Metalloid

Source: Wikipedia, the free encyclopedia.
  13 14
15
16 17
2 B
Boron
C
Carbon
N
Nitrogen
O
Oxygen
F
Fluorine
3 Al
Aluminium
Si
Silicon
P
Phosphorus
S
Sulfur
Cl
Chlorine
4 Ga
Gallium
Ge
Germanium
As
Arsenic
Se
Selenium
Br
Bromine
5 In
Indium
Sn
Tin
Sb
Antimony
Te
Tellurium
I
Iodine
6 Tl
Thallium
Pb
Lead
Bi
Bismuth
Po
Polonium
At
Astatine
 
  Commonly recognized (86–99%): B, Si, Ge, As, Sb, Te
  Irregularly recognized (40–49%): Po, At
  Less commonly recognized (24%): Se
  Rarely recognized (8–10%): C, Al
  (All other elements cited in less than 6% of sources)
  Arbitrary metal-nonmetal dividing line: between Be and B, Al and Si, Ge and As, Sb and Te, Po and At

Recognition status, as metalloids, of some elements in the p-block of the periodic table. Percentages are median appearance frequencies in the lists of metalloids.[n 1] The staircase-shaped line is a typical example of the arbitrary metal–nonmetal dividing line found on some periodic tables.

A metalloid is a type of

nonmetals. There is no standard definition of a metalloid and no complete agreement on which elements are metalloids. Despite the lack of specificity, the term remains in use in the literature of chemistry
.

The six commonly recognised metalloids are

p-block extending from boron at the upper left to astatine at lower right. Some periodic tables include a dividing line between metals and nonmetals
, and the metalloids may be found close to this line.

Typical metalloids have a metallic appearance, but they are brittle and only fair

, and electronics.

The electrical properties of silicon and germanium enabled the establishment of the semiconductor industry in the 1950s and the development of solid-state electronics from the early 1960s.[1]

The term metalloid originally referred to nonmetals. Its more recent meaning, as a category of elements with intermediate or hybrid properties, became widespread in 1940–1960. Metalloids are sometimes called semimetals, a practice that has been discouraged,[2] as the term semimetal has a different meaning in physics than in chemistry. In physics, it refers to a specific kind of electronic band structure of a substance. In this context, only arsenic and antimony are semimetals, and commonly recognised as metalloids.

Definitions

Judgment-based

A metalloid is an element that possesses a preponderance of properties in between, or that are a mixture of, those of metals and nonmetals, and which is therefore hard to classify as either a metal or a nonmetal. This is a generic definition that draws on metalloid attributes consistently cited in the literature.[n 2] Difficulty of categorisation is a key attribute. Most elements have a mixture of metallic and nonmetallic properties,[9] and can be classified according to which set of properties is more pronounced.[10][n 3] Only the elements at or near the margins, lacking a sufficiently clear preponderance of either metallic or nonmetallic properties, are classified as metalloids.[14]

Boron, silicon, germanium, arsenic, antimony, and tellurium are commonly recognised as metalloids.[15][n 4] Depending on the author, one or more from selenium, polonium, or astatine are sometimes added to the list.[17] Boron sometimes is excluded, by itself, or with silicon.[18] Sometimes tellurium is not regarded as a metalloid.[19] The inclusion of antimony, polonium, and astatine as metalloids has been questioned.[20]

Other elements are occasionally classified as metalloids. These elements include[21] hydrogen,[22] beryllium,[23] nitrogen,[24] phosphorus,[25] sulfur,[26] zinc,[27] gallium,[28] tin, iodine,[29] lead,[30] bismuth,[19] and radon.[31] The term metalloid has also been used for elements that exhibit metallic lustre and electrical conductivity, and that are amphoteric, such as arsenic, antimony, vanadium, chromium, molybdenum, tungsten, tin, lead, and aluminium.[32] The p-block metals,[33] and nonmetals (such as carbon or nitrogen) that can form alloys with metals[34] or modify their properties[35] have also occasionally been considered as metalloids.

Criteria-based

Element IE
(kcal/mol)
IE
(kJ/mol)
EN Band structure
Boron 191 801 2.04 semiconductor
Silicon 188 787 1.90 semiconductor
Germanium 182 762 2.01 semiconductor
Arsenic 226 944 2.18 semimetal
Antimony 199 831 2.05 semimetal
Tellurium 208 869 2.10 semiconductor
average 199 832 2.05
The elements commonly recognised as metalloids, and their ionization energies (IE);[36] electronegativities (EN, revised Pauling scale); and electronic band structures[37] (most thermodynamically stable forms under ambient conditions).

No widely accepted definition of a metalloid exists, nor any division of the periodic table into metals, metalloids, and nonmetals;[38] Hawkes[39] questioned the feasibility of establishing a specific definition, noting that anomalies can be found in several attempted constructs. Classifying an element as a metalloid has been described by Sharp[40] as "arbitrary".

The number and identities of metalloids depend on what classification criteria are used. Emsley[41] recognised four metalloids (germanium, arsenic, antimony, and tellurium); James et al.[42] listed twelve (Emsley's plus boron, carbon, silicon, selenium, bismuth, polonium, moscovium, and livermorium). On average, seven elements are included in such lists; individual classification arrangements tend to share common ground and vary in the ill-defined[43] margins.[n 5][n 6]

A single quantitative criterion such as electronegativity is commonly used,[46] metalloids having electronegativity values from 1.8 or 1.9 to 2.2.[47] Further examples include packing efficiency (the fraction of volume in a crystal structure occupied by atoms) and the Goldhammer–Herzfeld criterion ratio.[48] The commonly recognised metalloids have packing efficiencies of between 34% and 41%.[n 7] The Goldhammer–Herzfeld ratio, roughly equal to the cube of the atomic radius divided by the molar volume,[56][n 8] is a simple measure of how metallic an element is, the recognised metalloids having ratios from around 0.85 to 1.1 and averaging 1.0.[58][n 9] Other authors have relied on, for example, atomic conductance[n 10][62] or bulk coordination number.[63]

Jones, writing on the role of classification in science, observed that "[classes] are usually defined by more than two attributes".[64] Masterton and Slowinski[65] used three criteria to describe the six elements commonly recognised as metalloids: metalloids have ionization energies around 200 kcal/mol (837 kJ/mol) and electronegativity values close to 2.0. They also said that metalloids are typically semiconductors, though antimony and arsenic (semimetals from a physics perspective) have electrical conductivities approaching those of metals. Selenium and polonium are suspected as not in this scheme, while astatine's status is uncertain.[n 11]

In this context, Vernon proposed that a metalloid is a chemical element that, in its standard state, has (a) the electronic band structure of a semiconductor or a semimetal; and (b) an intermediate first ionization potential "(say 750−1,000 kJ/mol)"; and (c) an intermediate electronegativity (1.9–2.2).[68]

Periodic table territory

Distribution and recognition status
of elements classified as metalloids
1 2 12 13 14 15 16 17 18
H     He
Li Be B C N O F Ne
Na Mg Al Si P S Cl Ar
K Ca Zn Ga Ge As Se Br Kr
Rb Sr Cd In Sn Sb Te I Xe
Cs Ba Hg Tl Pb Bi Po At Rn
Fr Ra Cn Nh Fl Mc Lv Ts Og
 
  Commonly (93%) to rarely (9%) recognised as a
metalloid: B, C, Al, Si, Ge, As, Se, Sb, Te, Po, At
  Very rarely (1–5%): H, Be, P, S, Ga, Sn, I, Pb, Bi, Fl, Mc, Lv, Ts
  Sporadically: N, Zn, Rn
  Metal–nonmetal dividing line: between H and Li, Be and B, Al and Si, Ge and As, Sb and Te, Po and At, and Ts and Og

Periodic table extract showing groups 1–2 and 12–18, and a dividing line between metals and nonmetals. Percentages are median appearance frequencies in the

list of metalloid lists
. Sporadically recognised elements show that the metalloid net is sometimes cast very widely; although they do not appear in the list of metalloid lists, isolated references to their designation as metalloids can be found in the literature (as cited in this article).

Location

Metalloids lie on either side of the dividing line between metals and nonmetals. This can be found, in varying configurations, on some periodic tables. Elements to the lower left of the line generally display increasing metallic behaviour; elements to the upper right display increasing nonmetallic behaviour.[69] When presented as a regular stairstep, elements with the highest critical temperature for their groups (Li, Be, Al, Ge, Sb, Po) lie just below the line.[70]

The diagonal positioning of the metalloids represents an exception to the observation that elements with similar properties tend to occur in vertical

d-block
series.

This exception arises due to competing horizontal and vertical trends in the

nuclear charge. Going along a period, the nuclear charge increases with atomic number as do the number of electrons. The additional pull on outer electrons as nuclear charge increases generally outweighs the screening effect of having more electrons. With some irregularities, atoms therefore become smaller, ionization energy increases, and there is a gradual change in character, across a period, from strongly metallic, to weakly metallic, to weakly nonmetallic, to strongly nonmetallic elements.[73] Going down a main group, the effect of increasing nuclear charge is generally outweighed by the effect of additional electrons being further away from the nucleus. Atoms generally become larger, ionization energy falls, and metallic character increases.[74] The net effect is that the location of the metal–nonmetal transition zone shifts to the right in going down a group,[71] and analogous diagonal similarities are seen elsewhere in the periodic table, as noted.[75]

Alternative treatments

Elements bordering the metal–nonmetal dividing line are not always classified as metalloids, noting a binary classification can facilitate the establishment of rules for determining bond types between metals and nonmetals.[76] In such cases, the authors concerned focus on one or more attributes of interest to make their classification decisions, rather than being concerned about the marginal nature of the elements in question. Their considerations may or not be made explicit and may, at times, seem arbitrary.[40][n 12] Metalloids may be grouped with metals;[77] or regarded as nonmetals;[78] or treated as a sub-category of nonmetals.[79][n 13] Other authors have suggested classifying some elements as metalloids "emphasizes that properties change gradually rather than abruptly as one moves across or down the periodic table".[81] Some periodic tables distinguish elements that are metalloids and display no formal dividing line between metals and nonmetals. Metalloids are instead shown as occurring in a diagonal band[82] or diffuse region.[83] The key consideration is to explain the context for the taxonomy in use.

Properties

Metalloids usually look like metals but behave largely like nonmetals. Physically, they are shiny, brittle solids with intermediate to relatively good electrical conductivity and the electronic band structure of a semimetal or semiconductor. Chemically, they mostly behave as (weak) nonmetals, have intermediate ionization energies and electronegativity values, and amphoteric or weakly acidic oxides. They can form alloys with metals. Most of their other physical and chemical properties are intermediate in nature.

Compared to metals and nonmetals

Characteristic properties of metals, metalloids, and nonmetals are summarized in the table.[84] Physical properties are listed in order of ease of determination; chemical properties run from general to specific, and then to descriptive.

Properties of metals, metalloids and nonmetals
Physical property Metals Metalloids Nonmetals
Form solid; a few liquid at or near room temperature (Ga, Hg, Rb, Cs, Fr)[85][n 14] solid[87] majority gaseous[88]
Appearance lustrous (at least when freshly fractured) lustrous[87] several colourless; others coloured, or metallic grey to black
Elasticity typically elastic, ductile, malleable (when solid) brittle[89] brittle, if solid
Electrical conductivity
good to high[n 15] intermediate[91] to good[n 16] poor to good[n 17]
Band structure
metallic (Bi = semimetallic) are semiconductors or, if not (As, Sb = semimetallic), exist in semiconducting forms[95] semiconductor or insulator[96]
Chemical property Metals Metalloids Nonmetals
General chemical behaviour metallic nonmetallic[97] nonmetallic
Ionization energy relatively low intermediate ionization energies,[98] usually falling between those of metals and nonmetals[99] relatively high
Electronegativity usually low have electronegativity values close to 2[100] (revised Pauling scale) or within the range of 1.9–2.2 (Allen scale)[16][n 18] high
When mixed
with metals
give alloys can form alloys[103] ionic or
interstitial compounds
formed
Oxides lower oxides basic; higher oxides increasingly acidic amphoteric or weakly acidic[104] acidic

The above table reflects the hybrid nature of metalloids. The properties of form, appearance, and behaviour when mixed with metals are more like metals. Elasticity and general chemical behaviour are more like nonmetals. Electrical conductivity, band structure, ionization energy, electronegativity, and oxides are intermediate between the two.

Common applications

The focus of this section is on the recognised metalloids. Elements less often recognised as metalloids are ordinarily classified as either metals or nonmetals; some of these are included here for comparative purposes.

Metalloids are too brittle to have any structural uses in their pure forms.[105] They and their compounds are used as (or in) alloying components, biological agents (toxicological, nutritional, and medicinal), catalysts, flame retardants, glasses (oxide and metallic), optical storage media and optoelectronics, pyrotechnics, semiconductors, and electronics.[n 19]

Alloys

Several dozen metallic pellets, reddish-brown. They have a highly polished appearance, as if they had a cellophane coating.
Copper-germanium alloy pellets, likely ~84% Cu; 16% Ge.[107] When combined with silver the result is a tarnish resistant sterling silver. Also shown are two silver pellets.

Writing early in the history of

B metals
, "are probably best classed as alloys".

Among the lighter metalloids, alloys with

case hardening compositions for the engineering industry. Alloys of silicon with iron and with aluminium are widely used by the steel and automotive industries, respectively. Germanium forms many alloys, most importantly with the coinage metals.[111]

The heavier metalloids continue the theme. Arsenic can form alloys with metals, including

copper tellurium (40–50% tellurium).[116] Ferrotellurium is used as a stabilizer for carbon in steel casting.[117] Of the non-metallic elements less often recognised as metalloids, selenium – in the form of ferroselenium (50–58% selenium) – is used to improve the machinability of stainless steels.[118]

Biological agents

antileukaemic properties of white arsenic were first reported in 1878.[119]

All six of the elements commonly recognised as metalloids have toxic, dietary or medicinal properties.[120]

Arsenic and antimony compounds are especially toxic; boron, silicon, and possibly arsenic, are essential trace elements. Boron, silicon, arsenic, and antimony have medical applications, and germanium and tellurium are thought to have potential.

Boron is used in insecticides[121] and herbicides.[122] It is an essential trace element.[123] As boric acid, it has antiseptic, antifungal, and antiviral properties.[124]

Silicon is present in

silatrane, a highly toxic rodenticide.[125] Long-term inhalation of silica dust causes silicosis, a fatal disease of the lungs. Silicon is an essential trace element.[123] Silicone gel can be applied to badly burned patients to reduce scarring.[126]

Salts of germanium are potentially harmful to humans and animals if ingested on a prolonged basis.[127] There is interest in the pharmacological actions of germanium compounds but no licensed medicine as yet.[128]

Arsenic is notoriously poisonous and may also be an

acute promyelocytic leukaemia, a cancer of the blood and bone marrow.[131] Arsenic in drinking water, which causes lung and bladder cancer, has been associated with a reduction in breast cancer mortality rates.[132]

Metallic antimony is relatively non-toxic, but most antimony compounds are poisonous.[133] Two antimony compounds, sodium stibogluconate and stibophen, are used as antiparasitical drugs.[134]

Elemental tellurium is not considered particularly toxic; two grams of sodium tellurate, if administered, can be lethal.[135] People exposed to small amounts of airborne tellurium exude a foul and persistent garlic-like odour.[136] Tellurium dioxide has been used to treat seborrhoeic dermatitis; other tellurium compounds were used as antimicrobial agents before the development of antibiotics.[137] In the future, such compounds may need to be substituted for antibiotics that have become ineffective due to bacterial resistance.[138]

Of the elements less often recognised as metalloids, beryllium and lead are noted for their toxicity;

antibacterials.[144]

Catalysts

alumina have been used as catalysts for the removal of sulfur contaminants from natural gas.[155] Titanium doped aluminium has been identified as a substitute for expensive noble metal catalysts used in the production of industrial chemicals.[156]

Flame retardants

Compounds of boron, silicon, arsenic, and antimony have been used as

silica, and silicates, some of which were developed as alternatives to more toxic halogenated products, can considerably improve the flame retardancy of plastic materials.[158]
Arsenic compounds such as

Glass formation

Optical fibers, usually made of pure silicon dioxide glass, with additives such as boron trioxide or germanium dioxide
for increased sensitivity

The oxides

glass fibre additive,[166] and is also a component of borosilicate glass, widely used for laboratory glassware and domestic ovenware for its low thermal expansion.[167] Most ordinary glassware is made from silicon dioxide.[168] Germanium dioxide is used as a glass fibre additive, as well as in infrared optical systems.[169] Arsenic trioxide is used in the glass industry as a decolourizing and fining agent (for the removal of bubbles),[170] as is antimony trioxide.[171] Tellurium dioxide finds application in laser and nonlinear optics.[172]

metallic glasses are generally most easily prepared if one of the components is a metalloid or "near metalloid" such as boron, carbon, silicon, phosphorus or germanium.[173][n 20] Aside from thin films deposited at very low temperatures, the first known metallic glass was an alloy of composition Au75Si25 reported in 1960.[175] A metallic glass having a strength and toughness not previously seen, of composition Pd82.5P6Si9.5Ge2, was reported in 2011.[176]

Phosphorus, selenium, and lead, which are less often recognised as metalloids, are also used in glasses.

sodium lamps.[177] Selenium compounds can be used both as decolourising agents and to add a red colour to glass.[178] Decorative glassware made of traditional lead glass contains at least 30% lead(II) oxide (PbO); lead glass used for radiation shielding may have up to 65% PbO.[179] Lead-based glasses have also been extensively used in electronic components, enamelling, sealing and glazing materials, and solar cells. Bismuth based oxide glasses have emerged as a less toxic replacement for lead in many of these applications.[180]

Optical storage and optoelectronics

Varying compositions of

crystalline states. The change in optical and electrical properties can be used for information storage purposes.[181] Future applications for GeSbTe may include, "ultrafast, entirely solid-state displays with nanometre-scale pixels, semi-transparent 'smart' glasses, 'smart' contact lenses, and artificial retina devices."[182]

Pyrotechnics

A man is standing in the dark. He is holding out a short stick at mid-chest level. The end of the stick is alight, burning very brightly, and emitting smoke.
Archaic blue light signal, fuelled by a mixture of sodium nitrate, sulfur, and (red) arsenic trisulfide[183]

The recognised metalloids have either pyrotechnic applications or associated properties. Boron and silicon are commonly encountered;

blasting cap initiator compositions.[192]

Carbon, aluminium, phosphorus, and selenium continue the theme. Carbon, in

black powder, is a constituent of fireworks rocket propellants, bursting charges, and effects mixtures, and military delay fuses and igniters.[193][n 22] Aluminium is a common pyrotechnic ingredient,[184] and is widely employed for its capacity to generate light and heat,[195] including in thermite mixtures.[196] Phosphorus can be found in smoke and incendiary munitions, paper caps used in toy guns, and party poppers.[197] Selenium has been used in the same way as tellurium.[192]

Semiconductors and electronics

GaAs, for example) or as doping agents (B, Sb, Te
, for example).

All the elements commonly recognised as metalloids (or their compounds) have been used in the semiconductor or solid-state electronic industries.[198]

Some properties of boron have limited its use as a semiconductor. It has a high melting point, single crystals are relatively hard to obtain, and introducing and retaining controlled impurities is difficult.[199]

Silicon is the leading commercial semiconductor; it forms the basis of modern electronics (including standard solar cells)[200] and information and communication technologies.[201] This was despite the study of semiconductors, early in the 20th century, having been regarded as the "physics of dirt" and not deserving of close attention.[202]

Germanium has largely been replaced by silicon in semiconducting devices, being cheaper, more resilient at higher operating temperatures, and easier to work during the microelectronic fabrication process.

silicon-germanium "alloys" and these have been growing in use, particularly for wireless communication devices; such alloys exploit the higher carrier mobility of germanium.[107] The synthesis of gram-scale quantities of semiconducting germanane was reported in 2013. This consists of one-atom thick sheets of hydrogen-terminated germanium atoms, analogous to graphane. It conducts electrons more than ten times faster than silicon and five times faster than germanium, and is thought to have potential for optoelectronic and sensing applications.[203] The development of a germanium-wire based anode that more than doubles the capacity of lithium-ion batteries was reported in 2014.[204] In the same year, Lee et al. reported that defect-free crystals of graphene large enough to have electronic uses could be grown on, and removed from, a germanium substrate.[205]

Arsenic and antimony are not semiconductors in their

lithium-ion batteries to be replaced by more powerful sodium ion batteries.[207]

Tellurium, which is a semiconductor in its standard state, is used mainly as a component in type II/VI semiconducting-chalcogenides; these have applications in electro-optics and electronics.[208] Cadmium telluride (CdTe) is used in solar modules for its high conversion efficiency, low manufacturing costs, and large band gap of 1.44 eV, letting it absorb a wide range of wavelengths.[200] Bismuth telluride (Bi2Te3), alloyed with selenium and antimony, is a component of thermoelectric devices used for refrigeration or portable power generation.[209]

Five metalloids – boron, silicon, germanium, arsenic, and antimony – can be found in cell phones (along with at least 39 other metals and nonmetals).

n-type semiconductors.[212] The commercial use of gallium compounds is dominated by semiconductor applications – in integrated circuits, cell phones, laser diodes, light-emitting diodes, photodetectors, and solar cells.[213] Selenium is used in the production of solar cells[214] and in high-energy surge protectors.[215]

Boron, silicon, germanium, antimony, and tellurium,

Nomenclature and history

Derivation and other names

The word metalloid comes from the

chromium dioxide) or alloy that can act as a conductor and an insulator. "Meta-metal" is sometimes used instead to refer to certain metals (Be, Zn, Cd, Hg, In, Tl, β-Sn, Pb) located just to the left of the metalloids on standard periodic tables.[225] These metals are mostly diamagnetic[233] and tend to have distorted crystalline structures, electrical conductivity values at the lower end of those of metals, and amphoteric (weakly basic) oxides.[234] "Semimetal" sometimes refers, loosely or explicitly, to metals with incomplete metallic character in crystalline structure, electrical conductivity or electronic structure. Examples include gallium,[235] ytterbium,[236] bismuth[237] and neptunium.[238] The names amphoteric element and semiconductor are problematic as some elements referred to as metalloids do not show marked amphoteric behaviour (bismuth, for example)[239] or semiconductivity (polonium)[240]
in their most stable forms.

Origin and usage

The origin and usage of the term metalloid is convoluted. Its origin lies in attempts, dating from antiquity, to describe metals and to distinguish between typical and less typical forms. It was first applied in the early 19th century to metals that floated on water (sodium and potassium), and then more popularly to nonmetals. Earlier usage in mineralogy, to describe a mineral having a metallic appearance, can be sourced to as early as 1800.[241] Since the mid-20th century it has been used to refer to intermediate or borderline chemical elements.[242][n 23] The International Union of Pure and Applied Chemistry (IUPAC) previously recommended abandoning the term metalloid, and suggested using the term semimetal instead.[244] Use of this latter term has more recently been discouraged by Atkins et al.[2] as it has a different meaning in physics – one that more specifically refers to the electronic band structure of a substance rather than the overall classification of an element. The most recent IUPAC publications on nomenclature and terminology do not include any recommendations on the usage of the terms metalloid or semimetal.[245]

Elements commonly recognised as metalloids

Properties noted in this section refer to the elements in their most thermodynamically stable forms under ambient conditions.

Boron

allotrope)[246]

Pure boron is a shiny, silver-grey crystalline solid.[247] It is less dense than aluminium (2.34 vs. 2.70 g/cm3), and is hard and brittle. It is barely reactive under normal conditions, except for attack by fluorine,[248] and has a melting point of 2076 °C (cf. steel ~1370 °C).[249] Boron is a semiconductor;[250] its room temperature electrical conductivity is 1.5 × 10−6 S•cm−1[251] (about 200 times less than that of tap water)[252] and it has a band gap of about 1.56 eV.[253][n 24] Mendeleev commented that, "Boron appears in a free state in several forms which are intermediate between the metals and the nonmmetals."[255]

The structural chemistry of boron is dominated by its small atomic size, and relatively high ionization energy. With only three valence electrons per boron atom, simple covalent bonding cannot fulfil the octet rule.[256] Metallic bonding is the usual result among the heavier congenors of boron but this generally requires low ionization energies.[257] Instead, because of its small size and high ionization energies, the basic structural unit of boron (and nearly all of its allotropes)[n 25] is the icosahedral B12 cluster. Of the 36 electrons associated with 12 boron atoms, 26 reside in 13 delocalized molecular orbitals; the other 10 electrons are used to form two- and three-centre covalent bonds between icosahedra.[259] The same motif can be seen, as are deltahedral variants or fragments, in metal borides and hydride derivatives, and in some halides.[260]

The bonding in boron has been described as being characteristic of behaviour intermediate between metals and nonmetallic

covalent network solids (such as diamond).[261] The energy required to transform B, C, N, Si, and P from nonmetallic to metallic states has been estimated as 30, 100, 240, 33, and 50 kJ/mol, respectively. This indicates the proximity of boron to the metal-nonmetal borderline.[262]

Most of the chemistry of boron is nonmetallic in nature.

polymeric in structure,[274] weakly acidic,[275][n 29] and a glass former.[281] Organometallic compounds of boron[n 30] have been known since the 19th century (see organoboron chemistry).[283]

Silicon

A lustrous blue grey potato-shaped lump with an irregular corrugated surface.
Silicon has a blue-grey metallic lustre.

Silicon is a crystalline solid with a blue-grey metallic lustre.[284] Like boron, it is less dense (at 2.33 g/cm3) than aluminium, and is hard and brittle.[285] It is a relatively unreactive element.[284] According to Rochow,[286] the massive crystalline form (especially if pure) is "remarkably inert to all acids, including hydrofluoric".[n 31] Less pure silicon, and the powdered form, are variously susceptible to attack by strong or heated acids, as well as by steam and fluorine.[290] Silicon dissolves in hot aqueous alkalis with the evolution of hydrogen, as do metals[291] such as beryllium, aluminium, zinc, gallium or indium.[292] It melts at 1414 °C. Silicon is a semiconductor with an electrical conductivity of 10−4 S•cm−1[293] and a band gap of about 1.11 eV.[287] When it melts, silicon becomes a reasonable metal[294] with an electrical conductivity of 1.0–1.3 × 104 S•cm−1, similar to that of liquid mercury.[295]

The chemistry of silicon is generally nonmetallic (covalent) in nature.

organosilicon).[308]

Germanium

Greyish lustrous block with uneven cleaved surface.
Germanium is sometimes described as a metal

Germanium is a shiny grey-white solid.[309] It has a density of 5.323 g/cm3 and is hard and brittle.[310] It is mostly unreactive at room temperature[n 34] but is slowly attacked by hot concentrated sulfuric or nitric acid.[312] Germanium also reacts with molten caustic soda to yield sodium germanate Na2GeO3 and hydrogen gas.[313] It melts at 938 °C. Germanium is a semiconductor with an electrical conductivity of around 2 × 10−2 S•cm−1[312] and a band gap of 0.67 eV.[314] Liquid germanium is a metallic conductor, with an electrical conductivity similar to that of liquid mercury.[315]

Most of the chemistry of germanium is characteristic of a nonmetal.

oxoacid salts. A phosphate [(HPO4)2Ge·H2O] and highly stable trifluoroacetate Ge(OCOCF3)4 have been described, as have Ge2(SO4)2, Ge(ClO4)4 and GeH2(C2O4)3.[332] The oxide GeO2 is polymeric,[274] amphoteric,[333] and a glass former.[281] The dioxide is soluble in acidic solutions (the monoxide GeO, is even more so), and this is sometimes used to classify germanium as a metal.[334] Up to the 1930s germanium was considered to be a poorly conducting metal;[335] it has occasionally been classified as a metal by later writers.[336] As with all the elements commonly recognised as metalloids, germanium has an established organometallic chemistry (see Organogermanium chemistry).[337]

Arsenic

tarnishing

Arsenic is a grey, metallic looking solid. It has a density of 5.727 g/cm3 and is brittle, and moderately hard (more than aluminium; less than

atm, at 817 °C.[341] It is a semimetal with an electrical conductivity of around 3.9 × 104 S•cm−1[342] and a band overlap of 0.5 eV.[343][n 36] Liquid arsenic is a semiconductor with a band gap of 0.15 eV.[345]

The chemistry of arsenic is predominately nonmetallic.

hydrohalic acid.[357] The oxide is acidic but, as noted below, (weakly) amphoteric. The higher, less stable, pentavalent state has strongly acidic (nonmetallic) properties.[358] Compared to phosphorus, the stronger metallic character of arsenic is indicated by the formation of oxoacid salts such as AsPO4, As2(SO4)3[n 38] and arsenic acetate As(CH3COO)3.[361] The oxide As2O3 is polymeric,[274] amphoteric,[362][n 39] and a glass former.[281] Arsenic has an extensive organometallic chemistry (see Organoarsenic chemistry).[365]

Antimony

A glistening silver rock-like chunk, with a blue tint, and roughly parallel furrows.
Antimony, showing its brilliant lustre

Antimony is a silver-white solid with a blue tint and a brilliant lustre.

sulfate Sb2(SO4)3.[366] It is not affected by molten alkali.[367] Antimony is capable of displacing hydrogen from water, when heated: 2 Sb + 3 H2O → Sb2O3 + 3 H2.[368] It melts at 631 °C. Antimony is a semimetal with an electrical conductivity of around 3.1 × 104 S•cm−1[369] and a band overlap of 0.16 eV.[343][n 40] Liquid antimony is a metallic conductor with an electrical conductivity of around 5.3 × 104 S•cm−1.[371]

Most of the chemistry of antimony is characteristic of a nonmetal.

pentapositive antimony is (predominately) acidic.[380] Consistent with an increase in metallic character down group 15, antimony forms salts including an acetate Sb(CH3CO2)3, phosphate SbPO4, sulfate Sb2(SO4)3 and perchlorate Sb(ClO4)3.[381] The otherwise acidic pentoxide Sb2O5 shows some basic (metallic) behaviour in that it can be dissolved in very acidic solutions, with the formation of the oxycation SbO+
2
.[382] The oxide Sb2O3 is polymeric,[274] amphoteric,[383] and a glass former.[281] Antimony has an extensive organometallic chemistry (see Organoantimony chemistry).[384]

Tellurium

nonmetals[385]

Tellurium is a silvery-white shiny solid.[386] It has a density of 6.24 g/cm3, is brittle, and is the softest of the commonly recognised metalloids, being marginally harder than sulfur.[338] Large pieces of tellurium are stable in air. The finely powdered form is oxidized by air in the presence of moisture. Tellurium reacts with boiling water, or when freshly precipitated even at 50 °C, to give the dioxide and hydrogen: Te + 2 H2O → TeO2 + 2 H2.[387] It reacts (to varying degrees) with nitric, sulfuric, and hydrochloric acids to give compounds such as the sulfoxide TeSO3 or tellurous acid H2TeO3,[388] the basic nitrate (Te2O4H)+(NO3),[389] or the oxide sulfate Te2O3(SO4).[390] It dissolves in boiling alkalis, to give the tellurite and telluride: 3 Te + 6 KOH = K2TeO3 + 2 K2Te + 3 H2O, a reaction that proceeds or is reversible with increasing or decreasing temperature.[391]

At higher temperatures tellurium is sufficiently plastic to extrude.

Superheated liquid tellurium is a metallic conductor.[396]

Most of the chemistry of tellurium is characteristic of a nonmetal.[397] It shows some cationic behaviour. The dioxide dissolves in acid to yield the trihydroxotellurium(IV) Te(OH)3+ ion;

basic selenate 2TeO2·SeO3 and an analogous perchlorate and periodate 2TeO2·HXO4.[404] Tellurium forms a polymeric,[274] amphoteric,[383] glass-forming oxide[281] TeO2. It is a "conditional" glass-forming oxide – it forms a glass with a very small amount of additive.[281] Tellurium has an extensive organometallic chemistry (see Organotellurium chemistry).[405]

Elements less commonly recognised as metalloids

Carbon

A shiny grey-black cuboid nugget with a rough surface.
Carbon (as graphite). Delocalized valence electrons within the layers of graphite give it a metallic appearance.[406]

Carbon is ordinarily classified as a nonmetal

allotrope of carbon under ambient conditions.[409] It has a lustrous appearance[410] and is a fairly good electrical conductor.[411] Graphite has a layered structure. Each layer consists of carbon atoms bonded to three other carbon atoms in a hexagonal lattice arrangement. The layers are stacked together and held loosely by van der Waals forces and delocalized valence electrons.[412]

Like a metal, the conductivity of graphite in the direction of its planes decreases as the temperature is raised;[413][n 43] it has the electronic band structure of a semimetal.[413] The allotropes of carbon, including graphite, can accept foreign atoms or compounds into their structures via substitution, intercalation, or doping. The resulting materials are referred to as "carbon alloys".[417] Carbon can form ionic salts, including a hydrogen sulfate, perchlorate, and nitrate (C+
24
X.2HX, where X = HSO4, ClO4; and C+
24
NO
3
.3HNO3).[418][n 44] In organic chemistry, carbon can form complex cations – termed carbocations – in which the positive charge is on the carbon atom; examples are CH+
3
and CH+
5
, and their derivatives.[419]

Carbon is brittle,

sesquicarbide or allylenide), in compounds with metals of main groups 1–3, and with the lanthanides and actinides.[424] Its oxide CO2 forms carbonic acid H2CO3.[425][n 45]

Aluminium

alloys. People who handle it for the first time often ask if it is the real thing.[427]

Aluminium is ordinarily classified as a metal.

close-packed crystalline structure,[429] and forms a cation in aqueous solution.[430]

It has some properties that are unusual for a metal; taken together,[431] these are sometimes used as a basis to classify aluminium as a metalloid.[432] Its crystalline structure shows some evidence of directional bonding.[433] Aluminium bonds covalently in most compounds.[434] The oxide Al2O3 is amphoteric[435] and a conditional glass-former.[281] Aluminium can form anionic aluminates,[431] such behaviour being considered nonmetallic in character.[69]

Classifying aluminium as a metalloid has been disputed[436] given its many metallic properties. It is therefore, arguably, an exception to the mnemonic that elements adjacent to the metal–nonmetal dividing line are metalloids.[437][n 46]

Stott

high negative electrode potential". Moody[441]
says that, "aluminium is on the 'diagonal borderland' between metals and non-metals in the chemical sense."

Selenium

photoconductor, conducts electricity around 1,000 times better when light falls on it, a property used since the mid-1870s in various light-sensing applications[442]

Selenium shows borderline metalloid or nonmetal behaviour.[443][n 47]

Its most stable form, the grey

trigonal allotrope, is sometimes called "metallic" selenium because its electrical conductivity is several orders of magnitude greater than that of the red monoclinic form.[446] The metallic character of selenium is further shown by its lustre,[447] and its crystalline structure, which is thought to include weakly "metallic" interchain bonding.[448] Selenium can be drawn into thin threads when molten and viscous.[449] It shows reluctance to acquire "the high positive oxidation numbers characteristic of nonmetals".[450] It can form cyclic polycations (such as Se2+
8
) when dissolved in oleums[451] (an attribute it shares with sulfur and tellurium), and a hydrolysed cationic salt in the form of trihydroxoselenium(IV) perchlorate [Se(OH)3]+·ClO
4
.[452]

The nonmetallic character of selenium is shown by its brittleness[447] and the low electrical conductivity (~10−9 to 10−12 S•cm−1) of its highly purified form.[93] This is comparable to or less than that of bromine (7.95×10–12 S•cm−1),[453] a nonmetal. Selenium has the electronic band structure of a semiconductor[454] and retains its semiconducting properties in liquid form.[454] It has a relatively high[455] electronegativity (2.55 revised Pauling scale). Its reaction chemistry is mainly that of its nonmetallic anionic forms Se2−, SeO2−
3
and SeO2−
4
.[456]

Selenium is commonly described as a metalloid in the environmental chemistry literature.[457] It moves through the aquatic environment similarly to arsenic and antimony;[458] its water-soluble salts, in higher concentrations, have a similar toxicological profile to that of arsenic.[459]

Polonium

Polonium is "distinctly metallic" in some ways.[240] Both of its allotropic forms are metallic conductors.[240] It is soluble in acids, forming the rose-coloured Po2+ cation and displacing hydrogen: Po + 2 H+ → Po2+ + H2.[460] Many polonium salts are known.[461] The oxide PoO2 is predominantly basic in nature.[462] Polonium is a reluctant oxidizing agent, unlike its lightest congener oxygen: highly reducing conditions are required for the formation of the Po2− anion in aqueous solution.[463]

Whether polonium is ductile or brittle is unclear. It is predicted to be ductile based on its calculated elastic constants.[464] It has a simple cubic crystalline structure. Such a structure has few slip systems and "leads to very low ductility and hence low fracture resistance".[465]

Polonium shows nonmetallic character in its halides, and by the existence of

organic solvents).[466] Many metal polonides, obtained by heating the elements together at 500–1,000 °C, and containing the Po2− anion, are also known.[467]

Astatine

As a

face-centred cubic crystalline structure.[475]

Several authors have commented on the metallic nature of some of the properties of astatine. Since iodine is a semiconductor in the direction of its planes, and since the halogens become more metallic with increasing atomic number, it has been presumed that astatine would be a metal if it could form a condensed phase.[476][n 49] Astatine may be metallic in the liquid state on the basis that elements with an enthalpy of vaporization (∆Hvap) greater than ~42 kJ/mol are metallic when liquid.[478] Such elements include boron,[n 50] silicon, germanium, antimony, selenium, and tellurium. Estimated values for ∆Hvap of diatomic astatine are 50 kJ/mol or higher;[482] diatomic iodine, with a ∆Hvap of 41.71,[483] falls just short of the threshold figure.

"Like typical metals, it [astatine] is precipitated by

pseudohalide compounds ... complexes of astatine cations ... complex anions of trivalent astatine ... as well as complexes with a variety of organic solvents".[486] It has also been argued that astatine demonstrates cationic behaviour, by way of stable At+ and AtO+ forms, in strongly acidic aqueous solutions.[487]

Some of astatine's reported properties are nonmetallic. It has been extrapolated to have the narrow liquid range ordinarily associated with nonmetals (mp 302 °C; bp 337 °C),

Restrepo et al.[495] reported that astatine appeared to be more polonium-like than halogen-like. They did so on the basis of detailed comparative studies of the known and interpolated properties of 72 elements.

Related concepts

Near metalloids

electrical conductivity of 1.7 × 10−8 S•cm−1 at room temperature.[496] This is higher than selenium but lower than boron, the least electrically conducting of the recognised metalloids.[n 52]

In the periodic table, some of the elements adjacent to the commonly recognised metalloids, although usually classified as either metals or nonmetals, are occasionally referred to as near-metalloids[499] or noted for their metalloidal character. To the left of the metal–nonmetal dividing line, such elements include gallium,[500] tin[501] and bismuth.[502] They show unusual packing structures,[503] marked covalent chemistry (molecular or polymeric),[504] and amphoterism.[505] To the right of the dividing line are carbon,[506] phosphorus,[507] selenium[508] and iodine.[509] They exhibit metallic lustre, semiconducting properties[n 53] and bonding or valence bands with delocalized character. This applies to their most thermodynamically stable forms under ambient conditions: carbon as graphite; phosphorus as black phosphorus;[n 54] and selenium as grey selenium.

Allotropes

grey tin
(right). Both forms have a metallic appearance.

Different crystalline forms of an element are called allotropes. Some allotropes, particularly those of elements located (in periodic table terms) alongside or near the notional dividing line between metals and nonmetals, exhibit more pronounced metallic, metalloidal or nonmetallic behaviour than others.[515] The existence of such allotropes can complicate the classification of the elements involved.[516]

Tin, for example, has two allotropes:

polycrystalline forms. It is the stable form below 13.2 °C and has an electrical conductivity of between (2–5) × 102 S·cm−1 (~1/250th that of white tin).[518] Grey tin has the same crystalline structure as that of diamond. It behaves as a semiconductor (as if it had a band gap of 0.08 eV), but has the electronic band structure of a semimetal.[519] It has been referred to as either a very poor metal,[520] a metalloid,[521] a nonmetal[522] or a near metalloid.[502]

The diamond allotrope of carbon is clearly nonmetallic, being translucent and having a low electrical conductivity of 10−14 to 10−16 S·cm−1.[523] Graphite has an electrical conductivity of 3 × 104 S·cm−1,[524] a figure more characteristic of a metal. Phosphorus, sulfur, arsenic, selenium, antimony, and bismuth also have less stable allotropes that display different behaviours.[525]

Abundance, extraction, and cost

Z Element Grams
/tonne
8 Oxygen 461,000
14 Silicon 282,000
13 Aluminium 82,300
26 Iron 56,300
6 Carbon 200
29 Copper 60
5 Boron 10
33 Arsenic 1.8
32 Germanium 1.5
47 Silver 0.075
34 Selenium 0.05
51 Antimony 0.02
79 Gold 0.004
52 Tellurium 0.001
75 Rhenium 0.00000000077×10−10
54 Xenon 0.000000000033×10−11
84 Polonium 0.00000000000000022×10−16
85 Astatine 0.0000000000000000033×10−20

Abundance

The table gives

crustal abundances of the elements commonly to rarely recognised as metalloids.[526] Some other elements are included for comparison: oxygen and xenon (the most and least abundant elements with stable isotopes); iron and the coinage metals copper, silver, and gold; and rhenium, the least abundant stable metal (aluminium is normally the most abundant metal). Various abundance estimates have been published; these often disagree to some extent.[527]

Extraction

The recognised metalloids can be obtained by

electrolytic reduction: TeO2 + 2 NaOH → Na2TeO3 + H2O;[531] Na2TeO3 + H2O → Te + 2 NaOH + O2.[532] Another option is reduction of the oxide by roasting with carbon: TeO2 + C → Te + CO2.[533]

Production methods for the elements less frequently recognised as metalloids involve natural processing, electrolytic or chemical reduction, or irradiation. Carbon (as graphite) occurs naturally and is extracted by crushing the parent rock and floating the lighter graphite to the surface. Aluminium is extracted by dissolving its oxide Al2O3 in molten

soda ash to give the selenite: X2Se + O2 + Na2CO3 → Na2SeO3 + 2 X + CO2; the selenide is neutralized by sulfuric acid H2SO4 to give selenous acid H2SeO3; this is reduced by bubbling with SO2 to yield elemental selenium. Polonium and astatine are produced in minute quantities by irradiating bismuth.[534]

Cost

The recognised metalloids and their closer neighbours mostly cost less than silver; only polonium and astatine are more expensive than gold, on account of their significant radioactivity. As of 5 April 2014, prices for small samples (up to 100 g) of silicon, antimony and tellurium, and graphite, aluminium and selenium, average around one third the cost of silver (US$1.5 per gram or about $45 an ounce). Boron, germanium, and arsenic samples average about three-and-a-half times the cost of silver.[n 55] Polonium is available for about $100 per microgram.[535] Zalutsky and Pruszynski[536] estimate a similar cost for producing astatine. Prices for the applicable elements traded as commodities tend to range from two to three times cheaper than the sample price (Ge), to nearly three thousand times cheaper (As).[n 56]

Notes

  1. ^ Definitions and extracts by different authors, illustrating aspects of the generic definition, follow:
    • "In chemistry a metalloid is an element with properties intermediate between those of metals and nonmetals."[3]
    • "Between the metals and nonmetals in the periodic table we find elements ... [that] share some of the characteristic properties of both the metals and nonmetals, making it difficult to place them in either of these two main categories"[4]
    • "Chemists sometimes use the name metalloid ... for these elements which are difficult to classify one way or the other."[5]
    • "Because the traits distinguishing metals and nonmetals are qualitative in nature, some elements do not fall unambiguously in either category. These elements ... are called metalloids ..."[6]
    More broadly, metalloids have been referred to as:
    • "elements that ... are somewhat of a cross between metals and nonmetals";[7] or
    • "weird in-between elements".[8]
  2. cation
    formation), gold shows nonmetallic behaviour: On halogen character, see also Belpassi et al.,[12] who conclude that in the aurides MAu (M = Li–Cs) gold "behaves as a halogen, intermediate between Br and I"; on aurophilicity, see also Schmidbaur and Schier.[13]
  3. ^ Mann et al.[16] refer to these elements as "the recognized metalloids".
  4. ^ Jones[44] writes: "Though classification is an essential feature in all branches of science, there are always hard cases at the boundaries. Indeed, the boundary of a class is rarely sharp."
  5. ^ The lack of a standard division of the elements into metals, metalloids, and nonmetals is not necessarily an issue. There is more or less, a continuous progression from the metallic to the nonmetallic. A specified subset of this continuum could serve its particular purpose as well as any other.[45]
  6. ^ The packing efficiency of boron is 38%; silicon and germanium 34; arsenic 38.5; antimony 41; and tellurium 36.4.[49] These values are lower than in most metals (80% of which have a packing efficiency of at least 68%),[50] but higher than those of elements usually classified as nonmetals. (Gallium is unusual, for a metal, in having a packing efficiency of just 39%.)[51] Other notable values for metals are 42.9 for bismuth[52] and 58.5 for liquid mercury.[53]) Packing efficiencies for nonmetals are: graphite 17%,[54] sulfur 19.2,[55] iodine 23.9,[55] selenium 24.2,[55] and black phosphorus 28.5.[52]
  7. ^ More specifically, the Goldhammer–Herzfeld criterion is the ratio of the force holding an individual atom's valence electrons in place with the forces on the same electrons from interactions between the atoms in the solid or liquid element. When the interatomic forces are greater than, or equal to, the atomic force, valence electron itinerancy is indicated and metallic behaviour is predicted.[57] Otherwise nonmetallic behaviour is anticipated.
  8. crystalline structure, on relativistic grounds.[60] Even so it offers a first order rationalization for the occurrence of metallic character amongst the elements.[61]
  9. ^ Atomic conductance is the electrical conductivity of one mole of a substance. It is equal to electrical conductivity divided by molar volume.[5]
  10. ^ Selenium has an ionization energy (IE) of 225 kcal/mol (941 kJ/mol) and is sometimes described as a semiconductor. It has a relatively high 2.55 electronegativity (EN). Polonium has an IE of 194 kcal/mol (812 kJ/mol) and a 2.0 EN, but has a metallic band structure.[66] Astatine has an IE of 215 kJ/mol (899 kJ/mol) and an EN of 2.2.[67] Its electronic band structure is not known with any certainty.
  11. ^ Jones (2010, pp. 169–71): "Though classification is an essential feature of all branches of science, there are always hard cases at the boundaries. The boundary of a class is rarely sharp…Scientists should not lose sleep over the hard cases. As long as a classification system is beneficial to economy of description, to structuring knowledge and to our understanding, and hard cases constitute a small minority, then keep it. If the system becomes less than useful, then scrap it and replace it with a system based on different shared characteristics."
  12. ^ Oderberg[80] argues on ontological grounds that anything not a metal is therefore a nonmetal, and that this includes semi-metals (i.e. metalloids).
  13. ^ Copernicium is reportedly the only metal thought to be a gas at room temperature.[86]
  14. ^ Metals have electrical conductivity values of from 6.9 × 103 S•cm−1 for manganese to 6.3 × 105 for silver.[90]
  15. ^ Metalloids have electrical conductivity values of from 1.5 × 10−6 S•cm−1 for boron to 3.9 × 104 for arsenic.[92] If selenium is included as a metalloid the applicable conductivity range would start from ~10−9 to 10−12 S•cm−1.[93]
  16. ^ Nonmetals have electrical conductivity values of from ~10−18 S•cm−1 for the elemental gases to 3 × 104 in graphite.[94]
  17. Allred-Rochow scale). He included boron, silicon, germanium, arsenic, antimony, tellurium, polonium, and astatine in this category. In reviewing Chedd's work, Adler[102] described this choice as arbitrary, as other elements whose electronegativities lie in this range include copper
    , silver, phosphorus, mercury, and bismuth. He went on to suggest defining a metalloid as "a semiconductor or semimetal" and to include bismuth and selenium in this category.
  18. ^ Olmsted and Williams[106] commented that, "Until quite recently, chemical interest in the metalloids consisted mainly of isolated curiosities, such as the poisonous nature of arsenic and the mildly therapeutic value of borax. With the development of metalloid semiconductors, however, these elements have become among the most intensely studied".
  19. covalent bonding structures coexist.[174]
  20. MoO2. Adding arsenic or antimony (n-type electron donors) increases the rate of reaction; adding gallium or indium (p-type electron acceptors) decreases it.[188]
  21. ^ Ellern, writing in Military and Civilian Pyrotechnics (1968), comments that carbon black "has been specified for and used in a nuclear air-burst simulator."[194]
  22. ^ For a post-1960 example of the former use of the term metalloid to refer to nonmetals see Zhdanov,[243] who divides the elements into metals; intermediate elements (H, B, C, Si, Ge, Se, Te); and metalloids (of which the most typical are given as O, F, and Cl).
  23. ^ Boron, at 1.56 eV, has the largest band gap amongst the commonly recognised (semiconducting) metalloids. Of nearby elements in periodic table terms, selenium has the next highest band gap (close to 1.8 eV) followed by white phosphorus (around 2.1 eV).[254]
  24. ^ The synthesis of B40 borospherene, a "distorted fullerene with a hexagonal hole on the top and bottom and four heptagonal holes around the waist" was announced in 2014.[258]
  25. ^ The BH3 and Fe(CO4) species in these reactions are short-lived reaction intermediates.[266]
  26. ^ On the analogy between boron and metals, Greenwood[268] commented that: "The extent to which metallic elements mimic boron (in having fewer electrons than orbitals available for bonding) has been a fruitful cohering concept in the development of metalloborane chemistry ... Indeed, metals have been referred to as "honorary boron atoms" or even as "flexiboron atoms". The converse of this relationship is clearly also valid ..."
  27. ^ The bonding in boron trifluoride, a gas, has been referred to as predominately ionic[272] a description which was subsequently described as misleading.[273]
  28. P2O5.[280]
  29. ^ Organic derivatives of metalloids are traditionally counted as organometallic compounds.[282]
  30. ^ In air, silicon forms a thin coating of amorphous silicon dioxide, 2 to 3 nm thick.[287] This coating is dissolved by hydrogen fluoride at a very low pace – on the order of two to three hours per nanometre.[288] Silicon dioxide, and silicate glasses (of which silicon dioxide is a major component), are otherwise readily attacked by hydrofluoric acid.[289]
  31. ^ The bonding in silicon tetrafluoride, a gas, has been referred to as predominately ionic[272] a description which was subsequently described as misleading.[273]
  32. ^ Although SiO2 is classified as an acidic oxide, and hence reacts with alkalis to give silicates, it reacts with phosphoric acid to yield a silicon oxide orthophosphate Si5O(PO4)6,[305] and with hydrofluoric acid to give hexafluorosilicic acid H2SiF6.[306] The latter reaction "is sometimes quoted as evidence of basic [that is, metallic] properties".[307]
  33. ^ Temperatures above 400 °C are required to form a noticeable surface oxide layer.[311]
  34. GeO dissolves in dilute acids to give Ge+2 and in dilute bases to produce GeO2−2, all three entities being unstable in water". Sources dismissing germanium cations or further qualifying their presumed existence include: Jolly and Latimer[325] who assert that, "the germanous ion cannot be studied directly because no germanium (II) species exists in any appreciable concentration in noncomplexing aqueous solutions"; Lidin[326] who says that, "[germanium] forms no aquacations"; Ladd[327] who notes that the CdI2 structure is "intermediate in type between ionic and molecular compounds"; and Wiberg[328]
    who states that, "no germanium cations are known".
  35. ^ Arsenic also exists as a naturally occurring (but rare) allotrope (arsenolamprite), a crystalline semiconductor with a band gap of around 0.3 eV or 0.4 eV. It can also be prepared in a semiconducting amorphous form, with a band gap of around 1.2–1.4 eV.[344]
  36. arsenious acid HAsO2; the solubility…increases at pH's below 1 with the formation of 'arsenyl' ions AsO+…"; Kolthoff and Elving[351] who write that, "the As3+ cation exists to some extent only in strongly acid solutions; under less acid conditions the tendency is toward hydrolysis, so that the anionic form predominates"; Moody[352] who observes that, "arsenic trioxide, As4O6, and arsenious acid, H3AsO3, are apparently amphoteric but no cations, As3+, As(OH)2+ or As(OH)2+ are known"; and Cotton et al.[353]
    who write that (in aqueous solution) the simple arsenic cation As3+ "may occur to some slight extent [along with the AsO+ cation]" and that, "Raman spectra show that in acid solutions of As4O6 the only detectable species is the pyramidal As(OH)3".
  37. ^ The formulae of AsPO4 and As2(SO4)3 suggest straightforward ionic formulations, with As3+, but this is not the case. AsPO4, "which is virtually a covalent oxide", has been referred to as a double oxide, of the form As2O3·P2O5. It consists of AsO3 pyramids and PO4 tetrahedra, joined together by all their corner atoms to form a continuous polymeric network.[359] As2(SO4)3 has a structure in which each SO4 tetrahedron is bridged by two AsO3 trigonal pyramida.[360]
  38. ^ As2O3 is usually regarded as being amphoteric but a few sources say it is (weakly)[363] acidic. They describe its "basic" properties (its reaction with concentrated hydrochloric acid to form arsenic trichloride) as being alcoholic, in analogy with the formation of covalent alkyl chlorides by covalent alcohols (e.g., R-OH + HCl RCl + H2O)[364]
  39. ^ Antimony can also be prepared in an amorphous semiconducting black form, with an estimated (temperature-dependent) band gap of 0.06–0.18 eV.[370]
  40. ^ Lidin[375] asserts that SbO+ does not exist and that the stable form of Sb(III) in aqueous solution is an incomplete hydrocomplex [Sb(H2O)4(OH)2]+.
  41. ^ Cotton et al.[399] note that TeO2 appears to have an ionic lattice; Wells[400] suggests that the Te–O bonds have "considerable covalent character".
  42. ^ Liquid carbon may[414] or may not[415] be a metallic conductor, depending on pressure and temperature; see also.[416]
  43. oxidising agent, such as nitric acid, chromium trioxide or ammonium persulfate; in this instance the concentrated sulfuric acid is acting as an inorganic nonaqueous solvent
    .
  44. ^ Only a small fraction of dissolved CO2 is present in water as carbonic acid so, even though H2CO3 is a medium-strong acid, solutions of carbonic acid are only weakly acidic.[426]
  45. ^ A mnemonic that captures the elements commonly recognised as metalloids goes: Up, up-down, up-down, up ... are the metalloids![438]
  46. ^ Rochow,[444] who later wrote his 1966 monograph The metalloids,[445] commented that, "In some respects selenium acts like a metalloid and tellurium certainly does".
  47. ^ A further option is to include astatine both as a nonmetal and as a metalloid.[471]
  48. ^ A visible piece of astatine would be immediately and completely vaporized because of the heat generated by its intense radioactivity.[477]
  49. ^ The literature is contradictory as to whether boron exhibits metallic conductivity in liquid form. Krishnan et al.[479] found that liquid boron behaved like a metal. Glorieux et al.[480] characterised liquid boron as a semiconductor, on the basis of its low electrical conductivity. Millot et al.[481] reported that the emissivity of liquid boron was not consistent with that of a liquid metal.
  50. heavy metals
    ".
  51. ^ The separation between molecules in the layers of iodine (350 pm) is much less than the separation between iodine layers (427 pm; cf. twice the van der Waals radius of 430 pm).[497] This is thought to be caused by electronic interactions between the molecules in each layer of iodine, which in turn give rise to its semiconducting properties and shiny appearance.[498]
  52. ^ For example: intermediate electrical conductivity;[510] a relatively narrow band gap;[511] light sensitivity.[510]
  53. ^ White phosphorus is the least stable and most reactive form.[512] It is also the most common, industrially important,[513] and easily reproducible allotrope, and for these three reasons is regarded as the standard state of the element.[514]
  54. ^ Sample prices of gold, in comparison, start at roughly thirty-five times that of silver. Based on sample prices for B, C, Al, Si, Ge, As, Se, Ag, Sb, Te, and Au available on-line from Alfa Aesa; Goodfellow; Metallium; and United Nuclear Scientific.
  55. .

References

  1. ^ Chedd 1969, pp. 58, 78; National Research Council 1984, p. 43
  2. ^ a b Atkins et al. 2010, p. 20
  3. ^ Cusack 1987, p. 360
  4. ^ Kelter, Mosher & Scott 2009, p. 268
  5. ^ a b Hill & Holman 2000, p. 41
  6. ^ King 1979, p. 13
  7. ^ Moore 2011, p. 81
  8. ^ Gray 2010
  9. ^ Hopkins & Bailar 1956, p. 458
  10. ^ Glinka 1965, p. 77
  11. ^ Wiberg 2001, p. 1279
  12. ^ Belpassi et al. 2006, pp. 4543–44
  13. ^ Schmidbaur & Schier 2008, pp. 1931–51
  14. ^ Tyler Miller 1987, p. 59
  15. ^ Goldsmith 1982, p. 526; Kotz, Treichel & Weaver 2009, p. 62; Bettelheim et al. 2010, p. 46
  16. ^ a b Mann et al. 2000, p. 2783
  17. ^ Hawkes 2001, p. 1686; Segal 1989, p. 965; McMurray & Fay 2009, p. 767
  18. ^ Bucat 1983, p. 26; Brown c. 2007
  19. ^ a b Swift & Schaefer 1962, p. 100
  20. ^ Hawkes 2001, p. 1686; Hawkes 2010; Holt, Rinehart & Wilson c. 2007
  21. ^ Dunstan 1968, pp. 310, 409. Dunstan lists Be, Al, Ge (maybe), As, Se (maybe), Sn, Sb, Te, Pb, Bi, and Po as metalloids (pp. 310, 323, 409, 419).
  22. ^ Tilden 1876, pp. 172, 198–201; Smith 1994, p. 252; Bodner & Pardue 1993, p. 354
  23. ^ Bassett et al. 1966, p. 127
  24. ^ Rausch 1960
  25. ^ Thayer 1977, p. 604; Warren & Geballe 1981; Masters & Ela 2008, p. 190
  26. ^ Warren & Geballe 1981; Chalmers 1959, p. 72; US Bureau of Naval Personnel 1965, p. 26
  27. ^ Siebring 1967, p. 513
  28. ^ Wiberg 2001, p. 282
  29. ^ Rausch 1960; Friend 1953, p. 68
  30. ^ Murray 1928, p. 1295
  31. ^ Hampel & Hawley 1966, p. 950; Stein 1985; Stein 1987, pp. 240, 247–48
  32. ^ Hatcher 1949, p. 223; Secrist & Powers 1966, p. 459
  33. ^ Taylor 1960, p. 614
  34. ^ Considine & Considine 1984, p. 568; Cegielski 1998, p. 147; The American heritage science dictionary 2005, p. 397
  35. ^ Woodward 1948, p. 1
  36. ^ NIST 2010. Values shown in the above table have been converted from the NIST values, which are given in eV.
  37. ^ Berger 1997; Lovett 1977, p. 3
  38. ^ Goldsmith 1982, p. 526; Hawkes 2001, p. 1686
  39. ^ Hawkes 2001, p. 1687
  40. ^ a b Sharp 1981, p. 299
  41. ^ Emsley 1971, p. 1
  42. ^ James et al. 2000, p. 480
  43. ^ Chatt 1951, p. 417 "The boundary between metals and metalloids is indefinite ..."; Burrows et al. 2009, p. 1192: "Although the elements are conveniently described as metals, metalloids, and nonmetals, the transitions are not exact ..."
  44. ^ Jones 2010, p. 170
  45. ^ Kneen, Rogers & Simpson 1972, pp. 218–20
  46. ^ Rochow 1966, pp. 1, 4–7
  47. ^ Rochow 1977, p. 76; Mann et al. 2000, p. 2783
  48. ^ Askeland, Phulé & Wright 2011, p. 69
  49. ^ Van Setten et al. 2007, pp. 2460–61; Russell & Lee 2005, p. 7 (Si, Ge); Pearson 1972, p. 264 (As, Sb, Te; also black P)
  50. ^ Russell & Lee 2005, p. 1
  51. ^ Russell & Lee 2005, pp. 6–7, 387
  52. ^ a b Pearson 1972, p. 264
  53. ^ Okajima & Shomoji 1972, p. 258
  54. ^ Kitaĭgorodskiĭ 1961, p. 108
  55. ^ a b c Neuburger 1936
  56. ^ Edwards & Sienko 1983, p. 693
  57. ^ Herzfeld 1927; Edwards 2000, pp. 100–03
  58. ^ Edwards & Sienko 1983, p. 695; Edwards et al. 2010
  59. ^ Edwards 1999, p. 416
  60. ^ Steurer 2007, p. 142; Pyykkö 2012, p. 56
  61. ^ Edwards & Sienko 1983, p. 695
  62. ^ Hill & Holman 2000, p. 160. They characterise metalloids (in part) on the basis that they are "poor conductors of electricity with atomic conductance usually less than 10−3 but greater than 10−5 ohm−1 cm−4".
  63. ^ Bond 2005, p. 3: "One criterion for distinguishing semi-metals from true metals under normal conditions is that the bulk coordination number of the former is never greater than eight, while for metals it is usually twelve (or more, if for the body-centred cubic structure one counts next-nearest neighbours as well)."
  64. ^ Jones 2010, p. 169
  65. ^ Masterton & Slowinski 1977, p. 160 list B, Si, Ge, As, Sb, and Te as metalloids, and comment that Po and At are ordinarily classified as metalloids but add that this is arbitrary as so little is known about them.
  66. ^ Kraig, Roundy & Cohen 2004, p. 412; Alloul 2010, p. 83
  67. ^ Vernon 2013, p. 1704
  68. ^ Vernon 2013, p. 1703
  69. ^ a b Hamm 1969, p. 653
  70. ^ Horvath 1973, p. 336
  71. ^ a b Gray 2009, p. 9
  72. ^ Rayner-Canham 2011
  73. ^ Booth & Bloom 1972, p. 426; Cox 2004, pp. 17, 18, 27–28; Silberberg 2006, pp. 305–13
  74. ^ Cox 2004, pp. 17–18, 27–28; Silberberg 2006, pp. 305–13
  75. ^ Rodgers 2011, pp. 232–33; 240–41
  76. ^ Roher 2001, pp. 4–6
  77. ^ Tyler 1948, p. 105; Reilly 2002, pp. 5–6
  78. ^ Hampel & Hawley 1976, p. 174;
  79. ^ Goodrich 1844, p. 264; The Chemical News 1897, p. 189; Hampel & Hawley 1976, p. 191; Lewis 1993, p. 835; Hérold 2006, pp. 149–50
  80. ^ Oderberg 2007, p. 97
  81. ^ Brown & Holme 2006, p. 57
  82. ^ Wiberg 2001, p. 282; Simple Memory Art c. 2005
  83. ^ Chedd 1969, pp. 12–13
  84. ^ Kneen, Rogers & Simpson, 1972, p. 263. Columns 2 and 4 are sourced from this reference unless otherwise indicated.
  85. ^ Stoker 2010, p. 62; Chang 2002, p. 304. Chang speculates that the melting point of francium would be about 23 °C.
  86. ^ New Scientist 1975; Soverna 2004; Eichler et al. 2007; Austen 2012
  87. ^ a b Rochow 1966, p. 4
  88. ^ Hunt 2000, p. 256
  89. ^ McQuarrie & Rock 1987, p. 85
  90. ^ Desai, James & Ho 1984, p. 1160; Matula 1979, p. 1260
  91. ^ Choppin & Johnsen 1972, p. 351
  92. ^ Schaefer 1968, p. 76; Carapella 1968, p. 30
  93. ^ a b Kozyrev 1959, p. 104; Chizhikov & Shchastlivyi 1968, p. 25; Glazov, Chizhevskaya & Glagoleva 1969, p. 86
  94. ^ Bogoroditskii & Pasynkov 1967, p. 77; Jenkins & Kawamura 1976, p. 88
  95. ^ Hampel & Hawley 1976, p. 191; Wulfsberg 2000, p. 620
  96. ^ Swalin 1962, p. 216
  97. ^ Bailar et al. 1989, p. 742
  98. ^ Metcalfe, Williams & Castka 1974, p. 86
  99. ^ Chang 2002, p. 306
  100. ^ Pauling 1988, p. 183
  101. ^ Chedd 1969, pp. 24–25
  102. ^ Adler 1969, pp. 18–19
  103. ^ Hultgren 1966, p. 648; Young & Sessine 2000, p. 849; Bassett et al. 1966, p. 602
  104. ^ Rochow 1966, p. 4; Atkins et al. 2006, pp. 8, 122–23
  105. ^ Russell & Lee 2005, pp. 421, 423; Gray 2009, p. 23
  106. ^ Olmsted & Williams 1997, p. 975
  107. ^ a b c Russell & Lee 2005, p. 401; Büchel, Moretto & Woditsch 2003, p. 278
  108. ^ Desch 1914, p. 86
  109. ^ Phillips & Williams 1965, p. 620
  110. ^ Van der Put 1998, p. 123
  111. ^ Klug & Brasted 1958, p. 199
  112. ^ Good et al. 1813
  113. ^ Sequeira 2011, p. 776
  114. ^ Gary 2013
  115. ^ Russell & Lee 2005, pp. 405–06; 423–34
  116. ^ Davidson & Lakin 1973, p. 627
  117. ^ Wiberg 2001, p. 589
  118. ^ Greenwood & Earnshaw 2002, p. 749; Schwartz 2002, p. 679
  119. ^ Antman 2001
  120. ^ Řezanka & Sigler 2008; Sekhon 2012
  121. ^ Emsley 2001, p. 67
  122. ^ Zhang et al. 2008, p. 360
  123. ^ a b Science Learning Hub 2009
  124. ^ Skinner et al. 1979; Tom, Elden & Marsh 2004, p. 135
  125. ^ Büchel 1983, p. 226
  126. ^ Emsley 2001, p. 391
  127. ^ Schauss 1991; Tao & Bolger 1997
  128. ^ Eagleson 1994, p. 450; EVM 2003, pp. 197‒202
  129. ^ a b Nielsen 1998
  130. ^ MacKenzie 2015, p. 36
  131. ^ a b Jaouen & Gibaud 2010
  132. ^ Smith et al. 2014
  133. ^ Stevens & Klarner, p. 205
  134. ^ Sneader 2005, pp. 57–59
  135. ^ Keall, Martin and Tunbridge 1946
  136. ^ Emsley 2001, p. 426
  137. ^ Oldfield et al. 1974, p. 65; Turner 2011
  138. ^ Ba et al. 2010; Daniel-Hoffmann, Sredni & Nitzan 2012; Molina-Quiroz et al. 2012
  139. ^ Peryea 1998
  140. ^ Hager 2006, p. 299
  141. ^ Apseloff 1999
  142. ^ Trivedi, Yung & Katz 2013, p. 209
  143. ^ Emsley 2001, p. 382; Burkhart, Burkhart & Morrell 2011
  144. ^ Thomas, Bialek & Hensel 2013, p. 1
  145. ^ Perry 2011, p. 74
  146. ^ UCR Today 2011; Wang & Robinson 2011; Kinjo et al. 2011
  147. ^ Kauthale et al. 2015
  148. ^ Gunn 2014, pp. 188, 191
  149. ^ Gupta, Mukherjee & Cameotra 1997, p. 280; Thomas & Visakh 2012, p. 99
  150. ^ Muncke 2013
  151. ^ Mokhatab & Poe 2012, p. 271
  152. ^ Craig, Eng & Jenkins 2003, p. 25
  153. ^ McKee 1984
  154. ^ Hai et al. 2012
  155. ^ Kohl & Nielsen 1997, pp. 699–700
  156. ^ Chopra et al. 2011
  157. ^ Le Bras, Wilkie & Bourbigot 2005, p. v
  158. ^ Wilkie & Morgan 2009, p. 187
  159. ^ Locke et al. 1956, p. 88
  160. ^ Carlin 2011, p. 6.2
  161. ^ Evans 1993, pp. 257–28
  162. ^ Corbridge 2013, p. 1149
  163. ^ a b Kaminow & Li 2002, p. 118
  164. ^ Deming 1925, pp. 330 (As2O3), 418 (B2O3; SiO2; Sb2O3); Witt & Gatos 1968, p. 242 (GeO2)
  165. ^ Eagleson 1994, p. 421 (GeO2); Rothenberg 1976, 56, 118–19 (TeO2)
  166. ^ Geckeler 1987, p. 20
  167. ^ Kreith & Goswami 2005, pp. 12–109
  168. ^ Russell & Lee 2005, p. 397
  169. ^ Butterman & Jorgenson 2005, pp. 9–10
  170. ^ Shelby 2005, p. 43
  171. ^ Butterman & Carlin 2004, p. 22; Russell & Lee 2005, p. 422
  172. ^ Träger 2007, pp. 438, 958; Eranna 2011, p. 98
  173. ^ Rao 2002, p. 552; Löffler, Kündig & Dalla Torre 2007, p. 17–11
  174. ^ Guan et al. 2012; WPI-AIM 2012
  175. ^ Klement, Willens & Duwez 1960; Wanga, Dongb & Shek 2004, p. 45
  176. ^ Demetriou et al. 2011; Oliwenstein 2011
  177. ^ Karabulut et al. 2001, p. 15; Haynes 2012, pp. 4–26
  178. ^ Schwartz 2002, pp. 679–80
  179. ^ Carter & Norton 2013, p. 403
  180. ^ Maeder 2013, pp. 3, 9–11
  181. ^ Tominaga 2006, pp. 327–28; Chung 2010, pp. 285–86; Kolobov & Tominaga 2012, p. 149
  182. ^ New Scientist 2014; Hosseini, Wright & Bhaskaran 2014; Farandos et al. 2014
  183. ^ Ordnance Office 1863, p. 293
  184. ^ a b Kosanke 2002, p. 110
  185. ^ Ellern 1968, pp. 246, 326–27
  186. ^ a b Conkling & Mocella 2010, p. 82
  187. ^ Crow 2011; Mainiero 2014
  188. ^ Schwab & Gerlach 1967; Yetter 2012, p. 81; Lipscomb 1972, pp. 2–3, 5–6, 15
  189. ^ Ellern 1968, p. 135; Weingart 1947, p. 9
  190. ^ Conkling & Mocella 2010, p. 83
  191. ^ Conkling & Mocella 2010, pp. 181, 213
  192. ^ a b Ellern 1968, pp. 209–10, 322
  193. ^ Russell 2009, pp. 15, 17, 41, 79–80
  194. ^ Ellern 1968, p. 324
  195. ^ Ellern 1968, p. 328
  196. ^ Conkling & Mocella 2010, p. 171
  197. ^ Conkling & Mocella 2011, pp. 83–84
  198. ^ Berger 1997, p. 91; Hampel 1968, passim
  199. ^ Rochow 1966, p. 41; Berger 1997, pp. 42–43
  200. ^ a b Bomgardner 2013, p. 20
  201. ^ Russell & Lee 2005, p. 395; Brown et al. 2009, p. 489
  202. ^ Haller 2006, p. 4: "The study and understanding of the physics of semiconductors progressed slowly in the 19th and early 20th centuries ... Impurities and defects ... could not be controlled to the degree necessary to obtain reproducible results. This led influential physicists, including W. Pauli and I. Rabi, to comment derogatorily on the 'Physics of Dirt'."; Hoddeson 2007, pp. 25–34 (29)
  203. ^ Bianco et al. 2013
  204. ^ University of Limerick 2014; Kennedy et al. 2014
  205. ^ Lee et al. 2014
  206. ^ Russell & Lee 2005, pp. 421–22, 424
  207. ^ He et al. 2014
  208. ^ Berger 1997, p. 91
  209. ^ ScienceDaily 2012
  210. ^ Reardon 2005; Meskers, Hagelüken & Van Damme 2009, p. 1131
  211. ^ The Economist 2012
  212. ^ Whitten 2007, p. 488
  213. ^ Jaskula 2013
  214. ^ German Energy Society 2008, pp. 43–44
  215. ^ Patel 2012, p. 248
  216. ^ Moore 2104; University of Utah 2014; Xu et al. 2014
  217. ^ Yang et al. 2012, p. 614
  218. ^ Moore 2010, p. 195
  219. ^ Moore 2011
  220. ^ Liu 2014
  221. ^ Bradley 2014; University of Utah 2014
  222. ^ Oxford English Dictionary 1989, 'metalloid'; Gordh, Gordh & Headrick 2003, p. 753
  223. ^ Foster 1936, pp. 212–13; Brownlee et al. 1943, p. 293
  224. ^ Calderazzo, Ercoli & Natta 1968, p. 257
  225. ^ a b Klemm 1950, pp. 133–42; Reilly 2004, p. 4
  226. ^ Walters 1982, pp. 32–33
  227. ^ Tyler 1948, p. 105
  228. ^ Foster & Wrigley 1958, p. 218: "The elements may be grouped into two classes: those that are metals and those that are nonmetals. There is also an intermediate group known variously as metalloids, meta-metals, semiconductors, or semimetals."
  229. ^ Slade 2006, p. 16
  230. ^ Corwin 2005, p. 80
  231. ^ Barsanov & Ginzburg 1974, p. 330
  232. ^ Bradbury et al. 1957, pp. 157, 659
  233. ^ Miller, Lee & Choe 2002, p. 21
  234. ^ King 2004, pp. 196–98; Ferro & Saccone 2008, p. 233
  235. ^ Pashaey & Seleznev 1973, p. 565; Gladyshev & Kovaleva 1998, p. 1445; Eason 2007, p. 294
  236. ^ Johansen & Mackintosh 1970, pp. 121–24; Divakar, Mohan & Singh 1984, p. 2337; Dávila et al. 2002, p. 035411-3
  237. ^ Jezequel & Thomas 1997, pp. 6620–26
  238. ^ Hindman 1968, p. 434: "The high values obtained for the [electrical] resistivity indicate that the metallic properties of neptunium are closer to the semimetals than the true metals. This is also true for other metals in the actinide series."; Dunlap et al. 1970, pp. 44, 46: "... α-Np is a semimetal, in which covalency effects are believed to also be of importance ... For a semimetal having strong covalent bonding, like α-Np ..."
  239. ^ Lister 1965, p. 54
  240. ^ a b c Cotton et al. 1999, p. 502
  241. ^ Pinkerton 1800, p. 81
  242. ^ Goldsmith 1982, p. 526
  243. ^ Zhdanov 1965, pp. 74–75
  244. ^ Friend 1953, p. 68; IUPAC 1959, p. 10; IUPAC 1971, p. 11
  245. ^ IUPAC 2005; IUPAC 2006–
  246. ^ Van Setten et al. 2007, pp. 2460–61; Oganov et al. 2009, pp. 863–64
  247. ^ Housecroft & Sharpe 2008, p. 331; Oganov 2010, p. 212
  248. ^ Housecroft & Sharpe 2008, p. 333
  249. ^ Kross 2011
  250. ^ Berger 1997, p. 37
  251. ^ Greenwood & Earnshaw 2002, p. 144
  252. ^ Kopp, Lipták & Eren 2003, p. 221
  253. ^ Prudenziati 1977, p. 242
  254. ^ Berger 1997, pp. 84, 87
  255. ^ Mendeléeff 1897, p. 57
  256. ^ a b Rayner-Canham & Overton 2006, p. 291
  257. ^ Siekierski & Burgess 2002, p. 63
  258. ^ Wogan 2014
  259. ^ Siekierski & Burgess 2002, p. 86
  260. ^ Greenwood & Earnshaw 2002, p. 141; Henderson 2000, p. 58; Housecroft & Sharpe 2008, pp. 360–72
  261. ^ Parry et al. 1970, pp. 438, 448–51
  262. ^ a b Fehlner 1990, p. 202
  263. ^ Owen & Brooker 1991, p. 59; Wiberg 2001, p. 936
  264. ^ a b Greenwood & Earnshaw 2002, p. 145
  265. ^ Houghton 1979, p. 59
  266. ^ Fehlner 1990, p. 205
  267. ^ Fehlner 1990, pp. 204–05, 207
  268. ^ Greenwood 2001, p. 2057
  269. ^ Salentine 1987, pp. 128–32; MacKay, MacKay & Henderson 2002, pp. 439–40; Kneen, Rogers & Simpson 1972, p. 394; Hiller & Herber 1960, inside front cover; p. 225
  270. ^ Sharp 1983, p. 56
  271. ^ Fokwa 2014, p. 10
  272. ^ a b Gillespie 1998
  273. ^ a b Haaland et al. 2000
  274. ^ a b c d e f Puddephatt & Monaghan 1989, p. 59
  275. ^ Mahan 1965, p. 485
  276. ^ Danaith 2008, p. 81.
  277. ^ Lidin 1996, p. 28
  278. ^ Kondrat'ev & Mel'nikova 1978
  279. ^ Holderness & Berry 1979, p. 111; Wiberg 2001, p. 980
  280. ^ Toy 1975, p. 506
  281. ^ a b c d e f g h Rao 2002, p. 22
  282. ^ Fehlner 1992, p. 1
  283. ^ Haiduc & Zuckerman 1985, p. 82
  284. ^ a b Greenwood & Earnshaw 2002, p. 331
  285. ^ Wiberg 2001, p. 824
  286. ^ Rochow 1973, pp. 1337‒38
  287. ^ a b Russell & Lee 2005, p. 393
  288. ^ Zhang 2002, p. 70
  289. ^ Sacks 1998, p. 287
  290. ^ Rochow 1973, pp. 1337, 1340
  291. ^ Allen & Ordway 1968, p. 152
  292. ^ Eagleson 1994, pp. 48, 127, 438, 1194; Massey 2000, p. 191
  293. ^ Orton 2004, p. 7. This is a typical value for high-purity silicon.
  294. ^ Coles & Caplin 1976, p. 106
  295. ^ Glazov, Chizhevskaya & Glagoleva 1969, pp. 59–63; Allen & Broughton 1987, p. 4967
  296. ^ Cotton, Wilkinson & Gaus 1995, p. 393
  297. ^ Wiberg 2001, p. 834
  298. ^ Partington 1944, p. 723
  299. ^ a b c d e Cox 2004, p. 27
  300. ^ a b c d e Hiller & Herber 1960, inside front cover; p. 225
  301. ^ Kneen, Rogers and Simpson 1972, p. 384
  302. ^ a b c Bailar, Moeller & Kleinberg 1965, p. 513
  303. ^ Cotton, Wilkinson & Gaus 1995, pp. 319, 321
  304. ^ Smith 1990, p. 175
  305. ^ Poojary, Borade & Clearfield 1993
  306. ^ Wiberg 2001, pp. 851, 858
  307. ^ Barmett & Wilson 1959, p. 332
  308. ^ Powell 1988, p. 1
  309. ^ Greenwood & Earnshaw 2002, p. 371
  310. ^ Cusack 1967, p. 193
  311. ^ Russell & Lee 2005, pp. 399–400
  312. ^ a b Greenwood & Earnshaw 2002, p. 373
  313. ^ Moody 1991, p. 273
  314. ^ Russell & Lee 2005, p. 399
  315. ^ Berger 1997, pp. 71–72
  316. ^ Jolly 1966, pp. 125–6
  317. ^ Powell & Brewer 1938
  318. ^ Ladd 1999, p. 55
  319. ^ Everest 1953, p. 4120
  320. ^ Pan, Fu and Huang 1964, p. 182
  321. ^ Monconduit et al. 1992
  322. ^ Richens 1997, p. 152
  323. ^ Rupar et al. 2008
  324. ^ Schwietzer & Pesterfield 2010, p. 190
  325. ^ Jolly & Latimer 1951, p. 2
  326. ^ Lidin 1996, p. 140
  327. ^ Ladd 1999, p. 56
  328. ^ Wiberg 2001, p. 896
  329. ^ Schwartz 2002, p. 269
  330. ^ Eggins 1972, p. 66; Wiberg 2001, p. 895
  331. ^ Greenwood & Earnshaw 2002, p. 383
  332. ^ Glockling 1969, p. 38; Wells 1984, p. 1175
  333. ^ Cooper 1968, pp. 28–29
  334. ^ Steele 1966, pp. 178, 188–89
  335. ^ Haller 2006, p. 3
  336. ^ See, for example, Walker & Tarn 1990, p. 590
  337. ^ Wiberg 2001, p. 742
  338. ^ a b c Gray, Whitby & Mann 2011
  339. ^ a b Greenwood & Earnshaw 2002, p. 552
  340. ^ Parkes & Mellor 1943, p. 740
  341. ^ Russell & Lee 2005, p. 420
  342. ^ Carapella 1968, p. 30
  343. ^ a b Barfuß et al. 1981, p. 967
  344. ^ Greaves, Knights & Davis 1974, p. 369; Madelung 2004, pp. 405, 410
  345. ^ Bailar & Trotman-Dickenson 1973, p. 558; Li 1990
  346. ^ Bailar, Moeller & Kleinberg 1965, p. 477
  347. ^ Gillespie & Robinson 1963, p. 450
  348. ^ Paul et al. 1971; see also Ahmeda & Rucka 2011, pp. 2893–94
  349. ^ Gillespie & Passmore 1972, p. 478
  350. ^ Van Muylder & Pourbaix 1974, p. 521
  351. ^ Kolthoff & Elving 1978, p. 210
  352. ^ Moody 1991, pp. 248–49
  353. ^ Cotton & Wilkinson 1999, pp. 396, 419
  354. ^ Eagleson 1994, p. 91
  355. ^ a b Massey 2000, p. 267
  356. ^ Timm 1944, p. 454
  357. ^ Partington 1944, p. 641; Kleinberg, Argersinger & Griswold 1960, p. 419
  358. ^ Morgan 1906, p. 163; Moeller 1954, p. 559
  359. ^ Corbridge 2013, pp. 122, 215
  360. ^ Douglade 1982
  361. ^ Zingaro 1994, p. 197; Emeléus & Sharpe 1959, p. 418; Addison & Sowerby 1972, p. 209; Mellor 1964, p. 337
  362. ^ Pourbaix 1974, p. 521; Eagleson 1994, p. 92; Greenwood & Earnshaw 2002, p. 572
  363. ^ Wiberg 2001, pp. 750, 975; Silberberg 2006, p. 314
  364. ^ Sidgwick 1950, p. 784; Moody 1991, pp. 248–9, 319
  365. ^ Krannich & Watkins 2006
  366. ^ Greenwood & Earnshaw 2002, p. 553
  367. ^ Dunstan 1968, p. 433
  368. ^ Parise 1996, p. 112
  369. ^ Carapella 1968a, p. 23
  370. ^ Moss 1952, pp. 174, 179
  371. ^ Dupree, Kirby & Freyland 1982, p. 604; Mhiaoui, Sar, & Gasser 2003
  372. ^ Kotz, Treichel & Weaver 2009, p. 62
  373. ^ Cotton et al. 1999, p. 396
  374. ^ King 1994, p. 174
  375. ^ Lidin 1996, p. 372
  376. ^ Lindsjö, Fischer & Kloo 2004
  377. ^ Friend 1953, p. 87
  378. ^ Fesquet 1872, pp. 109–14
  379. ^ Greenwood & Earnshaw 2002, p. 553; Massey 2000, p. 269
  380. ^ King 1994, p. 171
  381. ^ Turova 2011, p. 46
  382. ^ Pourbaix 1974, p. 530
  383. ^ a b Wiberg 2001, p. 764
  384. ^ House 2008, p. 497
  385. ^ Mendeléeff 1897, p. 274
  386. ^ Emsley 2001, p. 428
  387. ^ a b Kudryavtsev 1974, p. 78
  388. ^ Bagnall 1966, pp. 32–33, 59, 137
  389. ^ Swink et al. 1966; Anderson et al. 1980
  390. ^ Ahmed, Fjellvåg & Kjekshus 2000
  391. ^ Chizhikov & Shchastlivyi 1970, p. 28
  392. ^ Kudryavtsev 1974, p. 77
  393. ^ Stuke 1974, p. 178; Donohue 1982, pp. 386–87; Cotton et al. 1999, p. 501
  394. ^ Becker, Johnson & Nussbaum 1971, p. 56
  395. ^ a b Berger 1997, p. 90
  396. ^ Chizhikov & Shchastlivyi 1970, p. 16
  397. ^ Jolly 1966, pp. 66–67
  398. ^ Schwietzer & Pesterfield 2010, p. 239
  399. ^ Cotton et al. 1999, p. 498
  400. ^ Wells 1984, p. 715
  401. ^ Wiberg 2001, p. 588
  402. ^ Mellor 1964a, p.  30; Wiberg 2001, p. 589
  403. ^ Greenwood & Earnshaw 2002, pp. 765–66
  404. ^ Bagnall 1966, pp. 134–51; Greenwood & Earnshaw 2002, p. 786
  405. ^ Detty & O'Regan 1994, pp. 1–2
  406. ^ Hill & Holman 2000, p. 124
  407. ^ Chang 2002, p. 314
  408. ^ Kent 1950, pp. 1–2; Clark 1960, p. 588; Warren & Geballe 1981
  409. ^ Housecroft & Sharpe 2008, p. 384; IUPAC 2006–, rhombohedral graphite entry
  410. ^ Mingos 1998, p. 171
  411. ^ Wiberg 2001, p. 781
  412. ^ Charlier, Gonze & Michenaud 1994
  413. ^ a b c Atkins et al. 2006, pp. 320–21
  414. ^ Savvatimskiy 2005, p. 1138
  415. ^ Togaya 2000
  416. ^ Savvatimskiy 2009
  417. ^ Inagaki 2000, p. 216; Yasuda et al. 2003, pp. 3–11
  418. ^ O'Hare 1997, p. 230
  419. ^ Traynham 1989, pp. 930–31; Prakash & Schleyer 1997
  420. ^ Olmsted & Williams 1997, p. 436
  421. ^ Bailar et al. 1989, p. 743
  422. ^ Moore et al. 1985
  423. ^ House & House 2010, p. 526
  424. ^ Wiberg 2001, p. 798
  425. ^ Eagleson 1994, p. 175
  426. ^ Atkins et al. 2006, p. 121
  427. ^ Russell & Lee 2005, pp. 358–59
  428. ^ Keevil 1989, p. 103
  429. ^ Russell & Lee 2005, pp. 358–60 et seq
  430. ^ Harding, Janes & Johnson 2002, p. 118
  431. ^ a b Metcalfe, Williams & Castka 1974, p. 539
  432. ^ Cobb & Fetterolf 2005, p. 64; Metcalfe, Williams & Castka 1974, p. 539
  433. ^ Ogata, Li & Yip 2002; Boyer et al. 2004, p. 1023; Russell & Lee 2005, p. 359
  434. ^ Cooper 1968, p. 25; Henderson 2000, p. 5; Silberberg 2006, p. 314
  435. ^ Wiberg 2001, p. 1014
  436. ^ Daub & Seese 1996, pp. 70, 109: "Aluminum is not a metalloid but a metal because it has mostly metallic properties."; Denniston, Topping & Caret 2004, p. 57: "Note that aluminum (Al) is classified as a metal, not a metalloid."; Hasan 2009, p. 16: "Aluminum does not have the characteristics of a metalloid but rather those of a metal."
  437. ^ Holt, Rinehart & Wilson c. 2007
  438. ^ Tuthill 2011
  439. ^ Stott 1956, p. 100
  440. ^ Steele 1966, p. 60
  441. ^ Moody 1991, p. 303
  442. ^ Emsley 2001, p. 382
  443. ^ Young et al. 2010, p. 9; Craig & Maher 2003, p. 391. Selenium is "near metalloidal".
  444. ^ Rochow 1957
  445. ^ Rochow 1966, p. 224
  446. ^ Moss 1952, p. 192
  447. ^ a b Glinka 1965, p. 356
  448. ^ Evans 1966, pp. 124–25
  449. ^ Regnault 1853, p. 208
  450. ^ Scott & Kanda 1962, p. 311
  451. ^ Cotton et al. 1999, pp. 496, 503–04
  452. ^ Arlman 1939; Bagnall 1966, pp. 135, 142–43
  453. ^ Chao & Stenger 1964
  454. ^ a b Berger 1997, pp. 86–87
  455. ^ Snyder 1966, p. 242
  456. ^ Fritz & Gjerde 2008, p. 235
  457. ^ Meyer et al. 2005, p. 284; Manahan 2001, p. 911; Szpunar et al. 2004, p. 17
  458. ^ US Environmental Protection Agency 1988, p. 1; Uden 2005, pp. 347‒48
  459. ^ De Zuane 1997, p. 93; Dev 2008, pp. 2‒3
  460. ^ Wiberg 2001, p. 594
  461. ^ Greenwood & Earnshaw 2002, p. 786; Schwietzer & Pesterfield 2010, pp. 242–43
  462. ^ Bagnall 1966, p. 41; Nickless 1968, p. 79
  463. ^ Bagnall 1990, pp. 313–14; Lehto & Hou 2011, p. 220; Siekierski & Burgess 2002, p. 117: "The tendency to form X2− anions decreases down the Group [16 elements] ..."
  464. ^ Legit, Friák & Šob 2010, pp. 214118–18
  465. ^ Manson & Halford 2006, pp. 378, 410
  466. ^ Bagnall 1957, p. 62; Fernelius 1982, p. 741
  467. ^ Bagnall 1966, p. 41; Barrett 2003, p. 119
  468. ^ Hawkes 2010; Holt, Rinehart & Wilson c. 2007; Hawkes 1999, p. 14; Roza 2009, p. 12
  469. ^ Keller 1985
  470. ^ Harding, Johnson & Janes 2002, p. 61
  471. ^ Long & Hentz 1986, p. 58
  472. ^ Vasáros & Berei 1985, p. 109
  473. ^ Haissinsky & Coche 1949, p. 400
  474. ^ Brownlee et al. 1950, p. 173
  475. ^ Hermann, Hoffmann & Ashcroft 2013
  476. ^ Siekierski & Burgess 2002, pp. 65, 122
  477. ^ Emsley 2001, p. 48
  478. ^ Rao & Ganguly 1986
  479. ^ Krishnan et al. 1998
  480. ^ Glorieux, Saboungi & Enderby 2001
  481. ^ Millot et al. 2002
  482. ^ Vasáros & Berei 1985, p. 117
  483. ^ Kaye & Laby 1973, p. 228
  484. ^ Samsonov 1968, p. 590
  485. ^ Korenman 1959, p. 1368
  486. ^ Rossler 1985, pp. 143–44
  487. ^ Champion et al. 2010
  488. ^ Borst 1982, pp. 465, 473
  489. ^ Batsanov 1971, p. 811
  490. ^ Swalin 1962, p. 216; Feng & Lin 2005, p. 157
  491. ^ Schwietzer & Pesterfield 2010, pp. 258–60
  492. ^ Hawkes 1999, p. 14
  493. ^ Olmsted & Williams 1997, p. 328; Daintith 2004, p. 277
  494. ^ Eberle1985, pp. 213–16, 222–27
  495. ^ Restrepo et al. 2004, p. 69; Restrepo et al. 2006, p. 411
  496. ^ Greenwood & Earnshaw 2002, p. 804
  497. ^ Greenwood & Earnshaw 2002, p. 803
  498. ^ Wiberg 2001, p. 416
  499. ^ Craig & Maher 2003, p. 391; Schroers 2013, p. 32; Vernon 2013, pp. 1704–05
  500. ^ Cotton et al. 1999, p. 42
  501. ^ Marezio & Licci 2000, p. 11
  502. ^ a b Vernon 2013, p. 1705
  503. ^ Russell & Lee 2005, p. 5
  504. ^ Parish 1977, pp. 178, 192–93
  505. ^ Eggins 1972, p. 66; Rayner-Canham & Overton 2006, pp. 29–30
  506. ^ Atkins et al. 2006, pp. 320–21; Bailar et al. 1989, pp. 742–43
  507. ^ Rochow 1966, p. 7; Taniguchi et al. 1984, p. 867: "... black phosphorus ... [is] characterized by the wide valence bands with rather delocalized nature."; Morita 1986, p. 230; Carmalt & Norman 1998, p. 7: "Phosphorus ... should therefore be expected to have some metalloid properties."; Du et al. 2010. Interlayer interactions in black phosphorus, which are attributed to van der Waals-Keesom forces, are thought to contribute to the smaller band gap of the bulk material (calculated 0.19 eV; observed 0.3 eV) as opposed to the larger band gap of a single layer (calculated ~0.75 eV).
  508. ^ Stuke 1974, p. 178; Cotton et al. 1999, p. 501; Craig & Maher 2003, p. 391
  509. ^ Steudel 1977, p. 240: "... considerable orbital overlap must exist, to form intermolecular, many-center ... [sigma] bonds, spread through the layer and populated with delocalized electrons, reflected in the properties of iodine (lustre, color, moderate electrical conductivity)."; Segal 1989, p. 481: "Iodine exhibits some metallic properties ..."
  510. ^ a b Lutz et al. 2011, p. 17
  511. ^ Yacobi & Holt 1990, p. 10; Wiberg 2001, p. 160
  512. ^ Greenwood & Earnshaw 2002, pp. 479, 482
  513. ^ Eagleson 1994, p. 820
  514. ^ Oxtoby, Gillis & Campion 2008, p. 508
  515. ^ Brescia et al. 1980, pp. 166–71
  516. ^ Fine & Beall 1990, p. 578
  517. ^ Wiberg 2001, p. 901
  518. ^ Berger 1997, p. 80
  519. ^ Lovett 1977, p. 101
  520. ^ Cohen & Chelikowsky 1988, p. 99
  521. ^ Taguena-Martinez, Barrio & Chambouleyron 1991, p. 141
  522. ^ Ebbing & Gammon 2010, p. 891
  523. ^ Asmussen & Reinhard 2002, p. 7
  524. ^ Deprez & McLachan 1988
  525. ^ Addison 1964 (P, Se, Sn); Marković, Christiansen & Goldman 1998 (Bi); Nagao et al. 2004
  526. ^ Lide 2005; Wiberg 2001, p. 423: At
  527. ^ Cox 1997, pp. 182‒86
  528. ^ MacKay, MacKay & Henderson 2002, p. 204
  529. ^ Baudis 2012, pp. 207–08
  530. ^ Wiberg 2001, p. 741
  531. ^ Chizhikov & Shchastlivyi 1968, p. 96
  532. ^ Greenwood & Earnshaw 2002, pp. 140–41, 330, 369, 548–59, 749: B, Si, Ge, As, Sb, Te
  533. ^ Kudryavtsev 1974, p. 158
  534. ^ Greenwood & Earnshaw 2002, pp. 271, 219, 748–49, 886: C, Al, Se, Po, At; Wiberg 2001, p. 573: Se
  535. ^ United Nuclear 2013
  536. ^ Zalutsky & Pruszynski 2011, p. 181

Sources

Further reading