-238, uranium-235, and thorium-232, respectively; the decay chains are called uranium series, actinium series, and thorium series, respectively. Since the amounts of them in nature depend also on other elements' presence, the isotopic composition of natural lead varies by sample: in particular, the relative amount of lead-206 may vary between 20.84% and 27.78%.[11] (For this reason, the atomic weight of lead is given with such imprecision, one decimal place.[15]) As time goes, relative amounts of lead-206 and lead-207 to that of lead-204 increase, since the former two are regenerated by radioactivity of heavier elements and the latter is not; this allows for the lead–lead dating. Analogously, as uranium decays (eventually) into lead, their relative amounts change; this allows for uranium–lead dating.

Apart from the stable isotopes, which make up almost all of lead that exists naturally, there are trace quantities of a few radioactive isotopes. One of them is lead-210; although it has a half-life of 22.3 years,[12] a period too short to allow any primordial lead-210 to still exist, some small quantities of it exist, because lead-210 is found in uranium series: even though it constantly decays away, its amount is also constantly regenerated by decay of its parent, polonium-214, which, while also constantly decaying, is also supplied by decay of its parent, and so on, all the way up to original uranium-238, which has been present for billions of years on Earth. Lead-210 is particularly well known for helping to identify ages of samples containing it, which is performed by measuring lead-210 to lead-206 ratios (both isotopes are present in a single decay chain); however, lead-214 is also present in that chain, and lead-212 is present in thorium series; therefore, traces of both isotopes exist naturally as well.[16]

In total, thirty-eight isotopes of lead have been synthesized, those with mass numbers of 178–215.[12] Lead-205 is the most stable radioisotope of lead, with a half-life of around 1.5×107 years.[d] Additionally, 47 nuclear isomers (long-lived excited nuclear states), corresponding to 24 lead isotopes, have been characterized. The longest-lived isomer is lead-204m2 (half-life of about 1.1 hours).[12]

Chemical characteristics[edit]

A lead atom has 82 electrons, arranged in an electronic configuration of [Xe]4f145d106s26p2. The first and second ionization energies—energies required to remove an electron from a neutral atom and an electron from a resulting singly charged ion—of lead combined are close to those of tin, its upper group 14 neighbor; this proximity is caused by the 4f shell—no f shell is present in previous group 14 elements atoms—and the thereby following lanthanide contraction. However, the first four ionization energies of lead combined exceed those of tin,[18] opposite to what the periodic trends would predict. For that reason, unlike tin,[19] lead is reluctant[19] to form the +4 oxidation state in inorganic compounds.

Such unusual behavior is rationalized by relativistic effects, which are increasingly stronger closer to the bottom of the periodic table;[19] one of such effects is the spin–orbit (SO) interaction, particularly the inert pair effect, which stabilizes the 6s orbital.[e] The inert pair effect in lead comes from the great difference in electronegativity between lead and the anions (oxide, halides, nitrides), which results in positive charge on lead and then leads to a stronger contraction of the 6s orbital than the 6p orbital, making the 6s orbital inert.[21] (However, this is not applicable to compounds in which lead forms covalent bonds; as such, lead, similar to carbon, is dominantly tetravalent in organolead compounds.) The SO interaction not only stabilizes the 6s electron levels, but also two of the six 6p levels; and lead has just two 6p electrons. This effect takes part in making lead slightly more stable chemically, but the effect is even more important for the period 7 elements, and as such, it is often mentioned in a chemical description of these elements but not lead.

The figures for electrode potential show that lead is only slightly easier to oxidize than hydrogen. Lead thus can dissolve in acids, but this is often impossible due to specific problems (such as the formation of insoluble salts).[22] Electronegativity, although often thought to be constant for each element, is a variable property; lead shows a high electronegativity difference between values for lead(II) and lead(IV)—1.87 and 2.33, accordingly. This marks the reversal of the trend of increasing stability of the +4 oxidation state in group 14 down the group into decreasing; tin, for comparison, has electronegativities of 1.80 and 1.96.[23]

Reactivity[edit]

Powdered lead burns with a bluish-white flame. As with many metals, finely divided powdered lead exhibits pyrophoricity.[24] Bulk lead released to the air forms a protective layer of insoluble lead oxide, which covers the metal from undergoing further reactions.[25] Other insoluble compounds, such as sulfate or chloride, may form the protective layer if lead is exposed to a different chemical environment.[26]

Fluorine reacts with lead at room temperature, forming lead(II) fluoride. The reaction with chlorine is similar, although it requires heating: the chloride layer diminishes the reactivity of the elements.[25][26] Molten lead reacts with chalcogens.[27]

Presence of carbonates or sulfates results in the formation of insoluble lead salts, which protect the metal from corrosion. So does carbon dioxide, as the insoluble lead carbonate is formed; however, an excess of the gas leads to the formation of the soluble bicarbonate, which makes the use of lead pipes dangerous.[22] Water in the presence of oxygen attacks lead to start an accelerating reaction.[28] Lead also dissolves in quite concentrated alkalis (≥10%) because of the amphoteric character and solubility of plumbites.[22]

The metal is normally not attacked by sulfuric acid; however, concentrated acid does dissolve lead thanks to complexation.[28] Lead does react with hydrochloric acid, albeit slowly, and nitric acid, quite actively, to form nitrogen oxides and lead(II) nitrate.[28] Organic acids, such as acetic acid, also dissolves lead, but this reaction requires oxygen as well.[26]

Inorganic compounds[edit]

In a vast majority of its compounds, lead occurs in oxidation states +2 or +4. Lead(II) compounds are normally ionic but lead(IV) compounds are often covalent. Even the strongest oxidizing elements (oxygen, fluorine) oxidize lead to only lead(II) initially.

Lead(II)[edit]

Most inorganic compounds lead forms are lead(II) compounds. This includes binary compounds; lead forms such compounds with many nonmetals, but not with every one: for example, there is no known lead carbide.

Even though most lead(II) compounds are ionic, they are not as ionic as those of many other metals. In particular, many lead(II) compounds are water-insoluble. In solution, lead(II) ions are colorless, but under specific conditions, lead is capable of changing its color.[29] Unlike tin(II) ions, these do not react as reducing agents in solution.

Lead monoxide exists in two allotropes, red α-PbO and yellow β-PbO, the latter being stable only from around 488 °C. It is the most commonly applicable compound of lead.[30] However, its hydroxide counterpart, lead(II) hydroxide, is not capable of existence outside solutions; in solution, it is known to form anions, plumbites. Lead commonly reacts with chalcogens other than oxygen. Lead sulfide can only be dissolved in strong acids.[31] It is a semiconductor, a photoconductor, and an extremely sensitive infrared radiation detector. A mixture of the monoxide and the monosulfide when heated forms the metal.[32] The other two chalcogenides are photoconducting as well.[33]

Lead dihalides are known and well-characterized; this refers to not only the binary halides, to some extent even including diastatide,[34] but also mixed ones, such as PbFCl, etc. The difluoride is the first ionically conducting compound to have been discovered. The other dihalides decompose on exposure to light, especially notably for the diiodide. There are anion counterparts for the heavier three dihalides, such as PbCl4−
6
.

Lead(IV)[edit]

Few lead(IV) compounds are known. Inorganic lead(IV) compounds are typically strong oxidants or exist only in highly acidic solutions.[19] Lead(II) oxide gives a mixed oxide on further oxidation, Pb
3
O
4
. It is described as lead(II,IV) oxide, or structurally 2PbOPbO
2
, and is the best-known mixed valence lead compound. Lead dioxide is a strong oxidizing agent, capable of oxidizing hydrochloric acid. Like lead monoxide, lead dioxide is capable of forming anions, plumbates. Lead tetrafluoride, a yellow crystalline powder, is stable, but less stable than the difluoride. Lead tetrachloride decomposes at room temperature, lead tetrabromide is less stable still and the existence of lead tetraiodide is questionable.[35][36] Lead disulfide, like the monosulfide, is a semiconductor.[37] Lead(IV) selenide is also known.[38]

Other oxidation states[edit]

A few compounds exist in oxidation states other than +2 and +4, but they don't have a great impact on lead chemistry from either theoretical or industrial perspective. Lead(III) may be obtained under specific conditions as an intermediate between lead(II) and lead(IV), in larger organolead complexes rather than by itself.[39][40] This oxidation state is not specifically stable, as lead(III) ion (as well as, consequently, larger complexes containing it) is a radical; same applies for lead(I), which can also be found in such species.[41] Negative oxidation states can occur as Zintl phases, as either free lead ions, for example, in Ba
2
Pb
, with lead formally being lead(−IV),[42] or cluster ions, for example, in a Pb5−
2
ion, where two lead atoms are lead(−I) and three are lead(0).[43]

Organolead[edit]

Main article: Organolead compound

Since lead is a heavier carbon homolog, it shares with carbon the property of being able to build long chains of atoms, bonded via single or multiple bonds: catenation. Lead may therefore behave in a similar way to carbon with regards to covalent chemistry. Lead atoms can build metal–metal bonds of order up to three,[44] although lower orders are also possible. Alternatively, lead is also known to build bonds to carbon; the carbon–lead bonds are covalent, and compounds containing such bonds thus resemble typical organic compounds.[45] Compound containing the lead–carbon bond are called organolead compounds. In general, such compounds are not very stable chemically.

The simplest lead analog of an organic compound is plumbane, the lead analog of methane. It is unstable against heat, decaying in heated tubes,[46] and thermodynamically;[47] in general, little is known about chemistry of plumbane, as it is so unstable. A lead analog of the next alkane, ethane, is not known.[46] Two simple plumbane derivatives, tetramethyllead and tetraethyllead, are the best-known organolead compounds. These compounds are relatively unstable against heating—tetraethyllead starts to decompose at only 100 °C (210 °F)[45]—as well as sunlight or ultraviolet light.[48] General oxidizing nature of organolead compounds find its use in chemistry: tetraethyllead is produced in larger quantities than any other organometallic compound;[49] lead tetraacetate is an important laboratory reagent for oxidation in organic chemistry.[citation needed] Other organolead compounds, including homologs of the said compounds, are less stable chemically still.[45]

Lead readily forms an equimolar alloy with sodium metal that reacts with alkyl halides to form organometallic compounds of lead such as tetraethyllead.[50] Plumbane may be obtained in a reaction between metallic lead and atomic (not molecular) hydrogen.[46] Atoms of chlorine or bromine displace alkyls in tetramethyllead and tetraethyllead; hydrogen chloride, a by-product of the previous reaction, further reacts with the halogenated molecules to complete mineralization—chemical reaction or a series of reactions transforming an organic compound into an inorganic one—of the original compounds, yielding lead dichloride.[48]

Origin and occurrence[edit]

In space[edit]

Chart representing the final part of the s-process. Red horizontal lines with a circle in their right ends represent neutron captures; blue arrows pointing up-left represent beta decays; green arrows pointing down-left represent alpha decays; cyan arrows pointing down-right represent electron captures.

Primordial lead—the isotopes lead-204, lead-206, lead-207, and lead-208—was created by the s-process and the r-process. The letter "s" stands for "slow" or "slow neutron capture", and the letter "r" stands for "rapid neutron capture": in the s-process another capture takes a long time, centuries or millennia, while the r-process takes only tens of seconds to result in a heavy nuclide of lead's mass. In the s-process, a nucleus in a star captures another slow neutron, and if the resulting nucleus is unstable, it typically undergoes a beta decay to become an element of the next atomic number. Lead-204 is created from short-lived thallium-204; on capturing another neutron, it becomes lead-205, which, while unstable, is stable enough to generally last longer than a capture takes (its half-life is around 15 million years). Further captures result in lead-206, lead-207, and lead-208. On capturing another neutron, lead-208 becomes lead-209, which quickly decays into bismuth-209, which on capturing another neutron becomes bismuth-210, which either undergoes an alpha decay to result in thallium-206, which would beta decay into lead-206, or a beta decay to yield polonium-210, which would inevitably alpha decay into lead-206 as well, and the cycle ends at lead-206, lead-207, lead-208, and bismuth-209. As a result, relative abundances of the three have stable lead isotopes are multiplied by a "cycling factor", which depends on the conditions of the process.[51]

Apart from the s-process, the latter three isotopes have been synthesized as a result of the r-process (lead-204 is not produced in this manner because its isobar mercury-204 is stable, and it is not formed as a decay product of r-process products). The r-process happens in mediums of great electron density. In such conditions, beta decay is blocked, because the high electron density fills all available free electron states up to a Fermi energy which is greater than the energy of nuclear beta decay. But nuclear capture of those free neutrons can still occur, and it causes neutron enrichment of matter. This results an extremely high density of free neutrons which cannot decay, and, correspondingly, a large neutron flux and high temperatures, which is the reason why neutron capture occurs much faster than beta decay. Furthermore, since these lead isotopes are also located at the end of three major decay chains (see above), they are created by the decay of the heavier elements as well, starting with thorium and uranium, and these elements are synthesized by the r-process as well.[51]

Solar System abundances[52]
Atomic
number
Element Relative
amount
42 Molybdenum 1.0
46 Palladium 0.3
50 Tin 0.9
52 Tellurium 1.6
56 Barium 1.2
80 Mercury 0.1
82 Lead 1
92 Uranium 0.0052

The isotopes at the end of the chains make up around 98.02% lead in the universe, with non-radiogenic lead-204 making up slightly less than two percent.[52] Lead is not an abundant element in general—its per-particle abundance in the Universe is 0.06 ppb[53]—still, it is an order of magnitude than that of mercury, and further exceeds those of many other elements of close atomic numbers. After element 40 (zirconium), no element is at least twofold as abundant as lead, and there is no element as abundant as lead starting after element 56 (barium). Lead is three times as abundant as platinum, ten times as mercury, and twenty times as gold.[52] Per mass, lead's abundance is 10 ppb[53]—the difference between the per-mass and per-particle abundances is justified by mass difference between lead isotopes and the most common elements: the most common nuclide in the Universe, hydrogen-1, has a mass of approximately one atomic mass unit, while those of lead isotopes have masses of over 200 atomic mass units.

On Earth[edit]

Lead is a quite common element in the Earth's crust for its high atomic number.

Since lead commonly reacts with sulfur (see above), it is classified as a chalcophile using the Goldschmidt classification. Lead is likely to form minerals that do not sink into the core but that stay above on Earth in its crust, even though without sinking deep into it. Lead's abundance in the Earth's crust is 16 ppm.[54] This results in a great availability of lead minerals and easy extraction of the metal; for this reason, the mineral form of its sulfide, galena, has been known for millennia, as was the metal itself (see below). Lead's pronounced chalcophilic character is close to those of zinc and copper; as such, it is usually found in ore and extracted together with these metals.[54] Metallic lead does occur in nature, but it is rare. As a result of lead's chemistry, it occurs in primary minerals exclusively as lead(II), unlike tin, which always occurs as tin(IV).[15][f] Lead deposits can be hydrothermal vein, impregnation, and replacement deposits; volcanogenic sedimentary deposits; and hydrothermal or marine sedimentary deposits. World resources of lead exceed 2 billion tons.[55] Massive resources are located in Australia, China, Ireland, Mexico, Peru, Portugal, Russia, and the United States. World reserves—resources ready to be mined for which that would be economically feasible—totaled 89 million tons in 2015, of which Australia had 35 million, China had 15.8 million, and Russia had 9.2 million.[55]

Lead and zinc bearing carbonate and clastic deposits

The main lead mineral is galena (PbS). Galena is mostly found with other minerals, mostly zinc ores.[54] Most other lead minerals are normally related to galena in some way; for example, boulangerite, Pb
5
Sb
4
S
11
, is a mixed sulfide derived from galena; anglesite, PbSO
4
, is a product of galena oxidation; cerussite or white lead ore, PbCO
3
, is a decomposition product of galena. Zinc, copper, arsenic, tin, anitmony, silver, gold, and bismuth are common impurities in lead minerals.[54]

History[edit]

Lead Roman pipes inscribed IMP.VESP.VIIII.T.IMP.VII.COS.CN.­IULIO.AGRICOLA.LEG.AUG.PR.PR.[g]

Lead has been commonly used for thousands of years because it is widespread, easy to extract, and easy to work with. It is highly malleable and easily smeltable. Metallic lead beads dating back to 7000–6500 BCE, if not before that, have been found in Asia Minor; this indicates lead was the first metal to be ever smelted.[56] Since then, the metal has been used by many ancient peoples. A major reason for the spread of lead production was its association with silver, which may be obtained by burning galena, a widespread lead mineral.[57] The Ancient Egyptians are thought to have used lead for sinkers in fishing nets, in glazes, glasses and enamels, and for ornaments. Various civilizations of the Fertile Crescent used lead as a writing material, as currency, and for construction. The Ancient Chinese used lead as a stimulant in the royal court,[57] a currency,[58] and a contraceptive;[59] lead also had a few uses, such as making amulets, for the Indus Valley civilization and the Mesoamericans.[57] Peoples of eastern and southern Africa are known to exercise wire drawing.[60]

Lead mines were worked in 2000 BCE in the Iberian peninsula by the Phoenicians;[61] and also in Athens, Carthage, and Sicily.[57] Lead was mined in Ancient China before 1000 BCE.[62] With the development of mining and its territorial expansion in Europe and across the Mediterranean, Rome became the greatest producer of lead during the classical era, with an estimated annual output equaling 80,000 tonnes. The Romans obtained lead mostly as a by-product of extensive silver smelting.[63][64][65] Lead mining occurred in Central Europe, Britain, the Balkans, Greece, Anatolia, and Hispania, which alone accounted for 40% of world production.[63] Lead was used for making water pipes in the Roman empire and consequently the Latin word for the metal, plumbum, was the origin of the English word "plumbing" and its derivatives[66]—even though some Romans, such as Vitruvius, were able to recognize its danger for health.[67] Nevertheless, a number of researchers suggest lead poisoning was one of the reasons behind the fall of Rome.[h] Lead poisoning—a condition in which one becomes dark and cynical—was called "saturnine", after the ghoulish god of Saturn; the metal was also considered the father of all metals. It was easily available in the Roman society, and as such, its social status was low.[69]

World lead production peaking in the Roman period and the rising Industrial Revolution
(present on the picture = 1980)[63]

During the ancient and classical eras, (and even far beyond them, until the 17th century), tin was often not distinguished from lead or seen as a different kind of the metal that lead is: Romans called lead plumbum nigrum (literally, "black lead"), while tin was called plumbum candidum (literally, "bright lead"). Their association through history can also be seen in other languages: the word olovo in Czech translates to "lead", but in Russian the cognate олово (olovo) means "tin".[70] In addition to that, lead also bore a close relation to antimony: Both elements commonly occur as sulfides (galena and stibnite), often together. Pliny declared stibnite would give lead on heating, whereas the mineral on heating actually produces antimony.[71] The originally South Asian surma—"galena" in English—spread across Asia with that meaning, and also gave its name to antimony in a number of Central Asian languages, as well as Russian.

Lead plumbing in Western Europe may have been continued beyond the fall of the Western Roman Empire into the medieval era,[72] but lead mining in Europe in general fell into decline,[73][74] and the largest lead production was conducted in South and East Asia, where lead output underwent a strong growth.[74] In European alchemy, lead continued its status of the oldest metal and its association with Saturn—this time, the planet named after the Roman god rather than the god himself. Alchemists accordingly used Saturn's symbol (the scythe, ♄) to refer to lead.[75] During the period, lead has become increasingly more used as for wine adulteration. This practice was declared forbidden in 1498 by a papal bull, but it continued long past the date, being a reason of various mass poisonings up to late 18th century.[73] In the wake of the Renaissance, the printing press was invented, and lead served as a key material for its parts, starting with the Johannes Gutenberg's press;[76] however, lead dust also was commonly inhaled by operators, causing lead poisoning.[77] Additionally, firearms were invented approximately at the same time, and lead, despite its expense over iron, became a chief material for making bullets, because it made less damage to iron gun barrels, had a higher density (which allowed better retaining velocity and energy), and its lower melting point made production much easier: bullets could be made on wooden fire.[78][79]

Papal bull of 1637 with a lead stamp

In the New World, lead was first produced soon after the European settlers had arrived; the earliest recorded lead production dates to 1621, in the Colony of Virginia that had been founded fourteen years earlier.[69][80] In Australia, mining was introduced by the colonists as well, and they opened the first mine on the continent—a lead mine—in 1841.[81] However, centuries before the Europeans were able to start the colonization of Africa in the late 19th century, lead mining was known in the Benue Trough[82] and the lower Congo basin, where lead was used for trade with the Europeans and as a currency.[83][i]

In the second half of the 18th century, Britain and later continental Europe and then the United States entered the Industrial Revolution. During the period, lead mining proved important; the Industrial Revolution was the first time to have greater lead production rates than those of Rome.[63] Britain was the leading producer during the period, losing the status of the greatest producer by the mid-19th century with depletion of its mines and development of lead mining in Germany, Spain, and the United States.[84] The United States took the lead by 1900;[85] other non-European nations—in particular, Canada, Mexico, and Australia—started their massive lead production, and by 1900, Europe's output of lead fell below that elsewhere.[86] A great share of demand of lead came from plumbing and painting—lead paints had been invented and regularly used; with invention of gasoline in late 19th century, lead was extensively used as an additive.[87] At this time, more people—the working class—contacted the metal, and this led to the increase of the numbers of those poisoned by lead. This also led to research of effects of lead intake: lead was proven to be more dangerous in its fume form than as a solid metal; lead poisoning and gout were linked (Alfred Baring Garrod noted a third of his gout patients was plumbers and painters); effects of chronic ingestion of lead, including mental disorders, were all studied in the 19th century. The first political acts to decrease the degree of lead poisoning in factories followed in the 1870s and 1880s in the United Kingdom.[87]

Lead mining in the upper Mississippi River region in the United States in 1865

Further evidence of the threat lead posed to human organisms were revealed in the late 19th and early 20th centuries—mechanisms of the harm were better realized, and lead blindness was documented[88]—and countries in Europe and the United States started efforts to reduce the amount of lead a regular person contacts with. The last major innovation to impose contact with lead on humans was adding tetraethyllead to gasoline, invented in the United States in 1921; it was phased out in the U.S. and the European Union by 2000.[87] Most European countries banned usage of lead paint for interiors by 1930.[89] The result of many regulations and bans put on lead products was significant: in the last quarter of the 20th century, percentage of people with excessive lead blood levels dropped from over three quarters of the population to slightly over two percent in the U.S.[87] By the end of the 20th century, the main good made of lead was the lead–acid battery,[90] which possesses no direct threat to humans. That allowed for a consistent lead production in the industrialized countries. From 1960 to 1990, lead output in the Western Bloc grew by 31%.[91] The share of the world's lead production of the Eastern Bloc increased from 10% and 30% from 1950 to 1990, with the Soviet Union being world's largest producer during the mid- and late 1970s and the 1980s, and China started a massive lead production in the late 20th century.[92] Unlike the European communist countries, China was largely unindustrialized by mid-20th century; in 2004, China surpassed Australia as the largest producer of lead.[93] However, in part similarly to the European industrialization, lead does have a negative effect on the global health in the country.[94]

Production[edit]

Ore processing[edit]

Historical evolution of the extracted lead ore grade extracted in Canada and Australia.

Most ores contain less than 10% lead, and ores containing as little as 3% lead can be economically exploited.[citation needed] During initial ore processing, ores typically undergo crushing, dense-medium separation, grinding, froth flotation, and drying of the resulting concentrate. The resulting concentrate is the initial quantitative metric of mined lead.[95] Sulfide ores are roasted, producing primarily lead oxide and a mixture of sulfates and silicates of lead and other metals contained in the ore.[96] Lead oxide from the roasting process is reduced in a coke-fired blast furnace to the metal.[97] Sulfate concentrate is more common for subsequent lead production than oxide concentrate; it commonly has a lead content fraction of 50%–60%, occasionally varying to up to 30% or 80%.[98]

Additional layers separate in the process and float to the top of the metallic lead. These are slag (silicates containing 1.5% lead), matte (sulfides containing 15% lead), and speiss (arsenides of iron and copper). These wastes contain concentrations of copper, zinc, cadmium, and bismuth that can be recovered economically, as can their content of unreduced lead.[99]

Galena, lead ore

Metallic lead that results from the roasting and blast furnace processes still contains significant contaminants of arsenic, antimony, bismuth, zinc, copper, silver, and gold. The melt is treated in a reverberatory furnace with air, steam, and sulfur, which oxidizes the contaminants except silver, gold, and bismuth. The oxidized contaminants are removed by drossing, where they float to the top and are skimmed off.[99][100] Since lead ores contain significant concentrations of silver, the smelted metal also is commonly contaminated with silver. Metallic silver as well as gold is removed and recovered economically by means of the Parkes process.[32][99][100] Desilvered lead is freed of bismuth according to the Betterton-Kroll process by treating it with metallic calcium and magnesium, which forms a bismuth dross that can be skimmed off.[99][100] Very pure lead can be obtained by processing smelted lead electrolytically by means of the Betts process. The process uses anodes of impure lead and cathodes of pure lead in an electrolyte of silica fluoride.[99][100]

Production and recycling[edit]

Historical evolution of the production of lead, as extracted in different countries.

Production and consumption of lead is increasing worldwide. Lead production generally is divided into two major categories, primary and secondary: the primary production is the production from concentrate from the previously mined ores, and the secondary production is the production from scrap. In 2013, 4.74 million metric tons came from the primary production, and 5.74 million tons came from secondary production. The top mining countries for lead in 2013 were China, Australia, Russia, India, Bolivia, Sweden, North Korea, South Africa, Poland, and Ireland. The top lead producing countries were China, United States, India, South Korea, Germany, Mexico, United Kingdom, Canada, Japan, and Australia.[95]

World's largest mining countries of lead, 2015[55]
Country Output
(thousand
tons)
 China 2,300
 Australia 633
 United States 385
 Peru 300
 Mexico 240
 India 130
 Russia 90
 Bolivia 82
 Sweden 76
 Turkey 54
 North Korea 45
 Poland 40
 South Africa 40
 Kazakhstan 38
 Ireland 33
Other countries 226

In 2007, New Scientist published an article predicting depletion of lead supply in 42 years based on then-current use rates.[101] A year before, environmental analyst Lester Brown suggested lead could run out within 18 years based on an extrapolation of 2% growth per year.[102] This may need to be reviewed to take account of renewed interest in recycling, and rapid progress in fuel cell technology. According to the International Resource Panel's Metal Stocks in Society report, the global per capita stock of lead in use in society is 8 kg. Much of this is in more-developed countries (20–150 kg per capita) rather than less-developed countries (1–4 kg per capita).[103]

Applications[edit]

Lead when mined contains an unstable isotope, lead-210, which has a half life of 22 years. This makes lead slightly radioactive. As such ancient lead which has almost no radioactivity is sometimes desired for scientific experimentation.[104][105]

Elemental form[edit]

Lead bricks are commonly used as radiation shielding.

Contrary to popular belief, pencil leads in wooden pencils have never been made from lead. The term comes from the Roman stylus, called the penicillus, a small brush used for painting.[106] When the pencil originated as a wrapped graphite writing tool, the particular type of graphite being used was named plumbago (lit. act for lead, or lead mockup).[107][108]

Lead is used in applications where its low melting point, ductility and high density are advantageous. In either pure form or alloyed with tin, or antimony, lead is the traditional material for bullets and shot in firearms use. The low melting point makes casting of lead easy, and therefore small arms ammunition and shotgun pellets can be cast with minimal technical equipment. It is also inexpensive and denser than other common metals.[109]

Because of its high density and resistance to corrosion, lead is used for the ballast keel of sailboats.[110] Its high density allows it to counterbalance the heeling effect of wind on the sails while at the same time occupying a small volume and thus offering the least underwater resistance. For the same reason it is used in scuba diving weight belts to counteract the diver's natural buoyancy and that of his equipment.[111] It does not have the weight-to-volume ratio of many heavy metals, but its low cost increases its use in these and other applications.

Roman lead water pipes with taps
Lead pipe in Roman baths
Multicolor lead-glazing in a Tang dynasty Chinese sancai ceramic cup dating from the 8th century CE
Punched lead cast in a Venice bridge wall fixing the hard-metal connecting bar

More than half of the US lead production (at least 1.15 million tonnes in 2000) is used for automobiles, mostly as electrodes in the lead–acid battery, used extensively as a car battery.[112]

Cathode (reduction)

PbO2 + 4 H+ + SO2−
4
+ 2e → PbSO4 + 2 H2O

Anode (oxidation)

Pb + SO2−
4
→ PbSO4 + 2e[113][114]

Lead is used as electrodes in the process of electrolysis. It is used in solder for electronics, although this usage is being phased out by some countries to reduce the amount of environmentally hazardous waste, and in high voltage power cables as sheathing material to prevent water diffusion into insulation. Lead is one of three metals used in the Oddy test for museum materials, helping detect organic acids, aldehydes, and acidic gases. It is also used as shielding from radiation (e.g., in X-ray rooms).[115] Molten lead is used as a coolant (e.g., for lead cooled fast reactors).[116]

Lead is added to brass to reduce machine tool wear. In the form of strips or tape, lead is used for the customization of tennis rackets. Tennis rackets in the past sometimes had lead added to them by the manufacturer to increase weight.[117] It is also used to form glazing bars for stained glass or other multi-lit windows. The practice has become less common, not for danger but for stylistic reasons. Lead, or sheet-lead, is used as a sound deadening layer in some areas in wall, floor and ceiling design in sound studios where levels of airborne and mechanically produced sound are targeted for reduction or virtual elimination.[118][119] It is the traditional base metal of organ pipes, mixed with varying amounts of tin to control the tone of the pipe.[120][121]

Lead has many uses in the construction industry (e.g., lead sheets are used as architectural metals in roofing material, cladding, flashing, gutters and gutter joints, and on roof parapets). Detailed lead moldings are used as decorative motifs used to fix lead sheet. Lead is still widely used in statues and sculptures. Lead is often used to balance the wheels of a car; this use is being phased out in favor of other materials for environmental reasons. Owing to its half-life of 22.20 years, the radioactive isotope 210Pb is used for dating material from marine sediment cores by radiometric methods.[122][123][124]

Compounds[edit]

Lead compounds are used as a coloring element in ceramic glazes, notably in the colors red and yellow.[125]

Lead tetraacetate (LTA) and lead dioxide have been used as oxidizing agents in organic chemistry. Geminal diols are cleaved to a pair of carbonyl compounds by stoichiometric LTA. LTA also is a selective oxidant of 5-methyl groups in 5-methylpyrrole-2-carboxylic esters, leading to 5-acetoxymethyl groups or 5-formyl groups with one or two equivalents of oxidant, respectively, to provide important intermediates for porphyrin synthesis.[126]

Lead is frequently used in polyvinyl chloride (PVC) plastic, which coats electrical cords.[127][128]

Lead is used in some candles to treat the wick to ensure a longer, more even burn. Because of the dangers, European and North American manufacturers use alternatives such as zinc.[129][130] Lead glass is composed of 12–28% lead oxide. It changes the optical characteristics of the glass and reduces the transmission of ionizing radiation.[131]

Some artists using oil-based paints continue to use lead carbonate white, citing its properties in comparison with the alternatives. Tetra-ethyl lead is used as an anti-knock additive for aviation fuel in piston-driven aircraft. Lead-based semiconductors, such as lead telluride, lead selenide and lead antimonide are finding applications in photovoltaic (solar energy) cells and infrared detectors.[132]

Historical applications[edit]

Lead pigments were used in lead paint for white as well as yellow, orange, and red. Most uses have been discontinued due to the dangers of lead poisoning. Beginning April 22, 2010, US federal law requires that contractors performing renovation, repair, and painting projects that disturb more than six square feet of paint in homes, child care facilities, and schools built before 1978 must be certified and trained to follow specific work practices to prevent lead contamination. Lead chromate is still in industrial use. Lead carbonate (white) is the traditional pigment for the priming medium for oil painting, but it has been largely displaced by the zinc and titanium oxide pigments. It was also quickly replaced in water-based painting mediums. Lead carbonate white was used by the Japanese geisha and in the West for face-whitening make-up, which was detrimental to health.[133][134][135]

Lead was the principal component of the alloy used in hot metal typesetting. It was used for plumbing (hence the name) as well as a preservative for food and drink in Ancient Rome. Until the early 1970s, lead was used for joining cast iron water pipes and used as a material for small diameter water pipes.[136]

Tetraethyllead was used in leaded fuels to reduce engine knocking, but this practice has been phased out across many countries of the world in efforts to reduce toxic pollution that affected humans and the environment.[137][138][139][140]

Lead was used to make bullets for slings. Lead is used for shotgun pellets (shot). Waterfowl hunting in the US with lead shot is illegal and it has been replaced with steel and other non-toxic shot for that purpose. In the Netherlands, the use of lead shot for hunting and sport shooting was banned in 1993, which caused a large drop in lead emission, from 230 tonnes in 1990 to 47.5 tonnes in 1995, two years after the ban.[141]

Lead was a component of the paint used on children's toys – now restricted in the United States and across Europe (ROHS Directive). Lead solder was used as a car body filler, which was used in many custom cars in the 1940s–60s; hence the term Leadsled. Lead is a superconductor with a transition temperature of 7.2 K, and therefore IBM tried to make a Josephson effect computer out of a lead alloy.[142]

Lead was also used in pesticides before the 1950s, when fruit orchards were treated especially against the codling moth.[143] A lead cylinder attached to a long line was used by sailors for the vital navigational task of determining water depth by heaving the lead at regular intervals. A soft tallow insert at its base allowed the nature of the sea bed to be determined, to assess its suitability for anchoring.[144]

Bioremediation[edit]

Fish bones are being researched for their ability to bioremediate lead in contaminated soil.[145][146] The fungus Aspergillus versicolor is both greatly effective and fast at removing lead ions.[147] Several bacteria have been researched for their ability to reduce lead; including the sulfate reducing bacteria Desulfovibrio and Desulfotomaculum; which are highly effective in aqueous solutions.[148]

Health effects[edit]

Main article: Lead poisoning

Lead is a highly poisonous metal (whether inhaled or swallowed), affecting almost every organ and system in the body. The component limit of lead (1.0 μg/g) is a test benchmark for pharmaceuticals, representing the maximum daily intake an individual should have. Even at this level, a prolonged intake can be hazardous to human beings.[149][150]

Much of its toxicity comes from how Pb2+ ions are confused for Ca2+ ions, and lead as a result gets into bones.

The main target for lead toxicity is the nervous system, both in adults and children. Long-term exposure of adults can result in decreased performance in some tests that measure functions of the nervous system.[151] Long-term exposure to lead or its salts (especially soluble salts or the strong oxidant PbO2) can cause nephropathy, and colic-like abdominal pains. It may also cause weakness in fingers, wrists, or ankles. Lead exposure also causes small increases in blood pressure, particularly in middle-aged and older people and can cause anemia. Exposure to high lead levels can cause severe damage to the brain and kidneys in adults or children and ultimately cause death. In pregnant women, high levels of exposure to lead may cause miscarriage. Chronic, high-level exposure has been shown to reduce fertility in males.[152] Lead also damages nervous connections (especially in young children) and causes blood and brain disorders. Lead poisoning typically results from ingestion of food or water contaminated with lead, but may also occur after accidental ingestion of contaminated soil, dust, or lead-based paint.[153] It is rapidly absorbed into the bloodstream and is believed to have adverse effects on the central nervous system, the cardiovascular system, kidneys, and the immune system.[154] The treatment for lead poisoning consists of dimercaprol and succimer.[155]

NFPA 704
"fire diamond"
Flammability code 1: Must be pre-heated before ignition can occur. Flash point over 93 °C (200 °F). E.g., canola oil Health code 3: Short exposure could cause serious temporary or residual injury. E.g., chlorine gas Reactivity code 0: Normally stable, even under fire exposure conditions, and is not reactive with water. E.g., liquid nitrogen Special hazards (white): no codeNFPA 704 four-colored diamond
Fire diamond for lead granules

The concern about lead's role in cognitive deficits in children has brought about widespread reduction in its use (lead exposure has been linked to learning disabilities).[156] Most cases of adult elevated blood lead levels are workplace-related.[157] High blood levels are associated with delayed puberty in girls.[158] Lead has been shown many times to permanently reduce the cognitive capacity of children at extremely low levels of exposure.[159]

During the 20th century, the use of lead in paint pigments was sharply reduced because of the danger of lead poisoning, especially to children.[160][161] By the mid-1980s, a significant shift in lead end-use patterns had taken place. Much of this shift was a result of the U.S. lead consumers' compliance with environmental regulations that significantly reduced or eliminated the use of lead in non-battery products, including gasoline, paints, solders, and water systems. Lead use is being further curtailed by the European Union's RoHS directive.[162] Lead may still be found in harmful quantities in stoneware,[163] vinyl[164] (such as that used for tubing and the insulation of electrical cords), and Chinese brass. Old houses may still contain substantial amounts of lead paint.[164] White lead paint has been withdrawn from sale in industrialized countries, but the yellow lead chromate is still in use. Old paint should not be stripped by sanding, as this produces inhalable dust.[165]

People can be exposed to lead in the workplace by breathing it in, swallowing it, skin contact, and eye contact. In the United States, the Occupational Safety and Health Administration (OSHA) has set the legal limit (permissible exposure limit) for lead exposure in the workplace as 0.050 mg/m3 over an 8-hour workday, which applies to metallic lead, inorganic lead compounds, and lead soaps. The National Institute for Occupational Safety and Health (NIOSH) has set a recommended exposure limit (REL) of 0.050 mg/m3 over an 8-hour workday, and recommends that workers' blood concentrations of lead stay below 0.060 mg per 100 g blood. At levels of 100 mg/m3, lead is immediately dangerous to life and health.[166]

Lead salts used in pottery glazes have on occasion caused poisoning, when acidic drinks, such as fruit juices, have leached lead ions out of the glaze.[167] It has been suggested that what was known as "Devon colic" arose from the use of lead-lined presses to extract apple juice in the manufacture of cider. Lead is considered to be particularly harmful for women's ability to reproduce. Lead(II) acetate (also known as sugar of lead) was used in the Roman Empire as a sweetener for wine, and some consider this a plausible explanation for the dementia of many Roman emperors, and that chronic lead poisoning contributed to the empire's gradual decline.[168]

Biochemistry of poisoning[edit]

In the human body, lead inhibits porphobilinogen synthase and ferrochelatase, preventing both porphobilinogen formation and the incorporation of iron into protoporphyrin IX, the final step in heme synthesis. This causes ineffective heme synthesis and subsequent microcytic anemia.[169] At lower levels, it acts as a calcium analog, interfering with ion channels during nerve conduction. This is one of the mechanisms by which it interferes with cognition. Acute lead poisoning is treated using disodium calcium edetate: the calcium chelate of the disodium salt of ethylene-diamine-tetracetic acid (EDTA). This chelating agent has a greater affinity for lead than for calcium and so the lead chelate is formed by exchange. This is then excreted in the urine leaving behind harmless calcium.[170] According to the Agency for Toxic Substance and Disease Registry, a small amount of ingested lead (1%) will store itself in bones, and the rest will be excreted by an adult through urine and feces within a few weeks of exposure. However, only about 32% of lead will be excreted by a child.[171]

Exposure to lead and lead chemicals can occur through inhalation, ingestion and dermal contact. Most exposure occurs through ingestion or inhalation; in the U.S. the skin exposure is unlikely as leaded gasoline additives are no longer used. Lead exposure is a global issue as lead mining and lead smelting are common in many countries. Most countries had stopped using lead-containing gasoline by 2007.[172] Lead exposure mostly occurs through ingestion. Lead paint is the major source of lead exposure for children. As lead paint deteriorates, it peels, is pulverized into dust and then enters the body through hand-to-mouth contact or through contaminated food, water or alcohol. Ingesting certain home remedy medicines may also expose people to lead or lead compounds.[172] Lead can be ingested through fruits and vegetables contaminated by high levels of lead in the soils they were grown in. Soil is contaminated through particulate accumulation from lead in pipes, lead paint and residual emissions from leaded gasoline that was used before the Environment Protection Agency issued the regulation around 1980.[173] The use of lead for water pipes is problematic in areas with soft or (and) acidic water. Hard water forms insoluble layers in the pipes while soft and acidic water dissolves the lead pipes.[174] Inhalation is the second major pathway of exposure, especially for workers in lead-related occupations. Almost all inhaled lead is absorbed into the body, the rate is 20–70% for ingested lead; children absorb more than adults.[172] Dermal exposure may be significant for a narrow category of people working with organic lead compounds, but is of little concern for general population. The rate of skin absorption is also low for inorganic lead.[172]

See also[edit]

Notes[edit]

  1. ^ Note that in contexts related to singular atoms and elements, words "heavy" and "light" normally refer to atomic numbers and not densities of the substances these elements form.
  2. ^ An even number of either protons or neutrons generally increases nuclear stability of isotopes, compared to isotopes with odd such numbers. For example, elements with odd atomic numbers have no more than two stable isotopes, while even-numbered elements have multiple stable isotopes, with tin (element 50) having the highest number of isotopes of all elements, ten.[12] See Even and odd atomic nuclei for more details.
  3. ^ The half-life found in the experiment was 1.9×1019 years.[13] A kilogram of natural bismuth, would thus be radioactive with an activity value of approximately 0.003 becquerels—decays per second. For comparison, the natural radiation within human body would make an adult human have radioactivity of 65 becquerels per kilogram of body weight (around 4500 becquerels on average).[14]
  4. ^ However, it decays via electron capture, which means when there are no electrons available and lead is accordingly fully ionized—has all 82 electrons removed—it cannot decay and becomes stable. Moreover, thallium-205, the isotope lead-205 would decay to, becomes unstable with respect to decaying into lead-205.[17]
  5. ^ About 10% of the lanthanide contraction has also been attributed to relativistic effects.[20]
  6. ^ However, in the oxidized zones of lead deposits, small quantities of lead(IV) species can be found, including the oxide minerals plattnerite, scrutinyite, and murdochite.
  7. ^ "Made when the Emperor Vespasian was consul for the ninth time and the Emperor Titus was consul for the seventh time, when Gnaeus Iulius Agricola was imperial governor [of Britain]".
  8. ^ It is suggested that the sweeteners the Romans made were often prepared in lead vessels; this led to formation of poisonous lead(II) acetate, which accumulated in the sweeteners and, accordingly, the products they were used for; in particular, wine. Lead containers did further sweeten the content as well as helped preserve it.[68] In comparison, copper vessels did throw off rust, which spoiled the taste of wine in them. The fact that Julius Caesar managed to incept only one child, as well as alleged sterility of his successor, Caesar Augustus, have also been alleged to lead poisoning.[58] However, the theory is criticized for the Romans were aware of the potential health problems lead could cause, as well as the fact copper has been used far more commonly for vessels than lead.
  9. ^ It is not known when mining was first performed in the region because no tradition of keeping written records was in place, but there are European 17th century records of trade with the Congolese, which indicates lead was first smelted no later than then.[83]

References[edit]

  1. ^ Standard Atomic Weights 2013. Commission on Isotopic Abundances and Atomic Weights
  2. ^ a b c d e f Polyanskiy 1986, p. 18.
  3. ^ Thurmer, K.; Williams, E; Reutt-Robey, J. (2002). "Autocatalytic Oxidation of Lead Crystallite Surfaces". Science 297 (5589): 2033–5. Bibcode:2002Sci...297.2033T. doi:10.1126/science.297.5589.2033. PMID 12242437. 
  4. ^ Tétreault, Jean; Sirois, Jane; Stamatopoulou, Eugénie (1998). "Studies of Lead Corrosion in Acetic Acid Environments". Studies in Conservation 43 (1): 17–32. doi:10.2307/1506633. JSTOR 1506633. 
  5. ^ Polyanskiy 1986, p. 14.
  6. ^ a b c Lide 2004, p. 12-220.
  7. ^ a b Lide 2004, p. 4-13.
  8. ^ Lide 2004, p. 12-219.
  9. ^ a b Lide 2004, p. 12-35.
  10. ^ Lide 2004, p. 12-37.
  11. ^ a b Polyanskiy 1986, p. 16.
  12. ^ a b c d e G. Audi; A. H. Wapstra; C. Thibault; J. Blachot & O. Bersillon (2003). "The NUBASE evaluation of nuclear and decay properties" (PDF). Nuclear Physics A 729 (1): 3–128. Bibcode:2003NuPhA.729....3A. doi:10.1016/j.nuclphysa.2003.11.001. 
  13. ^ Marcillac, Pierre de; Noël Coron; Gérard Dambier; Jacques Leblanc & Jean-Pierre Moalic (2003). "Experimental detection of α-particles from the radioactive decay of natural bismuth". Nature 422 (6934): 876–878. Bibcode:2003Natur.422..876D. doi:10.1038/nature01541. PMID 12712201. 
  14. ^ "Nuclear Radiation and Health Effects". World Nuclear Association. 2015. Retrieved 2015-11-12. 
  15. ^ a b Greenwood & Earnshaw 1998, p. 368.
  16. ^ UC Berkeley Nuclear Forensic Search Project. "Decay Chains". Nuclear Forensics: A Scientific Search Problem. Retrieved 2015-11-23. 
  17. ^ Takahashi, K; Boyd, R. N.; Mathews, G. J.; Yokoi, K. (October 1987). "Bound-state beta decay of highly ionized atoms" (PDF). Physical Review C (American Institute of Physics for the American Physical Society) 36 (4). ISSN 0556-2813. OCLC 1639677. Retrieved 2013-08-27. 
  18. ^ Lide 2004, p. 10-179.
  19. ^ a b c d Polyanskiy 1986, pp. 14–15.
  20. ^ Pekka Pyykko (1988). "Relativistic effects in structural chemistry". Chem. Rev. 88 (3): 563–594. doi:10.1021/cr00085a006. 
  21. ^ Synthesis of Organometallic Compounds: A Practical Guide Sanshiro Komiya Ed. 1997
  22. ^ a b c Polyanskiy 1986, p. 20.
  23. ^ Rappoport, Zvi; Marek, Ilan (2010). The Chemistry of Organocopper Compounds. John Wiley & Sons. p. 509. ISBN 978-0-470-77296-6. 
  24. ^ Charles, J.; Kopf, P. W.; Toby, S. (1966). "The Reaction of Pyrophoric Lead with Oxygen". Journal of Physical Chemistry 70 (5): 1478–1482. doi:10.1021/j100877a023. 
  25. ^ a b Polyanskiy 1986, p. 19.
  26. ^ a b c Greenwood & Earnshaw 1998, p. 373.
  27. ^ Greenwood & Earnshaw 1998, p. 374.
  28. ^ a b c Polyanskiy 1986, p. 32.
  29. ^ .Ensafi, Ali A.; Far, A. Katiraei; Meghdadi, S. (2009). "Highly selective optical-sensing film for lead(II) determination in water samples". Journal of Hazardous Materials 172 (2–3): 1069–75. doi:10.1016/j.jhazmat.2009.07.112. PMID 19709813. 
  30. ^ Greenwood & Earnshaw 1998, p. 384.
  31. ^ Lewis, Alison E. (2010). "Review of metal sulphide precipitation" (PDF). Hydrometallurgy 104 (2): 222–234. doi:10.1016/j.hydromet.2010.06.010. Retrieved 2014-10-14. 
  32. ^ a b Pauling, Linus (1947). General Chemistry. W.H. Freeman. ISBN 0-486-65622-5. 
  33. ^ Greenwood & Earnshaw 1998, p. 389.
  34. ^ Zuckerman, J. J.; Hagen, A. P. (1989). Inorganic Reactions and Methods, the Formation of Bonds to Halogens. John Wiley & Sons. p. 426. ISBN 978-0-471-18656-4. 
  35. ^ Greenwood & Earnshaw 1998, p. 398.
  36. ^ Macomber 1996, p. 230
  37. ^ Cava, R. J.; Hor, Y.S.; Cava, R.J. (2011). "Pressure Stabilized Se-Se Dimer Formation in PbSe2". Solid State Sciences 13: 38–41. Bibcode:2011SSSci..13...38B. doi:10.1016/j.solidstatesciences.2010.10.003. 
  38. ^ Silverman, M. S. (1966). "High-pressure (70-kilobar) Synthesis of New Crystalline Lead Dichalcogenides". Inorganic Chemistry 5 (11): 2067–9. doi:10.1021/ic50045a056. 
  39. ^ Becker, Marco; FöRster, Christoph; Franzen, Christian; Hartrath, Johannes; Kirsten, Enzio; Knuth, Jörn; Klinkhammer, Karl W.; Sharma, Ajay; Hinderberger, Dariush (2008). "Persistent Radicals of Trivalent Tin and Lead". Inorganic Chemistry 47 (21): 9965–78. doi:10.1021/ic801198p. PMID 18823115. 
  40. ^ Mosseri, Shlomo.; Henglein, Arnim.; Janata, Eberhard. (1990). "Trivalent lead as an intermediate in the oxidation of lead(II) and the reduction of lead(IV) species". The Journal of Physical Chemistry 94 (6): 2722–2726. doi:10.1021/j100369a089. 
  41. ^ Siew-Peng Chia; Hong-Wei Xi; Yongxin Li; Kok Hwa Lim; Cheuk-Wai So (2013). "A Base-Stabilized Lead(I) Dimer and an Aromatic Plumbylidenide Anion". Angew. Chem. Int. Ed. 52 (24): 6298–6301. doi:10.1002/anie.201301954. 
  42. ^ Universität Freiburg. "Binäre Zintl-Phasen" (in German). 
  43. ^ Alsfasser, Ralf (2007). Moderne anorganische Chemie: mit CD-ROM (in German). Walter de Gruyter. pp. 261–263. ISBN 978-3-11-019060-1. 
  44. ^ Stabenow, Frank; Saak, Wolfang; Weidenbruch, Manfred (2003). "Tris(triphenylplumbyl)plumbate: An anion with three stretched lead?lead bonds". Chemical Communications (18): 2342. doi:10.1039/B305217F. 
  45. ^ a b c Polyanskiy 1986, p. 43.
  46. ^ a b c Wiberg, Egon; Wiberg, Nils; Holleman, Arnold Frederick (2001). Inorganic Chemistry. Academic Press. p. 918. ISBN 978-0-12-352651-9. 
  47. ^ Hein, Thomas A.; Thiel, Walter; Lee, Timothy J. (1993). "Ab initio study of the stability and vibrational spectra of plumbane, methylplumbane, and homologous compounds". The Journal of Physical Chemistry 97 (17): 4381–4385. doi:10.1021/j100119a021. 
  48. ^ a b Polyanskiy 1986, p. 44.
  49. ^ Greenwood & Earnshaw 1998, p. 405.
  50. ^ Windholz, Martha (1976). Merck Index of Chemicals and Drugs (9th ed.). Merck. ISBN 0-911910-26-3. Monograph 8393. 
  51. ^ a b E. M. Burbidge; G. R. Burbidge; W. A. Fowler; F. Hoyle (1957). "Synthesis of the Elements in Stars" (PDF). Reviews of Modern Physics 29 (4): 547. Bibcode:1957RvMP...29..547B. doi:10.1103/RevModPhys.29.547. 
  52. ^ a b c Cameron, A. G. W. (1973). "Abundance of the Elements in the Solar System" (PDF). Space Science Review 15: 121–146. Bibcode:1973SSRv...15..121C. doi:10.1007/BF00172440. 
  53. ^ a b Winter, Mark. "Lead»geological information [WebElements Periodic Table]". www.webelements.com. Retrieved 2016-03-09. 
  54. ^ a b c d Sutherland et al. 2005, p. 5.
  55. ^ a b c U.S. Geological Survey (2016). Lead (PDF) (Report). Retrieved 2016-02-20. 
  56. ^ Rich 1994, p. 4.
  57. ^ a b c d "The History of Lead - Part 3". 2007-08-31. Archived from the original on August 31, 2007. Retrieved 2016-02-12. 
  58. ^ a b "Lead in history". corrosion-doctors.org. Retrieved 2016-02-13. 
  59. ^ "Toronto museum explores history of contraceptives". ABC News. Retrieved 2016-02-13. 
  60. ^ Bisson & Vogel 2000, p. 105.
  61. ^ Sutherland et al. 2005, p. 2.
  62. ^ "Exploring Chinese History :: History :: Ancient Chinese History :: Outline". www.ibiblio.org. Retrieved 2016-02-16. 
  63. ^ a b c d Hong, Sungmin; Candelone, Jean-Pierre; Patterson, Clair Cameron; Boutron, Claude F. (1994). "Greenland Ice Evidence of Hemispheric Lead Pollution Two Millennia Ago by Greek and Roman Civilizations" (PDF). Science 265 (5180): 1841–1843. Bibcode:1994Sci...265.1841H. doi:10.1126/science.265.5180.1841. PMID 17797222. 
  64. ^ Callataÿ, François de (2005). "The Graeco-Roman Economy in the Super Long-Run: Lead, Copper, and Shipwrecks". Journal of Roman Archaeology 18: 361–372 (361–365). 
  65. ^ Settle, Dorothy M.; Patterson, Clair C. (1980). "Lead in Albacore: Guide to Lead Pollution in Americans". Science 207 (4436): 1167–1176. Bibcode:1980Sci...207.1167S. doi:10.1126/science.6986654. PMID 6986654.  see 1170f.
  66. ^ "lead". vanderkrogt.net. Retrieved 2012-06-02. 
  67. ^ Hodge, Trevor, A. (October 1981). "Vitruvius, Lead Pipes and Lead Poisoning". American Journal of Archaeology (Archaeological Institute of America) 85 (4): 486–491. JSTOR 504874. 
  68. ^ "The History of Lead Part 4". 2015-04-05. Archived from the original on April 5, 2015. Retrieved 2016-02-12. 
  69. ^ a b EPA,OA,OEAEE,OWC, US. "Lead Poisoning: A Historical Perspective". www.epa.gov. Retrieved 2016-02-07. 
  70. ^ Polyanskiy 1986, p. 8.
  71. ^ Thomson, Thomas (1830-01-01). The History of Chemistry. H. Colburn and R. Bentley. p. 74. 
  72. ^ Squatriti, Paolo, ed. (2000). Working with water in medieval Europe: technology and resource use. Brill. pp. 134 ff. ISBN 978-90-04-10680-2. 
  73. ^ a b "The History Of Lead part 1". 2007-08-31. Archived from the original on August 31, 2007. Retrieved 2016-02-05. 
  74. ^ a b Rich 1994, p. 7.
  75. ^ Rehder, Dieter (2011-08-02). Chemistry in Space. p. 104. ISBN 978-3-527-63238-1. 
  76. ^ "Printing: History and Development". karmak.org. Retrieved 2016-02-13. 
  77. ^ Sinha, S.P.; Shelly; Sharma, Vibha; Meenakshi; Srivastava, Shalini; Srivastava, M.M. (1993). "Neurotoxic effects of lead exposure among printing press workers". Bulletin of Environmental Contamination and Toxicology 51 (4). doi:10.1007/BF00192162. 
  78. ^ "10 Innovations That Led to the Modern Bullet". HowStuffWorks. Retrieved 2016-02-13. 
  79. ^ Ramage, C. Kenneth (1980). Lyman Cast Bullet Handbook, 3rd Edition. Lyman Publications. p. 8. 
  80. ^ Beard, Michael E. (1995). Lead in Paint, Soil, and Dust: Health Risks, Exposure Studies, Control Measures, Measurement Methods, and Quality Assurance. ASTM International. p. 66. ISBN 9780803118843. 
  81. ^ "Request Rejected". www.mininghistory.asn.au. Retrieved 2016-02-15. 
  82. ^ Bisson & Vogel 2000, p. 85.
  83. ^ a b Bisson & Vogel 2000, pp. 131–132.
  84. ^ "Lead Mining news from the The Northern Echo". www.thenorthernecho.co.uk. Retrieved 2016-02-16. 
  85. ^ "Lead: Versatile Metal, Long Legacy - Dartmouth Toxic Metals Superfund Research Program". www.dartmouth.edu. Retrieved 2016-02-16. 
  86. ^ Rich 1994, p. 11.
  87. ^ a b c d Riva, Michele Augusto; Lafranconi, Alessandra; d'Orso, Marco Italo; Cesana, Giancarlo (2012). "Lead Poisoning: Historical Aspects of a Paradigmatic "Occupational and Environmental Disease"". Safety and Health at Work 3 (1): 11–6. doi:10.5491/SHAW.2012.3.1.11. PMC 3430923. PMID 22953225. 
  88. ^ "Lead poisoning in a historical perspective" (PDF). PMID 10940962. 
  89. ^ ""Cater to the children": the role of the lead industry in a public health tragedy, 1900-1955.". PMC 1446124. 
  90. ^ Rich 1994, p. 117.
  91. ^ Rich 1994, p. 17.
  92. ^ Rich 1994, pp. 91–92.
  93. ^ U.S. Geological Survey (2005). Lead (PDF) (Report). Retrieved 2016-02-20. 
  94. ^ . doi:10.1590/S0042-96862000000900003.  Missing or empty |title= (help)
  95. ^ a b Guberman, David E. (2015). "Lead". 2013 Minerals Yearbook (PDF) (Report). U.S. Geological Survey. 
  96. ^ CITEREFSutherland_et_al.2005
  97. ^ "Primary Extraction of Lead Technical Notes". LDA International. Archived from the original on 22 March 2007. Retrieved 7 April 2007. 
  98. ^ CITEREFSutherland_et_al.2005
  99. ^ a b c d e Sutherland, Charles A.; Milner, Edward F.; Kerby, Robert C.; Teindl, Herbert; Melin, Albert; Bolt, Hermann M. (2005). "Lead". Ullmann's Encyclopedia of Industrial Chemistry. doi:10.1002/14356007.a15_193.pub2. ISBN 3-527-30673-0. 
  100. ^ a b c d "Primary Lead Refining Technical Notes". LDA International. Archived from the original on 22 March 2007. Retrieved 7 April 2007. 
  101. ^ Reilly, Michael (2007). "How Long Will it Last?". New Scientist 194 (2605): 38–39. Bibcode:2007NewSc.194...38R. doi:10.1016/S0262-4079(07)61508-5. ISSN 0262-4079. 
  102. ^ Brown, Lester (2006). Plan B 2.0: Rescuing a Planet Under Stress and a Civilization in Trouble. New York: W.W. Norton. p. 109. ISBN 0-393-32831-7. 
  103. ^ "Metal stocks in Society – Scientific Synthesis" (PDF). International Resource Panel. Retrieved 2012-07-02. 
  104. ^ Fiorini, Ettor. 2.000 years-old Roman Lead for physics. INFN-Milano Bicocca
  105. ^ Nosengo, Nicola (2010). "Roman ingots to shield particle detector". Nature News. doi:10.1038/news.2010.186. 
  106. ^ "A history of pencils". www.pencils.com. Retrieved 7 April 2007. 
  107. ^ Electro-Plating on Non-Metallic Substances. Spons' Workshop Receipts. Vol. II: Dyeing to Japanning. Spon. 1921. p. 132. 
  108. ^ Evans, John W. (1908). "V.— the Meanings and Synonyms of Plumbago". Transactions of the Philological Society 26 (2): 133–179. doi:10.1111/j.1467-968X.1908.tb00513.x. 
  109. ^ Rooney, Corinne. "Contamination at Shooting Ranges" (PDF). The Lead Group, incorporated. Retrieved 7 April 2007. 
  110. ^ "Lead Ballast". Lead Ballast. Lead Ballast. 2007. Retrieved 3 July 2012. 
  111. ^ "Lead Shot Ballast". Lead Shot Ballast. Lead Shot Ballast. 2007. Retrieved 3 July 2012. 
  112. ^ Getting the Lead Out: Impacts of and Alternatives For Automotive Lead Uses, A report by Environmental Defense, Ecology Center, Clean Car Campaign (July 2003)
  113. ^ Crompton, T.R. (2000). Battery reference book. Newnes. pp. 18/2–18/4. ISBN 978-0-7506-4625-3. 
  114. ^ Stellman, Jeanne Mager (1998). Encyclopaedia of Occupational Health and Safety. International Labour Organization. pp. 81.2–81.4. ISBN 978-92-2-109816-4. 
  115. ^ Structural shielding design for medical X-ray imaging facilities. National Council on Radiation Protection and Measurement. 2004. pp. 16–17. ISBN 978-0-929600-83-3. 
  116. ^ Tuček, Kamil; Carlsson, Johan; Wider, Hartmut (2006). "Comparison of sodium and lead-cooled fast reactors regarding reactor physics aspects, severe safety and economical issues" (PDF). Nuclear Engineering and Design 236 (14–16): 1589–1598. doi:10.1016/j.nucengdes.2006.04.019. 
  117. ^ Hong, Youlian; Bartlett, Roger, eds. (2008). Routledge Handbook of Biomechanics and Human Movement Science. Routledge. p. 250. ISBN 978-0-415-40881-3. 
  118. ^ Guruswamy, Sivaraman (2000). Engineering properties and applications of lead alloys. Marcel Dekker. p. 31. ISBN 978-0-8247-8247-4. 
  119. ^ Lansdown, Richard; Yule, William, eds. (1986). The Lead debate : the environment, toxicology, and child health. Croom Helm. p. 240. ISBN 978-0-7099-1653-6. 
  120. ^ Audsley, George Ashdown (1988-04-01). The Art of Organ Building 2. pp. 250–251. ISBN 978-0-486-21315-6. 
  121. ^ Palmieri, Robert, ed. (2006). The Organ. Garland. pp. 412–413. ISBN 978-0-415-94174-7. 
  122. ^ http://www.nrcresearchpress.com/doi/abs/10.1139/f83-069 Evans, R. Douglas; Rigler, Frank H. (1983). "A Test of Lead-210 Dating for the Measurement of Whole Lake Soft Sediment Accumulation". Canadian Journal of Fisheries and Aquatic Sciences 40 (4): 506–515. doi:10.1139/f83-069. Retrieved 2012-07-03. 
  123. ^ "Dating of Sediments using Lead-210". dhigroup.com. DHI. Retrieved 2012-07-03. 
  124. ^ Noller, Jay Stratton (2000). "Lead-210 Geochronology". Quaternary Geochronology: Methods and Applications. pp. 115–120. ISBN 978-0-87590-950-9. 
  125. ^ Leonard, Alvin R.; Lynch, Glenn (1958). "Dishware as a Possible Source for Lead Poisoning". Calif. Med. 89 (6): 414–416. PMC 1512529. PMID 13608300. 
  126. ^ Siedel, Walter; Winkler, Fritz (1943). "Oxydation von Pyrrolderivaten mit Bleitetraacetat. Neuartige Porphyrinsynthesen". Justus Liebig's Annalen der Chemie 554 (1): 162–201. doi:10.1002/jlac.19435540111. ISSN 0075-4617. 
  127. ^ Zweifel, Hans (2009). Plastics Additives Handbook. Hanser Verlag. p. 438. ISBN 978-3-446-40801-2. 
  128. ^ Wilkes, C. E.; Summers, J. W.; Daniels, C. A.; Berard, M. T. (2005). PVC handbook. Hanser. p. 106. ISBN 978-1-56990-379-7. 
  129. ^ Randerson, James (June 2002). "Candle pollution". NewScientist.com (2348). Retrieved 2007-04-07. 
  130. ^ Nriagu, J; Kim, MJ (2000). "Emissions of lead and zinc from candles with metal-core wicks". The Science of the Total Environment 250 (1–3): 37–41. doi:10.1016/S0048-9697(00)00359-4. PMID 10811249. 
  131. ^ Amstock, Joseph S. (1997). Handbook of glass in construction. McGraw-Hill Professional. pp. 116–119. ISBN 978-0-07-001619-4. 
  132. ^ "Applications for Lead". Retrieved 7 April 2007. 
  133. ^ Nakashima, T; Matsuno, K; Matsushita, T (2007). "Lifestyle-determined gender and hierarchical differences in the lead contamination of bones from a feudal town of the Edo period". Journal of occupational health 49 (2): 134–9. doi:10.1539/joh.49.134. PMID 17429171. 
  134. ^ Nakashima, Tamiji; Hayashi, Haruki; Tashiro, Hiraku; Matsushita, Takayuki (1998). "Gender and Hierarchical Differences in Lead-Contaminated Japanese Bone from the Edo Period". Journal of Occupational Health 40: 55–60. doi:10.1539/joh.40.55. 
  135. ^ Ashikari, Mikiko (2003). "The memory of the women's white faces: Japaneseness and the ideal image of women". Japan Forum 15: 55–79. doi:10.1080/0955580032000077739. 
  136. ^ Hernberg, S (2000). "Lead poisoning in a historical perspective". American journal of industrial medicine 38 (3): 244–54. doi:10.1002/1097-0274(200009)38:3<244::AID-AJIM3>3.0.CO;2-F. PMID 10940962. 
  137. ^ "Lead replacement petrol phase-out – Information to motorists". Department for Transport (gov.uk). Archived from the original on 2009-05-20. 
  138. ^ "National phase out of leaded petrol: Some questions and answers". Department of the Environment and Heritage, Australian Government. 2001. 
  139. ^ "Oregon Stations Phase Out Use of Leaded Gasoline.(Originated from The Register-Guard, Eugene, Ore.)". Knight Ridder/Tribune Business News. 4 October 1995. Retrieved 23 September 2008. 
  140. ^ Seyferth, Dietmar (2003). "The Rise and Fall of Tetraethyllead. 2". Organometallics 22 (25): 5154–5178. doi:10.1021/om030621b. 
  141. ^ "Lood en zinkemissies door jacht" (PDF) (in Dutch). April 2010. 
  142. ^ Henkels, W. H.; Geppert, L. M.; Kadlec, J.; Epperlein, P. W.; Beha, H.; Chang, W. H.; Jaeckel, H. (September 1985). "Josephson 4 K-bit cache memory design for a prototype signal processor". Journal of Applied Physics (Harvard University) 58 (6): 2371. Bibcode:1985JAP....58.2371H. doi:10.1063/1.335960. ISSN 0021-8979. 
  143. ^ Tollestrup, Kristine; Daling, Janet R.; Allard, Jack (1995). "Mortality in a Cohort of Orchard Workers Exposed to Lead Arsenate Pesticide Spray". Archives of Environmental Health 50 (3): 221–229. doi:10.1080/00039896.1995.9940391. 
  144. ^ Burney, William (1830). A New Universal Dictionary of the Marine: Being, a Copious Explanation of the Technical Terms and Phrases ... With Such Parts of Astronomy, and Navigation, as Will be Found Useful to Practical Navigators. ... Together with Separate Views of the Masts, Yards, Sails, and Rigging. To which is Annexed a Vocabulary of French Sea-phrases and Terms of Art. p. 490. 
  145. ^ Kris S. Freeman (January 2012). "Remediating Soil Lead with Fishbones". Environmental Health Perspectives 120 (1): a20–a21. doi:10.1289/ehp.120-a20a. PMC 3261960. PMID 22214821. 
  146. ^ "Battling lead contamination, one fish bone at a time". Compass. 
  147. ^ Bairagi, HImadri; Motiar Khan; Lalitagauri Ray; Arun Guha (February 2011). "Adsorption profile of lead on Aspergillus versicolor: A mechanistic probing". Journal of Hazardous Materials 186 (1): 756–764. doi:10.1016/j.jhazmat.2010.11.064. PMID 21159429. 
  148. ^ Jin Hee Park; Nanthi Bolan; Mallavarapu Meghara; Ravi Naidu; Jae Woo Chung (2011). "Bacterial-Assisted Immobilization of Lead in Soils: Implications for Remediation" (PDF). Pedologist: 162–174. 
  149. ^ Heavy Metals Testing By Usp. Caspharma.com. Retrieved on 2012-01-23.
  150. ^ pharmaceutical – Britannica Online Encyclopedia. Britannica.com. Retrieved on 2012-01-23.
  151. ^ "Lead in Air". 
  152. ^ Golub, Mari S., ed. (2005). "Summary". Metals, fertility, and reproductive toxicity. Taylor and Francis. p. 153. ISBN 978-0-415-70040-5. 
  153. ^ "ToxFAQs: CABS/Chemical Agent Briefing Sheet: Lead" (PDF). Agency for Toxic Substances and Disease Registry/Division of Toxicology and Environmental Medicine. 2006. Archived from the original (PDF) on 2010-03-04. 
  154. ^ Bergeson, Lynn L. (2008). "The proposed lead NAAQS: Is consideration of cost in the clean air act's future?". Environmental Quality Management 18: 79–84. doi:10.1002/tqem.20197. 
  155. ^ Jagadish Prasad, P. (2010). Conceptual Pharmacology. Universities Press. p. 652. ISBN 978-81-7371-679-9. Retrieved 21 June 2012. 
  156. ^ Hu, Howard (1991). "Knowledge of diagnosis and reproductive history among survivors of childhood plumbism". American Journal of Public Health 81 (8): 1070–1072. doi:10.2105/AJPH.81.8.1070. PMC 1405695. PMID 1854006. 
  157. ^ "NIOSH Adult Blood Lead Epidemiology and Surveillance". United States National Institute for Occupational Safety and Health. Retrieved 2007-10-04. 
  158. ^ Schoeters, Greet; Den Hond, Elly; Dhooge, Willem; Van Larebeke, Nik; Leijs, Marike (2008). "Endocrine Disruptors and Abnormalities of Pubertal Development". Basic & Clinical Pharmacology & Toxicology 102 (2): 168–175. doi:10.1111/j.1742-7843.2007.00180.x. PMID 18226071. 
  159. ^ Needleman, Herbert L.; Schell, Alan; Bellinger, David; Leviton, Alan; Allred, Elizabeth N. (1990). "The long-term effects of exposure to low doses of lead in childhood. An 11-year follow-up report". New England Journal of Medicine 322 (2): 83–88. doi:10.1056/NEJM199001113220203. PMID 2294437. 
  160. ^ "Download: Lead paint: Cautionary note" (PDF). Queensland Government. Retrieved 7 April 2007. 
  161. ^ "Lead Paint Information". Master Painters, Australia. Archived from the original on 2008-02-12. Retrieved 7 April 2007. 
  162. ^ Smith, Donald R.; Flegal, A. Russell. "Lead in the Biosphere: Recent Trends" 24: 21–23. JSTOR 4314280. 
  163. ^ Grandjean, P. (1978). "Widening perspectives of lead toxicity". Environmental Research 17 (2): 303–321. doi:10.1016/0013-9351(78)90033-6. PMID 400972. 
  164. ^ a b Levin, R.; Brown, M. J.; Kashtock, M. E.; et al. (2008). "Lead Exposures in U.S. Children, 2008: Implications for Prevention". Environmental Health Perspectives 116 (10): 1285–1293. doi:10.1289/ehp.11241. PMC 2569084. PMID 18941567. 
  165. ^ Marino, P. E.; Landrigan, P. J.; Graef, J.; Nussbaum, A.; Bayan, G.; Boch, K.; Boch, S. (1990). "A case report of lead paint poisoning during renovation of a Victorian farmhouse". American Journal of Public Health 80 (10): 1183–1185. doi:10.2105/AJPH.80.10.1183. PMC 1404824. PMID 2119148. 
  166. ^ "CDC – NIOSH Pocket Guide to Chemical Hazards - Lead". www.cdc.gov. Retrieved 2015-11-19. 
  167. ^ "CPG Sec. 545.450 Pottery (Ceramics); Import and Domestic – Lead Contamination". U.S. Food and Drug Administration. Retrieved 2010-02-02. 
  168. ^ Angier, Natalie (August 21, 2007). "The Pernicious Allure of Lead". New York Times. Retrieved 7 May 2010. 
  169. ^ Cohen, Alan R.; Trotzky, Margret S.; Pincus, Diane (1981). "Reassessment of the Microcytic Anemia of Lead Poisoning". Pediatrics 67 (6): 904–906. PMID 7232054. 
  170. ^ Laurence, D. R. (1966). Clinical Pharmacology (Third ed.). 
  171. ^ "Toxic Substances Portal – Lead". Agency for Toxic Substance and Disease Registry. Archived from the original on June 6, 2011. 
  172. ^ a b c d "Case Studies in Environmental Medicine Lead (Pb) Toxicity: How are People Exposed to Lead?". Agency for Toxic Substances and Disease Registry. Archived from the original on 2011-06-06. 
  173. ^ "Information for the Community Lead Toxicity". Agency for Toxic Substances and Disease Registry. 
  174. ^ Moore, Michael R. (1977). "Lead in drinking water in soft water areas—health hazards". Science of the Total Environment 7 (2): 109–15. doi:10.1016/0048-9697(77)90002-X. PMID 841299. 

Bibliography[edit]

Further reading[edit]

External links[edit]