Jump to content

Metalloid: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
→‎Carbon: add content re carbon anions
(17 intermediate revisions by the same user not shown)
Line 7: Line 7:
Physically, metalloids usually have a metallic appearance but they are brittle and only fair conductors of electricity; chemically, they mostly behave as (weak) nonmetals. They can form alloys with metals. Ordinarily, most of the other physical and chemical properties of metalloids are intermediate in nature.
Physically, metalloids usually have a metallic appearance but they are brittle and only fair conductors of electricity; chemically, they mostly behave as (weak) nonmetals. They can form alloys with metals. Ordinarily, most of the other physical and chemical properties of metalloids are intermediate in nature.


Being too brittle to have any structural uses, metalloids and their compounds find use in alloys, biological agents, flame retardants, glasses, optical storage and semiconductors. The electrical properties of silicon and germanium enabled the establishment of the [[semiconductor industry]] in the 1950s and the development of [[Solid-state (electronics)|solid-state electronics]] from the early 1960s onward.<ref>[[#Chedd1969|Chedd 1969, pp.&nbsp;58, 78]]; [[#NRC1984|National Research Council 1984, p.&nbsp; 43]]</ref>
Metalloids and their compounds are too brittle to have any structural uses. They are used in alloys, biological agents, flame retardants, glasses, optical storage and semiconductors. The electrical properties of silicon and germanium enabled the establishment of the [[semiconductor industry]] in the 1950s and the development of [[Solid-state (electronics)|solid-state electronics]] from the early 1960s onward.<ref>[[#Chedd1969|Chedd 1969, pp.&nbsp;58, 78]]; [[#NRC1984|National Research Council 1984, p.&nbsp; 43]]</ref>


The term ''metalloid'' originally referred to nonmetals. Its more recent meaning, as a category of elements with intermediate or hybrid properties, did not become widespread until 1940–1960. Metalloids are sometimes called semimetals, a practice that has been discouraged.<ref name=Atkins2010p20/> This is because the term ''[[semimetal]]'' 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 a chemical element.
The term ''metalloid'' originally referred to nonmetals. Its more recent meaning, as a category of elements with intermediate or hybrid properties, did not become widespread until 1940–1960. Metalloids are sometimes called semimetals, a practice that has been discouraged.<ref name=Atkins2010p20/> This is because the term ''[[semimetal]]'' 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 a chemical element.


==Definition==
==Definition==
At a conceptual level, metalloids are usually regarded as a third category of chemical elements alongside, and occupying a [[fuzzy logic|fuzzy]] 'buffer zone' between, those of metals and nonmetals.<ref name=roher>[[#Roher2001|Roher 2001, pp.&nbsp;4–6]]</ref>{{#tag:ref|On the fuzziness of metalloids see, for example: Rouvray;<ref>[[#Rouvray1995|Rouvray 1995, p.&nbsp;546]]. Rouvray submits that classifying the [[electrical conductivity]] of the elements using the overlapping domains of metals, metalloids, and nonmetals better reflects reality than a strictly black or white paradigm.</ref> Cobb & Fetterolf;<ref>[[#Cobb2005|Cobb & Fetterolf 2005, p.&nbsp;64]]: 'The division between metals and nonmetals is rather fuzzy, so the elements in the immediate vicinity of the zigzag staircase line are called metalloids, which means they don't fit either definition exactly.'</ref> and Fellet.<ref>[[#Fellet2011|Fellet 2011]]: 'Chemistry has all sorts of fuzzy definitions'.</ref> For the 'buffer zone' terminology see [[Eugene G. Rochow|Rochow]].<ref>[[#Rochow1977|Rochow 1977, p.&nbsp;14]]</ref>|group=n}} At a practical level, there is no universally agreed, rigorous definition of a metalloid.<ref>[[#Goldsmith1982|Goldsmith 1982, p.&nbsp;526]]; [[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]</ref> The feasibility of establishing a specific definition has also been questioned, noting anomalies that can be found in several such attempted constructs.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1687]]</ref> Classifying any particular element as a metalloid has been described as 'arbitrary'.<ref>[[#Sharp1981|Sharp 1981, p.&nbsp;299]]</ref>
Metalloids are usually regarded as a third category of chemical elements occupying a [[fuzzy logic|fuzzy]] 'buffer zone' between metals and nonmetals.<ref name=roher>[[#Roher2001|Roher 2001, pp.&nbsp;4–6]]</ref>{{#tag:ref|On the fuzziness of metalloids see, for example: Rouvray;<ref>[[#Rouvray1995|Rouvray 1995, p.&nbsp;546]]. Rouvray submits that classifying the [[electrical conductivity]] of the elements using the overlapping domains of metals, metalloids, and nonmetals better reflects reality than a strictly black or white paradigm.</ref> Cobb & Fetterolf;<ref>[[#Cobb2005|Cobb & Fetterolf 2005, p.&nbsp;64]]: 'The division between metals and nonmetals is rather fuzzy, so the elements in the immediate vicinity of the zigzag staircase line are called metalloids, which means they don't fit either definition exactly.'</ref> and Fellet.<ref>[[#Fellet2011|Fellet 2011]]: 'Chemistry has all sorts of fuzzy definitions'.</ref> For the 'buffer zone' terminology see [[Eugene G. Rochow|Rochow]].<ref>[[#Rochow1977|Rochow 1977, p.&nbsp;14]]</ref>|group=n}} There is no universally agreed, rigorous definition of a metalloid.<ref>[[#Goldsmith1982|Goldsmith 1982, p.&nbsp;526]]; [[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]</ref> The feasibility of establishing a specific definition has also been questioned, as anomalies can be found in such attempted constructs.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1687]]</ref> Classifying an element as a metalloid has been described as 'arbitrary'.<ref>[[#Sharp1981|Sharp 1981, p.&nbsp;299]]</ref>


===Generic===
===Generic===
A generic definition of a metalloid, based on attributes consistently cited in the literature, is that a metalloid is a chemical element that has properties that are in between or a mixture of those of metals and nonmetals and is consequently difficult to classify unambiguously as either a metal or a nonmetal. 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.'<ref>[[#Cusack1987|Cusack 1987, p.&nbsp;360]]</ref> The half-way character of these elements is also characterised by their mixed properties, and categorisation difficulties: '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.'<ref>[[#Kelter2009|Kelter, Mosher & Scott 2009, p.&nbsp;268]]</ref></I> These difficulties can be accommodated by recognising another category of elements: 'Chemists sometimes use the name metalloid...for these elements difficult to classify one way or the other.'<ref name="Hill 2000, p. 41">[[#Hill2000|Hill & Holman 2000, p.&nbsp;41]]</ref> A few authors are more explicit: '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...'.<ref>[[#King1979|King 1979, p.&nbsp;13]]</ref> Finally, some authors take a broader or more whimsical view as to the nature of metalloids, referring to them as 'elements that...are somewhat of a cross between metals and nonmetals'<ref>[[#Moore2011|Moore 2011, p. 81]]</ref> or 'weird in-between elements.'<ref>[[#Gray2010|Gray 2010]]</ref>
A metalloid is a chemical element that has properties in between or a mixture of those of metals and nonmetals and is difficult to classify as a metal or a nonmetal. 'In chemistry a metalloid is an element with properties intermediate between those of metals and nonmetals.'<ref>[[#Cusack1987|Cusack 1987, p.&nbsp;360]]</ref> These elements are characterised by their mixed properties, and categorisation difficulties: '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.'<ref>[[#Kelter2009|Kelter, Mosher & Scott 2009, p.&nbsp;268]]</ref></I> These difficulties can be accommodated by recognising another category of elements: 'Chemists sometimes use the name metalloid...for these elements difficult to classify one way or the other.'<ref name="Hill 2000, p. 41">[[#Hill2000|Hill & Holman 2000, p.&nbsp;41]]</ref> A few authors are more explicit: '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...'.<ref>[[#King1979|King 1979, p.&nbsp;13]]</ref> Metalloids have also been referred to as 'elements that...are somewhat of a cross between metals and nonmetals'<ref>[[#Moore2011|Moore 2011, p. 81]]</ref> or 'weird in-between elements.'<ref>[[#Gray2010|Gray 2010]]</ref>


The criterion that metalloids be difficult to unambiguously classify one way or the other is a key tenet. In contrast, elements such as sodium and potassium 'have metallic properties to a high degree' and fluorine, chlorine and oxygen 'are almost exclusively nonmetallic.'<ref name=Hopkins>[[#Hopkins1956|Hopkins & Bailar 1956, p.&nbsp;458]]</ref> Although most other elements have a mixture of metallic and nonmetallic properties,<ref name=Hopkins/> most such elements can be classified as either metals or nonmetals according to which set of properties is more pronounced.<ref>[[#Glinka1965|Glinka 1965, p.&nbsp;77]]</ref>{{#tag:ref|[[Gold]], for example, has mixed properties but is still recognised as 'king of metals.' Besides ''metallic'' behaviour (such as high electrical conductivity, and [[cation]] formation), gold shows marked ''nonmetallic'' behaviour:
The criterion that metalloids be difficult to unambiguously classify one way or the other is a key tenet. In contrast, elements such as sodium and potassium 'have metallic properties to a high degree' and fluorine, chlorine and oxygen 'are almost exclusively nonmetallic.'<ref name=Hopkins>[[#Hopkins1956|Hopkins & Bailar 1956, p.&nbsp;458]]</ref> Although most elements have a mixture of metallic and nonmetallic properties,<ref name=Hopkins/> they can be classified as either metals or nonmetals according to which set of properties is more pronounced.<ref>[[#Glinka1965|Glinka 1965, p.&nbsp;77]]</ref>{{#tag:ref|[[Gold]], for example, has mixed properties but is still recognised as 'king of metals.' Besides ''metallic'' behaviour (such as high electrical conductivity, and [[cation]] formation), gold shows marked ''nonmetallic'' behaviour:
* It has the most positive [[electrode potential]].
* It has the most positive [[electrode potential]].
* Its electronegativity of 2.54 is highest among the metals and exceeds that of some nonmetals ([[hydrogen]] 2.2; [[phosphorus]] 2.19; [[radon]] 2.2).
* Its electronegativity of 2.54 is highest among the metals and exceeds that of some nonmetals ([[hydrogen]] 2.2; [[phosphorus]] 2.19; and [[radon]] 2.2).
* It has the most negative [[electron affinity]].
* It has the most negative [[electron affinity]].
* It has the highest [[ionization energy]] among the metals (but for [[zinc]] and [[mercury (element)|mercury]]).
* It has the third-highest [[ionization energy]] among the metals (after [[zinc]] and [[mercury (element)|mercury]]).
* It forms the Au<sup>–</sup> auride [[anion]] thereby behaving analogously to the [[halogen]]s.
* It forms the Au<sup>–</sup> auride [[anion]] thereby behaving analogously to the [[halogen]]s.
* It sometimes has a tendency, known as '[[aurophilicity]]', to bond to itself.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;1279]]</ref>
* It sometimes has a tendency, known as '[[aurophilicity]]', to bond to itself.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;1279]]</ref>
On halogen character, see also Belpassi et al.<ref>[[#Belpassi2006|Belpassi et al. 2006, pp.&nbsp;4543–4]]</ref> who conclude that in the aurides MAu (M = [[alkali metal|Li–Cs]]) gold 'behaves as a halogen, intermediate between [[bromine|Br]] and [[iodine|I]]'. On aurophilicity, see also Schmidbaur and Schier.<ref>[[#Schmidbaur2008|Schmidbaur & Schier 2008, pp.&nbsp;1931–51]]</ref>|group=n}} It is only the elements at or near the margins, ordinarily those that are regarded as lacking a sufficiently clear preponderance of metallic or nonmetallic properties, which are classified as metalloids.<ref>[[#TM1987|Tyler Miller 1987, p.&nbsp;59]]</ref>
On halogen character, see also Belpassi et al.<ref>[[#Belpassi2006|Belpassi et al. 2006, pp.&nbsp;4543–4]]</ref> who conclude that in the aurides MAu (M = [[alkali metal|Li–Cs]]) gold 'behaves as a halogen, intermediate between [[bromine|Br]] and [[iodine|I]]'. On aurophilicity, see also Schmidbaur and Schier.<ref>[[#Schmidbaur2008|Schmidbaur & Schier 2008, pp.&nbsp;1931–51]]</ref>|group=n}} Only the elements at or near the margins, regarded as lacking a sufficiently clear preponderance of metallic or nonmetallic properties, are classified as metalloids.<ref>[[#TM1987|Tyler Miller 1987, p.&nbsp;59]]</ref>


[[Boron]], [[silicon]], [[germanium]], [[arsenic]], [[antimony]] and [[tellurium]] are commonly recognised as metalloids.<ref>[[#Goldsmith1982|Goldsmith 1982, p.&nbsp;526]]; [[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]; [[#Boylan1962|Boylan 1962, p.&nbsp;493]]; [[#Sherman1966|Sherman & Weston 1966, p.&nbsp;64]]; [[#Wulfsberg1991|Wulfsberg 1991, p.&nbsp;201]]; [[#Kotz2009|Kotz, Treichel & Weaver 2009, p.&nbsp;62]]</ref>{{#tag:ref|Mann et al.<ref name=Mann/> refer to these elements as 'the recognized metalloids'.|group=n}} Depending on the author, one or more from among [[selenium]], [[polonium]] or [[astatine]] are sometimes added to this list.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]; [[#Segal1989|Segal 1989, p.&nbsp;965]]; [[#McMurray2009|McMurray & Fay 2009, p.&nbsp;767]]</ref> Boron is sometimes excluded, by itself or together with silicon.<ref>[[#Bucat1983|Bucat 1983, p.&nbsp;26]]; [[#Brown2007|Brown c. 2007]]</ref> Tellurium is sometimes not regarded as a metalloid.<ref name="Swift EH 1962, p.,[object Object], ,[object Object],100">[[#Swift1962|Swift & Schaefer 1962, p.&nbsp;100]]</ref> The inclusion of antimony, polonium and astatine as metalloids has also been questioned.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]; [[#Hawkes2010|Hawkes 2010]]; [[#Holt2007|Holt, Rinehart & Wilson c. 2007]]</ref>
Boron, silicon, germanium, arsenic, antimony and tellurium are commonly recognised as metalloids.<ref>[[#Goldsmith1982|Goldsmith 1982, p.&nbsp;526]]; [[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]; [[#Boylan1962|Boylan 1962, p.&nbsp;493]]; [[#Sherman1966|Sherman & Weston 1966, p.&nbsp;64]]; [[#Wulfsberg1991|Wulfsberg 1991, p.&nbsp;201]]; [[#Kotz2009|Kotz, Treichel & Weaver 2009, p.&nbsp;62]]</ref>{{#tag:ref|Mann et al.<ref name=Mann/> refer to these elements as 'the recognized metalloids'.|group=n}} Depending on the author, one or more from selenium, polonium or astatine are sometimes added to this list.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]; [[#Segal1989|Segal 1989, p.&nbsp;965]]; [[#McMurray2009|McMurray & Fay 2009, p.&nbsp;767]]</ref> Boron is sometimes excluded, by itself or together with silicon.<ref>[[#Bucat1983|Bucat 1983, p.&nbsp;26]]; [[#Brown2007|Brown c. 2007]]</ref> Tellurium is sometimes not regarded as a metalloid.<ref name="Swift EH 1962, p.,[object Object], ,[object Object],100">[[#Swift1962|Swift & Schaefer 1962, p.&nbsp;100]]</ref> The inclusion of antimony, polonium and astatine as metalloids has also been questioned.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]; [[#Hawkes2010|Hawkes 2010]]; [[#Holt2007|Holt, Rinehart & Wilson c. 2007]]</ref>


Some other elements are occasionally classified as metalloids. These elements include<ref>[[#Dunstan1968|Dunstan 1968, pp.&nbsp;310, 409]]. Dunstan lists Be, Al, Ge (maybe), As, Se (maybe), Sn, Sb, Te, Pb, Bi and Po as metalloids (pp.&nbsp;310, 323, 409, 419).</ref> [[hydrogen]],<ref>[[#Tilden1876|Tilden 1876, pp.&nbsp;172, 198–201]]; [[#Smith1994|Smith 1994, p.&nbsp;252]]; [[#Bodner1993|Bodner & Pardue 1993, p.&nbsp;354]]</ref> [[beryllium]],<ref>[[#Bassett1966|Bassett et al. 1966, p.&nbsp;127]]</ref> [[nitrogen]],<ref name=rausch>[[#Rausch1960|Rausch 1960]]</ref> [[phosphorus]],<ref>[[#Thayer1977|Thayer 1977, p.&nbsp;604]]; [[#Warren1981|Warren & Geballe 1981]]; [[#M&E|Masters & Ela 2008, p.&nbsp;190]]</ref> [[sulfur]],<ref>[[#Warren1981|Warren & Geballe 1981]]; [[#Chalmers1959|Chalmers 1959, p.&nbsp;72]]; [[#United1965|US Bureau of Naval Personnel 1965, p.&nbsp;26]]</ref> [[zinc]],<ref>[[#Siebring1967|Siebring 1967, p.&nbsp;513]]</ref> [[gallium]],<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;282]]</ref> [[tin]], [[iodine]],<ref>[[#Rausch1960|Rausch 1960]]; [[#Friend1953|Friend 1953, p.&nbsp;68]]</ref> [[lead]],<ref>[[#Murray1928|Murray 1928, p.&nbsp;1295]]</ref> [[bismuth]]<ref name="Swift EH 1962, p.,[object Object], ,[object Object],100" /> and [[radon]].<ref>[[#Hampel&H1966|Hampel & Hawley 1966, p.&nbsp;950]];
Some other elements are occasionally classified as metalloids. These elements include<ref>[[#Dunstan1968|Dunstan 1968, pp.&nbsp;310, 409]]. Dunstan lists Be, Al, Ge (maybe), As, Se (maybe), Sn, Sb, Te, Pb, Bi and Po as metalloids (pp.&nbsp;310, 323, 409, 419).</ref> hydrogen,<ref>[[#Tilden1876|Tilden 1876, pp.&nbsp;172, 198–201]]; [[#Smith1994|Smith 1994, p.&nbsp;252]]; [[#Bodner1993|Bodner & Pardue 1993, p.&nbsp;354]]</ref> [[beryllium]],<ref>[[#Bassett1966|Bassett et al. 1966, p.&nbsp;127]]</ref> [[nitrogen]],<ref name=rausch>[[#Rausch1960|Rausch 1960]]</ref> phosphorus,<ref>[[#Thayer1977|Thayer 1977, p.&nbsp;604]]; [[#Warren1981|Warren & Geballe 1981]]; [[#M&E|Masters & Ela 2008, p.&nbsp;190]]</ref> [[sulfur]],<ref>[[#Warren1981|Warren & Geballe 1981]]; [[#Chalmers1959|Chalmers 1959, p.&nbsp;72]]; [[#United1965|US Bureau of Naval Personnel 1965, p.&nbsp;26]]</ref> [[zinc]],<ref>[[#Siebring1967|Siebring 1967, p.&nbsp;513]]</ref> [[gallium]],<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;282]]</ref> [[tin]], [[iodine]],<ref>[[#Rausch1960|Rausch 1960]]; [[#Friend1953|Friend 1953, p.&nbsp;68]]</ref> [[lead]],<ref>[[#Murray1928|Murray 1928, p.&nbsp;1295]]</ref> [[bismuth]]<ref name="Swift EH 1962, p.,[object Object], ,[object Object],100" /> and radon.<ref>[[#Hampel&H1966|Hampel & Hawley 1966, p.&nbsp;950]];
[[#Stein1985|Stein 1985]]; [[#Stein1987|Stein 1987, pp.&nbsp;240, 247–8]]</ref> The term metalloid has also been used to refer to elements that exhibit metallic lustre and electrical conductivity, and that are [[amphoterism|amphoteric]]; examples include arsenic, antimony, [[vanadium]], [[chromium]], [[molybdenum]], [[tungsten]], tin, lead and aluminium.<ref>[[#Hatcher1949|Hatcher 1949, p.&nbsp;223]]; [[#Secrist|Secrist & Powers 1966, p. 459]]</ref> Other elements that have been referred to as metalloids are the [[poor metal]]s,<ref>[[#Taylor1960|Taylor 1960, p.&nbsp;614]]</ref> and nonmetals (such as carbon or nitrogen) that can form [[alloy]]s with,<ref>[[#Considine1984|Considine & Considine 1984, p.&nbsp;568]]; [[#Cegielski1998|Cegielski 1998, p.&nbsp;147]]; [[#TheAmerican2005|''The American heritage science dictionary 2005'' p.&nbsp;397]]</ref> or modify the properties of,<ref>[[#Woodward1948|Woodward 1948, p.&nbsp;1]]</ref> metals.
[[#Stein1985|Stein 1985]]; [[#Stein1987|Stein 1987, pp.&nbsp;240, 247–8]]</ref> The term metalloid has also been used to refer to elements that exhibit metallic lustre and electrical conductivity, and that are [[amphoterism|amphoteric]]; examples include arsenic, antimony, [[vanadium]], [[chromium]], [[molybdenum]], [[tungsten]], tin, lead and aluminium.<ref>[[#Hatcher1949|Hatcher 1949, p.&nbsp;223]]; [[#Secrist|Secrist & Powers 1966, p. 459]]</ref> The [[poor metal]]s,<ref>[[#Taylor1960|Taylor 1960, p.&nbsp;614]]</ref> and nonmetals (such as carbon or nitrogen) that can form [[alloy]]s with,<ref>[[#Considine1984|Considine & Considine 1984, p.&nbsp;568]]; [[#Cegielski1998|Cegielski 1998, p.&nbsp;147]]; [[#TheAmerican2005|''The American heritage science dictionary 2005'' p.&nbsp;397]]</ref> or modify the properties of,<ref>[[#Woodward1948|Woodward 1948, p.&nbsp;1]]</ref> metals have also occasionally been considered as metalloids.


===Specific===
===Specific===
Line 100: Line 100:
| colspan="6" height=10 |
| colspan="6" height=10 |
|}
|}
Metalloids tend to be collectively characterised in terms of generalities or a few broadly indicative physical or chemical properties.<ref name="ReferenceA">[[#Goldsmith1982|Goldsmith 1982, p.&nbsp;526]]</ref> A single quantitative criterion, such as [[electronegativity]], is also occasionally mentioned.{{#tag:ref|[[Eugene G. Rochow|Rochow]]<ref>[[#Rochow1966|Rochow 1966, pp.&nbsp;1, 4–7]]</ref> concluded that no single measurement indicates exactly which elements are properly classified as metalloids, and that, therefore, present-day students and teachers usually agree to use electronegativity as a compromise criterion. He described metalloids as a collection of ''in between'' elements, of electronegativity 1.8 to 2.2 (classical [[Linus Pauling|Pauling]] scale), that "...resemble metals, yet are not completely metallic either in appearance or in properties," and "...are neither metals nor nonmetals." In mentioning 'appearance', Rochow referred to the intermediate reflectivity values of the commonly recognised metalloids,<ref>[[#Lagrenaudie|Lagrenaudie 1953]]; [[#Rochow1966|Rochow 1966, pp.&nbsp;23, 25]]</ref> as compared to the intermediate to typically high<ref>[[#Askeland|Askeland, Fulay & Wright 2011, p.&nbsp;806]]</ref> values of the metals,<ref>[[#Born|Born & Wolf 1999, p.&nbsp;746]]</ref> and the zero or low (mostly)<ref>[[#Burakowski|Burakowski & Wierzchoń 1999, p.&nbsp;336]]</ref> to intermediate reflectivity values of the non-metals.<ref>[[#Olechna|Olechna & Knox 1965, pp. A991‒2]]</ref> See also, for example:
Metalloids tend to be characterised in terms of generalities or a few broadly indicative physical or chemical properties.<ref name="ReferenceA">[[#Goldsmith1982|Goldsmith 1982, p.&nbsp;526]]</ref> A single quantitative criterion, such as [[electronegativity]], is also occasionally mentioned.{{#tag:ref|[[Eugene G. Rochow|Rochow]]<ref>[[#Rochow1966|Rochow 1966, pp.&nbsp;1, 4–7]]</ref> concluded that no single measurement indicates exactly which elements are properly classified as metalloids, and that, therefore, present-day students and teachers usually agree to use electronegativity as a compromise criterion. He described metalloids as a collection of ''in between'' elements, of electronegativity 1.8 to 2.2 (classical [[Linus Pauling|Pauling]] scale), that "...resemble metals, yet are not completely metallic either in appearance or in properties," and "...are neither metals nor nonmetals." In mentioning 'appearance', Rochow referred to the intermediate reflectivity values of the commonly recognised metalloids,<ref>[[#Lagrenaudie|Lagrenaudie 1953]]; [[#Rochow1966|Rochow 1966, pp.&nbsp;23, 25]]</ref> compared to the intermediate to typically high<ref>[[#Askeland|Askeland, Fulay & Wright 2011, p.&nbsp;806]]</ref> values of the metals,<ref>[[#Born|Born & Wolf 1999, p.&nbsp;746]]</ref> and the zero or low (mostly)<ref>[[#Burakowski|Burakowski & Wierzchoń 1999, p.&nbsp;336]]</ref> to intermediate reflectivity values of the non-metals.<ref>[[#Olechna|Olechna & Knox 1965, pp. A991‒2]]</ref> See also, for example:
* Hill and Hollman,<ref name="Hill 2000, p. 41"/> who characterise metalloids (in part) on the basis that they are 'poor conductors of electricity with atomic conductance usually less than 10<sup>−3</sup> but greater than 10<sup>−5</sup>&nbsp;ohm<sup>−1</sup> cm<sup>−4</sup>'.
* Hill and Hollman,<ref name="Hill 2000, p. 41"/> who characterise metalloids (in part) on the basis that they are 'poor conductors of electricity with atomic conductance usually less than 10<sup>−3</sup> but greater than 10<sup>−5</sup>&nbsp;ohm<sup>−1</sup> cm<sup>−4</sup>'.
* Bond,<ref>[[#Bond2005|Bond 2005, p.&nbsp;3]]</ref> who suggests that 'one criterion for distinguishing semi-metals from true metals under normal conditions is that the co-ordination number of the former is never greater than eight'.
* Bond,<ref>[[#Bond2005|Bond 2005, p.&nbsp;3]]</ref> who suggests that 'one criterion for distinguishing semi-metals from true metals under normal conditions is that the co-ordination number of the former is never greater than eight'.
* Edwards et al.,<ref>[[#Edwards2010|Edwards et al. 2010, p.&nbsp;958]]</ref> who state that, 'Using the [[#Gold|Goldhammer-Herzfeld criterion]] with measured atomic electronic polarizabilities and condensed phase molar volumes allows one to readily predict which elements are metallic, which are nonmetallic, and which are borderline when in their condensed phases (solid or liquid).'|group=n}} In contrast, Jones<ref>[[#Jones2010|Jones 2010, p.&nbsp;169]]</ref> (writing on the role of classification in science) observed that, 'Classes are usually defined by more than two attributes.'
* Edwards et al.,<ref>[[#Edwards2010|Edwards et al. 2010, p.&nbsp;958]]</ref> who state that, 'Using the [[#Gold|Goldhammer-Herzfeld criterion]] with measured atomic electronic polarizabilities and condensed phase molar volumes allows one to readily predict which elements are metallic, which are nonmetallic, and which are borderline when in their condensed phases (solid or liquid).'|group=n}} Jones<ref>[[#Jones2010|Jones 2010, p.&nbsp;169]]</ref> (writing on the role of classification in science) observed that, 'Classes are usually defined by more than two attributes.'


Masterton and Slowinski<ref>[[#Masterton1977|Masterton & Slowinski 1977, p.&nbsp;160]]. They list B, Si, Ge, As, Sb and Te as metalloids, and comment that Po and At are ordinarily classified as metalloids but add that, 'since very little is known about their chemical and physical properties, and such classification must be rather arbitrary.'</ref> offer a more specific treatment. They wrote that metalloids have ionization energies clustering around 200 kcal/mol, and electronegativity values close to 2.0. They also said that metalloids are typically semiconductors, though antimony and arsenic (semimetals in the physics sense) have electrical conductivities that approach those of metals. Their description, using these three more or less clearly defined properties, encompasses the six elements commonly recognised as metalloids (see table, upper right). Selenium and polonium are probably excluded from this scheme; astatine may or may not be included.{{#tag:ref|Selenium has an ionization energy (IE) of ~226&nbsp;kcal/mol and is sometimes described as a semiconductor. However it has a relatively high 2.55 electronegativity (EN). Polonium has an IE of ~196&nbsp;kcal/mol and a 2.0 EN, but has a metallic band structure.<ref>[[#Kraig2004|Kraig, Roundy & Cohen 2004, p.&nbsp;412]]; [[#Alloul2010|Alloul 2010, p.&nbsp;83]]</ref> Astatine has an estimated IE of ~210±10&nbsp;kcal/mol<ref>[[#NIST2011|NIST 2011]]. They cite [[#Finkelnburg1955|Finkelnburg & Humbach (1955)]] who give a figure of 9.2±0.4&nbsp;eV = 212.2±9.224&nbsp;kcal/mol.</ref> and an EN of 2.2. However its electronic band structure is not known with any great degree of certainty.|group=n}}
Masterton and Slowinski<ref>[[#Masterton1977|Masterton & Slowinski 1977, p.&nbsp;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, 'since very little is known about their chemical and physical properties, and such classification must be rather arbitrary.'</ref> offer a more specific treatment. They write that metalloids have ionization energies around 200 kcal/mol, and electronegativity values close to 2.0. They also say that metalloids are typically semiconductors, though antimony and arsenic (semimetals in the physics sense) have electrical conductivities that approach those of metals. Their description, using these three more or less clearly defined properties, encompasses the six elements commonly recognised as metalloids (see table, upper right). Selenium and polonium are probably excluded from this scheme; astatine may or may not be included.{{#tag:ref|Selenium has an ionization energy (IE) of ~226&nbsp;kcal/mol and is sometimes described as a semiconductor. It has a relatively high 2.55 electronegativity (EN). Polonium has an IE of ~196&nbsp;kcal/mol and a 2.0 EN, but has a metallic band structure.<ref>[[#Kraig2004|Kraig, Roundy & Cohen 2004, p.&nbsp;412]]; [[#Alloul2010|Alloul 2010, p.&nbsp;83]]</ref> Astatine has an estimated IE of ~210±10&nbsp;kcal/mol<ref>[[#NIST2011|NIST 2011]]. They cite [[#Finkelnburg1955|Finkelnburg & Humbach (1955)]] who give a figure of 9.2±0.4&nbsp;eV = 212.2±9.224&nbsp;kcal/mol.</ref> and an EN of 2.2. Its electronic band structure is not known with any great certainty.|group=n}}


The elements commonly recognised as metalloids can also be quantitatively described in terms of their intermediate [[atomic packing factor|packing efficiencies]] (between 34% to 41%) and Goldhammer-Herzfeld criterion metallization ratios (between ~0.85 to 1.1; average 1.0).<ref>[[#Edwards1983|Edwards & Sienko 1983, p.&nbsp;695]]; [[#Edwards2010|Edwards et al. 2010]]</ref>{{#tag:ref|The <span id="Gold"></span>''Goldhammer-[[Karl Herzfeld|Herzfeld]] criterion'' is a ratio that compares the force holding an individual atom's valence electrons in place with the forces, acting on the same electrons, arising 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. Metallic behaviour is then predicted.<ref>[[#Herzfeld1927|Herzfeld 1927]]; [[#Edwards2000|Edwards 2000, pp.&nbsp;100–3]]</ref> Otherwise nonmetallic behaviour is anticipated.|group=n}}
The commonly recognised metalloids can also be quantitatively described in terms of their intermediate [[atomic packing factor|packing efficiencies]] (between 34% to 41%) and Goldhammer-Herzfeld criterion metallization ratios (between ~0.85 to 1.1; average 1.0).<ref>[[#Edwards1983|Edwards & Sienko 1983, p.&nbsp;695]]; [[#Edwards2010|Edwards et al. 2010]]</ref>{{#tag:ref|The <span id="Gold"></span>''Goldhammer-[[Karl Herzfeld|Herzfeld]] criterion'' is a ratio that compares the force holding an individual atom's valence electrons in place with the forces, acting on the same electrons, arising 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. Metallic behaviour is then predicted.<ref>[[#Herzfeld1927|Herzfeld 1927]]; [[#Edwards2000|Edwards 2000, pp.&nbsp;100–3]]</ref> Otherwise nonmetallic behaviour is anticipated.|group=n}}
The packing efficiency of boron is 38%; silicon and germanium 34; arsenic 38.5; antimony 41; and tellurium 36.4.<ref>[[#VanSetten2007|Van Setten et al. 2007, pp.&nbsp;2460–1]]; [[#Russell2005|Russell & Lee 2005, p.&nbsp;7]] (Si, Ge); [[#Pearson1972|Pearson 1972, p.&nbsp;264]] (As, Sb, Te; also black P)</ref> These values are lower than the values of most metals (at least 80% of which have a packing efficiency of at least 68%)<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;1]]</ref>{{#tag:ref|Gallium is unusual (for a metal) in having a packing efficiency of just 39%.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;6–7, 387]]</ref> Other notable values are 42.9 for bismuth<ref name="ReferenceB">[[#Pearson1972|Pearson 1972, p.&nbsp;264]]</ref> and 58.5 for liquid [[mercury (element)]].<ref>[[#Okakjima1972|Okakjima & Shomoji 1972, p.&nbsp;258]]</ref>|group=n}} but higher than those of elements usually classified as nonmetals. Packing efficiencies for nonmetals are: graphite 17%,<ref>[[#Kitaĭgorodskiĭ1961|Kitaĭgorodskiĭ 1961, p.&nbsp;108]]</ref> sulfur 19.2,<ref name="Neuburger">[[#Neuburger1936|Neuburger 1936]]</ref> iodine 23.9,<ref name="Neuburger"/> selenium 24.2,<ref name="Neuburger"/> and black phosphorus 28.5.<ref name="ReferenceB"/> The Goldhammer-Herzfeld criterion ratio values of the recognised metalloids are lower than those of representative and transition metals and higher than those of nonmetals.{{#tag:ref|As the ratio is based on classical arguments<ref>[[#Edwards1999|Edwards 1999, p.&nbsp;416]]</ref> it does not accommodate the finding that polonium, which has a value of ~0.95, adopts a metallic (rather than covalent) crystalline structure, on relativistic grounds.<ref>[[#Steurer2007|Steurer 2007, p.&nbsp;142]]; [[#Pyykkö|Pyykkö 2012, p.&nbsp;56]]</ref> It nevertheless offers a relatively simple first order rationalization for the occurrence of metallic character amongst the elements.<ref name=edwards695>[[#Edwards1983|Edwards & Sienko 1983, p.&nbsp;695]]</ref>|group=n}}
The packing efficiency of boron is 38%; silicon and germanium 34; arsenic 38.5; antimony 41; and tellurium 36.4.<ref>[[#VanSetten2007|Van Setten et al. 2007, pp.&nbsp;2460–1]]; [[#Russell2005|Russell & Lee 2005, p.&nbsp;7]] (Si, Ge); [[#Pearson1972|Pearson 1972, p.&nbsp;264]] (As, Sb, Te; also black P)</ref> These values are lower than the values of most metals (at least 80% of which have a packing efficiency of at least 68%)<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;1]]</ref>{{#tag:ref|Gallium is unusual (for a metal) in having a packing efficiency of just 39%.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;6–7, 387]]</ref> Other notable values are 42.9 for bismuth<ref name="ReferenceB">[[#Pearson1972|Pearson 1972, p.&nbsp;264]]</ref> and 58.5 for liquid mercury.<ref>[[#Okakjima1972|Okakjima & Shomoji 1972, p.&nbsp;258]]</ref>|group=n}} but higher than those of elements usually classified as nonmetals. Packing efficiencies for nonmetals are: graphite 17%,<ref>[[#Kitaĭgorodskiĭ1961|Kitaĭgorodskiĭ 1961, p.&nbsp;108]]</ref> sulfur 19.2,<ref name="Neuburger">[[#Neuburger1936|Neuburger 1936]]</ref> iodine 23.9,<ref name="Neuburger"/> selenium 24.2,<ref name="Neuburger"/> and black phosphorus 28.5.<ref name="ReferenceB"/> The Goldhammer-Herzfeld criterion ratio values of the recognised metalloids are lower than those of representative and transition metals and higher than those of nonmetals.{{#tag:ref|As the ratio is based on classical arguments<ref>[[#Edwards1999|Edwards 1999, p.&nbsp;416]]</ref> it does not accommodate the finding that polonium, which has a value of ~0.95, adopts a metallic (rather than covalent) crystalline structure, on relativistic grounds.<ref>[[#Steurer2007|Steurer 2007, p.&nbsp;142]]; [[#Pyykkö|Pyykkö 2012, p.&nbsp;56]]</ref> It nevertheless offers a relatively simple first order rationalization for the occurrence of metallic character amongst the elements.<ref name=edwards695>[[#Edwards1983|Edwards & Sienko 1983, p.&nbsp;695]]</ref>|group=n}}
{{clear}}
{{clear}}


==Periodic table territory==
==Periodic table territory==

===Location===
===Location===
{{periodic table (metalloid border)}}
{{periodic table (metalloid border)}}
Metalloids cluster on either side of the [[dividing line between metals and nonmetals]]. This can be found, in varying configurations, on some [[periodic table]]s (see the mini-example, in this section). Elements to the lower left of the line generally display increasing metallic behaviour; elements to the upper right display increasing nonmetallic behaviour.<ref name="Hamm 1969, p.,[object Object], ,[object Object],653" /> 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.<ref>[[#Horvath1973|Horvath 1973, p.&nbsp;336]]</ref>
Metalloids lie on either side of the [[dividing line between metals and nonmetals]]. This can be found, in varying configurations, on some [[periodic table]]s. Elements to the lower left of the line generally display increasing metallic behaviour; elements to the upper right display increasing nonmetallic behaviour.<ref name="Hamm 1969, p.,[object Object], ,[object Object],653" /> 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.<ref>[[#Horvath1973|Horvath 1973, p.&nbsp;336]]</ref>


The diagonal positioning of the metalloids represents somewhat of an exception to the phenomenon that elements with similar properties tend to occur in vertical columns.<ref name="Gray91">[[#Gray2009|Gray 2009, p.&nbsp; 9]]</ref> Going across a periodic table row, the nuclear charge increases with atomic number just as there is as a corresponding increase in electrons. The additional 'pull' on outer electrons with increasing nuclear charge generally outweighs the screening efficacy 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.<ref>[[#Booth1972|Booth & Bloom 1972, p.&nbsp;426]]; [[#Cox2004|Cox 2004, pp.&nbsp;17, 18, 27–8]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;305–13]]</ref> Going down a [[main group element|main group]] periodic table column, the effect of increasing nuclear charge is generally outweighed by the effect of additional electrons being further away from the nucleus. With some irregularities, atoms therefore become larger, ionization energy falls, and metallic character increases.<ref>[[#Cox2004|Cox 2004, pp.&nbsp;17–18, 27–8]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;305–13]]</ref> The combined effect of these competing horizontal and vertical [[periodic trends|trends]] is that the location of the metal-nonmetal transition zone shifts to the right in going down a period.<ref name=Gray91/> A related effect can be seen in other [[diagonal relationship|diagonal similarities]] that occur between some elements and their lower right neighbours, such as lithium-magnesium, beryllium-aluminium, carbon-phosphorus, and nitrogen-sulfur.<ref>[[#Rayner2011|Rayner-Canham 2011]]</ref>
The diagonal positioning of the metalloids represents an exception to the observation that elements with similar properties tend to occur in vertical columns.<ref name="Gray91">[[#Gray2009|Gray 2009, p.&nbsp; 9]]</ref> Going along a [[period (periodic table)|period]], the nuclear charge increases with atomic number as there is as an increase in electrons. The additional 'pull' on outer electrons with increasing nuclear charge generally outweighs the screening efficacy 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.<ref>[[#Booth1972|Booth & Bloom 1972, p.&nbsp;426]]; [[#Cox2004|Cox 2004, pp.&nbsp;17, 18, 27–8]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;305–13]]</ref> Going down a [[main group element|main group]], the effect of increasing nuclear charge is generally outweighed by the effect of additional electrons being further away from the nucleus. With some irregularities, atoms become larger, ionization energy falls, and metallic character increases.<ref>[[#Cox2004|Cox 2004, pp.&nbsp;17–18, 27–8]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;305–13]]</ref> The combined effect of these competing horizontal and vertical [[periodic trends|trends]] is that the location of the metal-nonmetal transition zone shifts to the right in going down a group.<ref name=Gray91/> A related effect can be seen in other [[diagonal relationship|diagonal similarities]] between some elements and their lower right neighbours, such as lithium-magnesium, beryllium-aluminium, carbon-phosphorus, and nitrogen-sulfur.<ref>[[#Rayner2011|Rayner-Canham 2011]]</ref>


===Number, composition and alternative treatments===
===Number, composition and alternative treatments===
How many and which elements are metalloids depends on the classification criteria being used. [[Emsley]],<ref>[[#Emsley1971|Emsley 1971, p.&nbsp;1]]</ref> for example, recognised only four: germanium, arsenic, antimony and tellurium. James et al.,<ref>[[#James2000|James et al. 2000, p.&nbsp;480]]</ref> on the other hand, listed twelve: boron, carbon, silicon, germanium, arsenic, selenium, antimony, tellurium, bismuth, polonium, [[ununpentium]] and [[livermorium]]. On average, seven elements are included in [[List of metalloid lists|such lists]].
How many and which elements are metalloids depends on the classification criteria being used. [[Emsley]],<ref>[[#Emsley1971|Emsley 1971, p.&nbsp;1]]</ref> for example, recognised only four: germanium, arsenic, antimony and tellurium; James et al.,<ref>[[#James2000|James et al. 2000, p.&nbsp;480]]</ref> listed twelve: boron, carbon, silicon, germanium, arsenic, selenium, antimony, tellurium, bismuth, polonium, [[ununpentium]] and [[livermorium]]. On average, seven elements are included in [[List of metalloid lists|such lists]].


The absence of a standardized division of the elements into metals, metalloids and nonmetals is not necessarily an issue. There is a more or less continuous progression from the metallic to the nonmetallic. A specified subset of this continuum can potentially serve its particular purpose as well as any other.<ref>[[#Kneen1972|Kneen, Rogers & Simpson 1972, pp.&nbsp;218–220]]</ref> Individual metalloid classification arrangements tend to share common ground with most variations occurring around the indistinct<ref>[[#Chatt1951|Chatt 1951, p.&nbsp;417]]: 'The boundary between metals and metalloids is indefinite<span style="white-space: nowrap">...</span>'; [[#Burrows2009|Burrows et al. 2009, p.&nbsp;1192]]: 'Although the elements are conveniently described as metals, metalloids, and nonmetals, the transitions are not exact<span style="white-space: nowrap">...</span>'.</ref> margins.{{#tag:ref|Jones<ref>[[#Jones2010|Jones 2010, p.&nbsp;170]]</ref> 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'.|group=n}}
With no standardized division of the elements into metals, metalloids and nonmetals, a specified subset of the continuum from the metallic to the nonmetallic can potentially serve its particular purpose as well as any other.<ref>[[#Kneen1972|Kneen, Rogers & Simpson 1972, pp.&nbsp;218–220]]</ref> Individual metalloid classification arrangements tend to share common ground, with most variations occurring around the indistinct<ref>[[#Chatt1951|Chatt 1951, p.&nbsp;417]]: 'The boundary between metals and metalloids is indefinite<span style="white-space: nowrap">...</span>'; [[#Burrows2009|Burrows et al. 2009, p.&nbsp;1192]]: 'Although the elements are conveniently described as metals, metalloids, and nonmetals, the transitions are not exact<span style="white-space: nowrap">...</span>'.</ref> margins.{{#tag:ref|Jones<ref>[[#Jones2010|Jones 2010, p.&nbsp;170]]</ref> 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'.|group=n}}


Some authors do not classify elements bordering the metal-nonmetal dividing line as metalloids noting that a binary classification can facilitate the establishment of some simple rules for determining bond types between metals and nonmetals.<ref name=roher/> Metalloids are grouped instead with metals,<ref>[[#Tyler1948|Tyler 1948, p.&nbsp;105]]; [[#Reilly2002|Reilly 2002, pp.&nbsp;5–6]]</ref> regarded as nonmetals<ref>[[#Hampel1976|Hampel & Hawley 1976, p.&nbsp;174]]</ref> or treated as a sub-category of same.<ref>[[#Goodrich1844|Goodrich 1844, p.&nbsp;264]]; [[#TheChemical1897|''The Chemical News'' 1897, p.&nbsp;189]]; [[#Hampel1976|Hampel & Hawley 1976, p.&nbsp;191]]; [[#Lewis1993|Lewis 1993, p.&nbsp;835]]; [[#Hérold2006|Hérold 2006, pp.&nbsp;149–50]]</ref>{{#tag:ref|Oderberg<ref>[[#Oderberg2007|Oderberg 2007, p.&nbsp;97]]</ref> argues on [[ontology|ontological]] grounds that anything that is not a metal, is a nonmetal and that this includes semi-metals (i.e. metalloids).|group=n}} Other authors, in contrast, have suggested that classifying some elements as metalloids 'emphasizes that properties change gradually rather than abruptly as one moves across or down the periodic table'.<ref name=brown>[[#Brown2006|Brown & Holme 2006, p.&nbsp;57]]</ref> Alternatively, some periodic tables distinguish elements that are metalloids in the absence of any formal dividing line between metals and nonmetals. Metalloids are instead shown as occurring in a diagonal band<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;282]]; [[#Simple2005|Simple Memory Art c. 2005]]</ref> or diffuse region.<ref>[[#Chedd1969|Chedd 1969, pp.&nbsp;12–13]]</ref>
Some authors do not classify elements bordering the metal-nonmetal dividing line as metalloids, noting that a binary classification can facilitate the establishment of some simple rules for determining bond types between metals and nonmetals.<ref name=roher/> Metalloids are grouped instead with metals,<ref>[[#Tyler1948|Tyler 1948, p.&nbsp;105]]; [[#Reilly2002|Reilly 2002, pp.&nbsp;5–6]]</ref> regarded as nonmetals<ref>[[#Hampel1976|Hampel & Hawley 1976, p.&nbsp;174]]</ref> or treated as a sub-category of them.<ref>[[#Goodrich1844|Goodrich 1844, p.&nbsp;264]]; [[#TheChemical1897|''The Chemical News'' 1897, p.&nbsp;189]]; [[#Hampel1976|Hampel & Hawley 1976, p.&nbsp;191]]; [[#Lewis1993|Lewis 1993, p.&nbsp;835]]; [[#Hérold2006|Hérold 2006, pp.&nbsp;149–50]]</ref>{{#tag:ref|Oderberg<ref>[[#Oderberg2007|Oderberg 2007, p.&nbsp;97]]</ref> argues on [[ontology|ontological]] grounds that anything not a metal is therefore a nonmetal, and that this includes semi-metals (i.e. metalloids).|group=n}} Other authors have suggested that classifying some elements as metalloids 'emphasizes that properties change gradually rather than abruptly as one moves across or down the periodic table'.<ref name=brown>[[#Brown2006|Brown & Holme 2006, p.&nbsp;57]]</ref> Some periodic tables distinguish elements that are metalloids in the absence of any formal dividing line between metals and nonmetals. Metalloids are shown as occurring in a diagonal band<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;282]]; [[#Simple2005|Simple Memory Art c. 2005]]</ref> or diffuse region.<ref>[[#Chedd1969|Chedd 1969, pp.&nbsp;12–13]]</ref>
{{clear}}
{{clear}}


==Properties of metalloids==
==Properties of metalloids==

===Physical and chemical===
===Physical and chemical===
Metalloids are usually characterised as metallic-looking brittle solids with intermediate to relatively good electrical conductivities, and each has the electronic band structure of a semimetal or semiconductor. Chemically, they mostly behave as (weak) nonmetals, have intermediate ionization energies and electronegativity values, and have amphoteric or weakly acidic [[oxide]]s. They can also form alloys with metals. Ordinarily, most of the other physical and chemical properties of metalloids are [[Metalloid (comparison of properties with those of metals and nonmetals)|intermediate in nature]].
Metalloids are usually characterised as metallic-looking 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 have amphoteric or weakly acidic [[oxide]]s. They can form alloys with metals. Most of the other physical and chemical properties of metalloids are [[Metalloid (comparison of properties with those of metals and nonmetals)|intermediate in nature]].


===Distinctive===
===Distinctive===
Of the above properties, brittleness<ref>[[#Nickelès|Nickelès 1861, p.&nbsp;416]]; [[#United1966|US Air Force Medical Service 1966, p.&nbsp;3-3]]</ref> or semiconductivity<ref>[[#Brady|Brady, Senese & Jespersen 2009, p.&nbsp;53]]</ref> or both<ref>[[#Remy1956|Remy 1956, p.&nbsp;1]]</ref> have been cited or used as markedly distinguishing indicators of metalloid status. Metallic lustre together with very marked dualistic chemical behaviour—by way of, for example, [[amphoterism|amphoteric oxides]]—has also been cited as a benchmark.<ref>[[#Johnston1992|Johnston 1992, p.&nbsp;57]]</ref>
Brittleness,<ref>[[#Nickelès|Nickelès 1861, p.&nbsp;416]]; [[#United1966|US Air Force Medical Service 1966, p.&nbsp;3-3]]</ref> semiconductivity<ref>[[#Brady|Brady, Senese & Jespersen 2009, p.&nbsp;53]]</ref> or both<ref>[[#Remy1956|Remy 1956, p.&nbsp;1]]</ref> have been used as distinguishing indicators of metalloid status. Metallic lustre along with dualistic chemical behaviour—for example, amphoterism—has also been cited as a benchmark.<ref>[[#Johnston1992|Johnston 1992, p.&nbsp;57]]</ref>


The concepts of ''metalloid'' and ''[[semiconductor]]'' should not be confused. 'Metalloid' is chemistry-based concept referring to the physical (including electronic) and chemical properties of certain [[periodic table elements]]. 'Semiconductor' is a physics-based concept referring to the electronic properties of materials (including both elements and compounds).<ref>[[#Malerba1985|Malerba 1985, p.&nbsp;13]]</ref> Not all elements classified in the literature as metalloids display semiconductivity, although most do.<ref>[[#Rochow1966|Rochow 1966, p.&nbsp;14]]</ref>
A ''metalloid'' is a chemistry-based concept describing the physical (including electronic) and chemical properties of elements. ''[[Semiconductor]]'' is a physics-based concept referring to the electronic properties of materials (including both elements and compounds).<ref>[[#Malerba1985|Malerba 1985, p.&nbsp;13]]</ref> Not all elements classified in the literature as metalloids display semiconductivity, although most do.<ref>[[#Rochow1966|Rochow 1966, p.&nbsp;14]]</ref>


Though metalloids are all considered solid,<ref>[[#Boikess1985|Boikess & Edelson 1985, p.&nbsp;85]]</ref> and have metallic lustre, their other properties vary.<ref>[[#Aldridge1998|Aldridge 1998, p.&nbsp;290]]</ref> Given metallic character (for example) is a combination of several properties, it has been suggested that metalloid status be judged separately for each element. This could be done based on the extent to which an element exhibits properties relevant to such status.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]</ref>
Metalloids are all solid,<ref>[[#Boikess1985|Boikess & Edelson 1985, p.&nbsp;85]]</ref> and have metallic lustre, but their other properties vary.<ref>[[#Aldridge1998|Aldridge 1998, p.&nbsp;290]]</ref> Given that metallic character (for example) is a combination of several properties, it has been suggested that metalloid status be judged separately for each element. This could be done based on the extent to which an element exhibits properties relevant to such status.<ref>[[#Hawkes2001|Hawkes 2001, p.&nbsp;1686]]</ref>


===Compared to those of metals and nonmetals===
===Compared to those of metals and nonmetals===
{{Main|Metalloid (comparison of properties with those of metals and nonmetals)}}
{{Main|Metalloid (comparison of properties with those of metals and nonmetals)}}
Characteristic properties of metals, metalloids and nonmetals are summarized in the following table.<ref>[[#Kneen1972|Kneen, Rogers & Simpson, 1972, p.&nbsp;263.]] Columns 2 and 4 are sourced from this reference unless otherwise indicated.</ref> Physical properties are listed in loose order of ease of determination; chemical properties run from general to specific, and then to descriptive.
Characteristic properties of metals, metalloids and nonmetals are summarized in the table.<ref>[[#Kneen1972|Kneen, Rogers & Simpson, 1972, p.&nbsp;263.]] Columns 2 and 4 are sourced from this reference unless otherwise indicated.</ref> Physical properties are listed in order of ease of determination; chemical properties run from general to specific, and then to descriptive.


{|class="wikitable"
{|class="wikitable"
Line 166: Line 168:
| scope="row"| [[Electrical conductivity]]
| scope="row"| [[Electrical conductivity]]
| good to high{{#tag:ref|Metals have electrical conductivity values of from 6.9 × 10<sup>3</sup>&nbsp;S•cm<sup>−1</sup> for [[manganese]] to 6.3 × 10<sup>5</sup> for [[silver]].<ref>[[#Desai1984|Desai, James & Ho 1984, p.&nbsp;1160]]; [[#Matula1979|Matula 1979, p.&nbsp;1260]]</ref>|group=n}}
| good to high{{#tag:ref|Metals have electrical conductivity values of from 6.9 × 10<sup>3</sup>&nbsp;S•cm<sup>−1</sup> for [[manganese]] to 6.3 × 10<sup>5</sup> for [[silver]].<ref>[[#Desai1984|Desai, James & Ho 1984, p.&nbsp;1160]]; [[#Matula1979|Matula 1979, p.&nbsp;1260]]</ref>|group=n}}
| intermediate<ref>[[#Choppin1972|Choppin & Johnsen 1972, p.&nbsp;351]]</ref> to good{{#tag:ref|Metalloids have electrical conductivity values of from 1.5 × 10<sup>−6</sup>&nbsp;S•cm<sup>−1</sup> for boron to 3.9 × 10<sup>4</sup> for [[arsenic]].<ref>[[#Schaefer1968|Schaefer 1968, p.&nbsp;76]]; [[#Carapella1968|Carapella 1968, p.&nbsp;30]]</ref> If [[selenium]] is included as a metalloid the applicable conductivity range would start from ~10<sup>−9</sup> to 10<sup>−12</sup>&nbsp;S•cm<sup>−1</sup>.<ref name="Kozyrev">[[#Kozyrev1959|Kozyrev 1959, p.&nbsp;104]]; [[#Chizhikov1968|Chizhikov & Shchastlivyi 1968, p.&nbsp;25]];
| intermediate<ref>[[#Choppin1972|Choppin & Johnsen 1972, p.&nbsp;351]]</ref> to good{{#tag:ref|Metalloids have electrical conductivity values of from 1.5 × 10<sup>−6</sup>&nbsp;S•cm<sup>−1</sup> for boron to 3.9 × 10<sup>4</sup> for arsenic.<ref>[[#Schaefer1968|Schaefer 1968, p.&nbsp;76]]; [[#Carapella1968|Carapella 1968, p.&nbsp;30]]</ref> If selenium is included as a metalloid the applicable conductivity range would start from ~10<sup>−9</sup> to 10<sup>−12</sup>&nbsp;S•cm<sup>−1</sup>.<ref name="Kozyrev">[[#Kozyrev1959|Kozyrev 1959, p.&nbsp;104]]; [[#Chizhikov1968|Chizhikov & Shchastlivyi 1968, p.&nbsp;25]];
[[#Glazov1969|Glazov, Chizhevskaya & Glagoleva 1969, p.&nbsp;86]]</ref>|group=n}}
[[#Glazov1969|Glazov, Chizhevskaya & Glagoleva 1969, p.&nbsp;86]]</ref>|group=n}}
| poor to good{{#tag:ref|Nonmetals have electrical conductivity values of from ~10<sup>−18</sup>&nbsp;S•cm<sup>−1</sup> for the elemental gases to 3 × 10<sup>4</sup> in graphite.<ref>[[#Bogoroditskii1967|Bogoroditskii & Pasynkov 1967, p.&nbsp;77]]; [[#Jenkins1976|Jenkins & Kawamura 1976, p.&nbsp;88]]</ref>|group=n}}
| poor to good{{#tag:ref|Nonmetals have electrical conductivity values of from ~10<sup>−18</sup>&nbsp;S•cm<sup>−1</sup> for the elemental gases to 3 × 10<sup>4</sup> in graphite.<ref>[[#Bogoroditskii1967|Bogoroditskii & Pasynkov 1967, p.&nbsp;77]]; [[#Jenkins1976|Jenkins & Kawamura 1976, p.&nbsp;88]]</ref>|group=n}}
Line 192: Line 194:
| scope="row" |[[Electronegativity]]
| scope="row" |[[Electronegativity]]
| usually low
| usually low
| have electronegativity values close to 2<ref>[[#Pauling1988|Pauling 1988, p.&nbsp;183]]</ref> (revised [[Linus Pauling|Pauling]] scale) or within the narrow range of 1.9–2.2 (Allen scale)<ref name="Mann">[[#Mann2000|Mann et al. 2000, p.&nbsp;2783]]</ref>{{#tag:ref|Chedd<ref>[[#Chedd1969|Chedd 1969, pp.&nbsp;24–5]]</ref> defines metalloids as having electronegativity values of 1.8 to 2.2 [[Allred-Rochow scale|(Allred-Rochow scale).]] He included [[boron]], [[silicon]], [[germanium]], [[arsenic]], [[antimony]], [[tellurium]], [[polonium]] and [[astatine]] in this category. In reviewing Chedd's work, Adler<ref>[[#Adler1969|Adler 1969, pp.&nbsp;18–19]]</ref> described this choice as arbitrary, given other elements have electronegativities in this range, including [[copper]], [[silver]], phosphorus, [[mercury (element)|mercury]] and [[bismuth]]. He went on to suggest defining a metalloid simply as, 'a semiconductor or semimetal' and 'to have included the interesting materials bismuth and [[selenium]] in the book'.|group=n}}
| have electronegativity values close to 2<ref>[[#Pauling1988|Pauling 1988, p.&nbsp;183]]</ref> (revised Pauling scale) or within the range of 1.9–2.2 (Allen scale)<ref name="Mann">[[#Mann2000|Mann et al. 2000, p.&nbsp;2783]]</ref>{{#tag:ref|Chedd<ref>[[#Chedd1969|Chedd 1969, pp.&nbsp;24–5]]</ref> defines metalloids as having electronegativity values of 1.8 to 2.2 [[Allred-Rochow scale|(Allred-Rochow scale).]] He included boron, silicon, germanium, arsenic, antimony, tellurium, polonium and [[astatine]] in this category. In reviewing Chedd's work, Adler<ref>[[#Adler1969|Adler 1969, pp.&nbsp;18–19]]</ref> described this choice as arbitrary, as other elements with electronegativities in this range include [[copper]], silver, phosphorus, mercury and bismuth. He went on to suggest defining a metalloid as 'a semiconductor or semimetal' and included bismuth and selenium in his book.|group=n}}
| high
| high
|- valign=top
|- valign=top
Line 205: Line 207:
| acidic
| acidic
|}
|}
Of the above properties, three (form, appearance, and when mixed with metals) are shared with metals and two (elasticity and general chemical behaviour) with nonmetals. The other five (electrical conductivity, band structure, ionization energy, electronegativity, and oxides) are intermediate in nature.
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.


==Typical or shared applications==
==Typical or shared applications==
:''The focus of this section is on the recognised metalloids. Although elements less commonly recognised as metalloids are ordinarly classified as either metals or nonmetals, some of these are included here for comparative purposes, as and where appropriate. For prevalent and speciality applications of individual metalloids see the main article for each element. ''
<!-- :''The focus of this section is on the recognised metalloids. Although elements less commonly recognised as metalloids are ordinarly classified as either metals or nonmetals, some of these are included here for comparative purposes, as and where appropriate. For prevalent and speciality applications of individual metalloids see the main article for each element. '' -->


Metalloids are too brittle to have any structural uses in their pure forms.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;421, 423]]; [[#Gray2009|Gray 2009, p.&nbsp;23]]</ref> Typical or shared applications have instead encompassed their presence in, or specific uses as, alloying components;
Metalloids are too brittle to have any structural uses in their pure forms.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;421, 423]]; [[#Gray2009|Gray 2009, p.&nbsp;23]]</ref> Theyand their compounds are used as alloying components, biological agents (toxicological, nutritional and medicinal), flame retardants, glasses (oxide and metallic), optical storage media and electronics, and semiconductors.{{#tag:ref|Olmsted and Williams<ref>[[#Olmsted1997|Olmsted & Williams 1997, p.&nbsp;975]]</ref> 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'.|group=n}}
biological agents (toxicological; nutritional; medicinal); flame retardants; glasses (oxide and metallic); optical storage media and electronics; and
semiconductors.{{#tag:ref|Olmsted and Williams<ref>[[#Olmsted1997|Olmsted & Williams 1997, p.&nbsp;975]]</ref> 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'.|group=n}}


===Alloys===
===Alloys===
[[File:Copper germanium.jpg|thumb|right|Copper-germanium master alloy{{#tag:ref|A master alloy is an alloy usually comprised of a [[base metal]] (such as aluminium, nickel or copper) and a relatively high percentage of one or two other elements (only Ge in this example), which is added to a melt to raise the percentage of a desired constituent (i.e. Ge) in a final alloy.<ref>[[#Seybolt1953|Seybolt & Burke 1953, p.&nbsp;169]]</ref> They are usually available commercially, or can be made to order.<ref>[[#Isbell1998|Isbell 1998, p.&nbsp;106]]</ref>|group=n}} pellets, likely ~84% Cu; 16% Ge.<ref name="Russell2005401"/> When combined with silver the result is a [[argentium sterling silver|tarnish resistant]] [[sterling silver]]. Also present are two silver pellets.|alt=Several dozen metallic pellets, reddish-brown in colour with a highly polished appearance, as if they had a cellophane coating.]]
[[File:Copper germanium.jpg|thumb|right|Copper-germanium alloy pellets, likely ~84% Cu; 16% Ge.<ref name="Russell2005401"/> When combined with silver the result is a [[argentium sterling silver|tarnish resistant sterling silver]]. Also present are two silver pellets.|alt=Several dozen metallic pellets, reddish-brown with a highly polished appearance, as if they had a cellophane coating.]]
Writing early in the history of [[intermetallic|intermetallic compounds]], the British metallurgist Cecil Desch observed that 'certain non-metallic elements are capable of forming compounds of distinctly metallic character with metals, and these elements may therefore enter into the composition of alloys'. He associated silicon, arsenic and tellurium—in particular—with the alloy-forming elements.<ref>[[#Desch1914|Desch 1914, p.&nbsp;86]]</ref> More specifically, compounds of silicon, germanium, arsenic and antimony with the poor metals, it has been suggested, 'are probably best classed as alloys.'<ref>[[#Phillips1965|Phillips & Williams 1965, p.&nbsp;620]]</ref>
Writing early in the history of [[intermetallic|intermetallic compounds]], the British metallurgist Cecil Desch observed that 'certain non-metallic elements are capable of forming compounds of distinctly metallic character with metals, and these elements may therefore enter into the composition of alloys'. He associated silicon, arsenic and tellurium—in particular—with the alloy-forming elements.<ref>[[#Desch1914|Desch 1914, p.&nbsp;86]]</ref> Compounds of silicon, germanium, arsenic and antimony with the poor metals, it has been suggested, 'are probably best classed as alloys.'<ref>[[#Phillips1965|Phillips & Williams 1965, p.&nbsp;620]]</ref>


In terms of individual alloy types, those with [[transition metals]] are well-represented. Boron can form intermetallic compounds and alloys with such metals, of the composition M<sub>''n''</sub>B, if ''n'' > 2.<ref>[[#Vanderput1998|Van der Put 1998, p.&nbsp;123]]</ref> Ferroboron (15% boron) is used to introduce boron into steel; nickel-boron alloys are ingredients in welding alloys and face-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 [[Group 11 element|coinage metals]].<ref>[[#Klug1958|Klug & Brasted 1958, p.&nbsp;199]]</ref> Arsenic can form alloys with metals, including platinum and copper;<ref>[[#Good1813|Good et al. 1813]]</ref> it is also added to copper and copper alloys to improve corrosion resistance<ref>[[#Sequeira|Sequeira 2011, p.&nbsp;776]]</ref> and appears to confer the same benefit when added to magnesium.<ref>[[#Gary|Gary 2013]]</ref> Antimony is well known as an alloy former, including with the coinage metals. Its alloys are exemplified by [[pewter]] (a tin alloy with up to 20% antimony) and [[type metal]] (a lead alloy with up to 25% antimony).<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;423–4; 405–6]]</ref> Tellurium readily forms alloys with iron, in the form of ferrotellurium (50–58% tellurium), and with copper, in the form of copper tellurium (40–50% tellurium).<ref>[[#Davidson1973|Davidson & Lakin 1973, p.&nbsp;627]]</ref> Ferrotellurium, in particular, is used as a stabilizer for carbon in steel casting.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;589]]</ref> Of the non-metallic elements that are less often recognised as metalloids, selenium—in the form of ferroselenium (50–58% selenium)—is used to improve the workability and machinability of stainless steels.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;749]]; [[#Schwartz2002|Schwartz 2002, p.&nbsp;679]]</ref>
Alloys with [[transition metals]] are well-represented. Boron can form intermetallic compounds and alloys with such metals, of the composition M<sub>''n''</sub>B, if ''n'' > 2.<ref>[[#Vanderput1998|Van der Put 1998, p.&nbsp;123]]</ref> Ferroboron (15% boron) is used to introduce boron into steel; nickel-boron alloys are ingredients in welding alloys and face-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 [[Group 11 element|coinage metals]].<ref>[[#Klug1958|Klug & Brasted 1958, p.&nbsp;199]]</ref> Arsenic can form alloys with metals, including platinum and copper;<ref>[[#Good1813|Good et al. 1813]]</ref> it is also added to copper and its alloys to improve corrosion resistance<ref>[[#Sequeira|Sequeira 2011, p.&nbsp;776]]</ref> and appears to confer the same benefit when added to magnesium.<ref>[[#Gary|Gary 2013]]</ref> Antimony is well known as an alloy former, including with the coinage metals. Its alloys are exemplified by [[pewter]] (a tin alloy with up to 20% antimony) and [[type metal]] (a lead alloy with up to 25% antimony).<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;423–4; 405–6]]</ref> Tellurium readily alloys with iron, as ferrotellurium (50–58% tellurium), and with copper, in the form of copper tellurium (40–50% tellurium).<ref>[[#Davidson1973|Davidson & Lakin 1973, p.&nbsp;627]]</ref> Ferrotellurium is used as a stabilizer for carbon in steel casting.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;589]]</ref> Of the non-metallic elements that are less often recognised as metalloids, selenium—in the form of ferroselenium (50–58% selenium)—is used to improve the workability of stainless steels.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;749]]; [[#Schwartz2002|Schwartz 2002, p.&nbsp;679]]</ref>


===Biological agents===
===Biological agents===
[[File:Arsenic trioxide.jpg|thumb|left|[[Arsenic trioxide]] or ''white arsenic,'' one of the most toxic and prevalent forms of arsenic. The antileukaemic properties of white arsenic were first reported in 1878.<ref>[[#Antman|Antman 2001]]</ref>|alt=A clear glass dish on which is a small mound of a white crystalline powder.]] All six of the elements commonly recognised as metalloids have toxic, dietary or medicinal properties<ref>[[#Řezanka|Řezanka & Sigler 2008]]; [[#Sekhon|Sekhon 2012]]</ref> to varying degrees. Two (arsenic and antimony) are notably toxic; two (boron and silicon) or three (arsenic) are essential trace elements; and four (boron, silicon, arsenic and antimony) have medical applications (with germanium and tellurium being thought to have potential)
[[File:Arsenic trioxide.jpg|thumb|left|[[Arsenic trioxide]] or ''white arsenic,'' one of the most toxic and prevalent forms of arsenic. The antileukaemic properties of white arsenic were first reported in 1878.<ref>[[#Antman|Antman 2001]]</ref>|alt=A clear glass dish on which is a small mound of a white crystalline powder.]] All six of the elements commonly recognised as metalloids have toxic, dietary or medicinal properties<ref>[[#Řezanka|Řezanka & Sigler 2008]]; [[#Sekhon|Sekhon 2012]]</ref> to varying degrees. Arsenic and antimony are toxic; boron, silicon, and possibly arsenic, are essential trace elements. Boron, silicon, arsenic and antimony have medical applications (with germanium and tellurium being thought to have potential).


Boron is used in insecticides<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;67]]</ref> and herbicides.<ref>[[#Zhang2008|Zhang et al. 2008, p.&nbsp;360]]</ref> Even so, it is an essential trace element.<ref name=SLH>[[#SLH|Science Learning Hub 2009]]</ref> As [[boric acid]], it also has antiseptic, antifungal, and antiviral properties.
Boron is used in insecticides<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;67]]</ref> and herbicides.<ref>[[#Zhang2008|Zhang et al. 2008, p.&nbsp;360]]</ref> It is an essential trace element.<ref name=SLH>[[#SLH|Science Learning Hub 2009]]</ref> As [[boric acid]], it has antiseptic, antifungal, and antiviral properties.


Silicon is not toxic although it makes up the central component of [[silatrane]], a highly toxic rodenticide.<ref>[[#Büchel|Büchel 1983, p.&nbsp;226]]</ref> Long-term inhalation of silica dust also causes [[silicosis]], a fatal disease of the lungs. Silicon is, however, an essential trace element.<ref name=SLH/> It can also be applied to badly burned patients, in the form a [[silicone]] gel, to reduce scarring.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;391]]</ref>
Silicon is not toxic. It is present in [[silatrane]], a highly toxic rodenticide.<ref>[[#Büchel|Büchel 1983, p.&nbsp;226]]</ref> Long-term inhalation of silica dust causes [[silicosis]], a fatal disease of the lungs. Silicon is an essential trace element.<ref name=SLH/> [[Silicone]] gel can be applied to badly burned patients to reduce scarring.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;391]]</ref>


Salts of germanium are potentially harmful to humans and animals if ingested on a prolonged basis.<ref>[[#Schauss1991|Schauss 1991]]; [[#Tao1997|Tao & Bolger 1997]]</ref> It is not an essential trace element. Although interest in the pharmacological actions of germanium compounds is ongoing there is (as yet) no licensed medicine.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;450]]; [[#EVM|EVM 2003, pp.&nbsp;197‒202]]</ref>
Salts of germanium are potentially harmful to humans and animals if ingested on a prolonged basis.<ref>[[#Schauss1991|Schauss 1991]]; [[#Tao1997|Tao & Bolger 1997]]</ref> It is not an essential trace element. Although interest in the pharmacological actions of germanium compounds is ongoing there is (as yet) no licensed medicine.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;450]]; [[#EVM|EVM 2003, pp.&nbsp;197‒202]]</ref>


Arsenic is notoriously poisonous. Even so it may possibly be an essential element in ultratrace amounts.<ref name=Neilsen>[[#Nielsen|Nielsen 1998]]</ref> Arsenic has been used as a pharmaceutical agent since antiquity and notably for the treatment of [[syphilis]] prior to the development of antibiotics.<ref name=Jaouen>[[#Jaouen|Jaouen & Gibaud 2010]]</ref> Arsenic is also a component of [[melarsoprol]], a medicinal drug used in the treatment of [[human African trypanosomiasis]] or sleeping sickness. In 2003, arsenic trioxide (under the trade name [[Trisenox]]) was re-introduced for the treatment of [[acute promyelocytic leukaemia]], a cancer of the blood and bone marrow.<ref name=Jaouen/>
Arsenic is notoriously poisonous and may possibly also be an essential element in ultratrace amounts.<ref name=Neilsen>[[#Nielsen|Nielsen 1998]]</ref> It has been used as a pharmaceutical agent since antiquity for the treatment of [[syphilis]] prior to the development of antibiotics.<ref name=Jaouen>[[#Jaouen|Jaouen & Gibaud 2010]]</ref> Arsenic is also a component of [[melarsoprol]], a medicinal drug used in the treatment of [[human African trypanosomiasis]] or sleeping sickness. In 2003, arsenic trioxide (under the trade name [[Trisenox]]) was re-introduced for the treatment of [[acute promyelocytic leukaemia]], a cancer of the blood and bone marrow.<ref name=Jaouen/>


While metallic animony is relatively non-toxic, most antimony compounds are poisonous.<ref>[[#Stevens1990|Stevens & Klarner, p.&nbsp;205]]</ref> It is not an essential element. Compounds of antimony are however used as antiprotozoan drugs, and in some veterinary preparations.
Metallic antimony is relatively non-toxic; most antimony compounds are poisonous.<ref>[[#Stevens1990|Stevens & Klarner, p.&nbsp;205]]</ref> It is not an essential element. Compounds of antimony are used as antiprotozoan drugs, and in some veterinary preparations.


Tellurium is not particularly noted for its toxicity although as little as two grams of sodium tellurate, if administered, can be lethal.<ref>[[#Keall1946|Keall, Martin and Tunbridge 1946]]</ref> As well, people exposed to small amounts of airborne tellurium exude a foul and persistent garlic-like odour.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;426]]</ref> It is not an essential element. Tellurium dioxide has been used to treat [[seborrhoeic dermatitis]]; other tellurium compounds were used as antimicrobial agents before the development of antibiotics.<ref>[[#Oldfield1974|Oldfield et al. 1974, p.&nbsp;65]]; [[#Turner2011|Turner 2011]]</ref> Ironically, such compounds may have the potential to act as substitutes for antibiotics that have become ineffective due to increasing bacterial resistance.<ref>[[#Ba|Ba et al. 2010]]; [[#Daniel-Hoffmann|Daniel-Hoffmann, Sredni & Nitzan 2012]]; [[#Molina-Quiroz|Molina-Quiroz et al. 2012]]</ref>
Tellurium is not considered particularly toxic although as little as two grams of sodium tellurate, if administered, can be lethal.<ref>[[#Keall1946|Keall, Martin and Tunbridge 1946]]</ref> People exposed to small amounts of airborne tellurium exude a foul and persistent garlic-like odour.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;426]]</ref> It is not an essential element. Tellurium dioxide has been used to treat [[seborrhoeic dermatitis]]; other tellurium compounds were used as antimicrobial agents before the development of antibiotics.<ref>[[#Oldfield1974|Oldfield et al. 1974, p.&nbsp;65]]; [[#Turner2011|Turner 2011]]</ref> Such compounds may have the potential to act as substitutes for antibiotics that have become ineffective due to increasing bacterial resistance.<ref>[[#Ba|Ba et al. 2010]]; [[#Daniel-Hoffmann|Daniel-Hoffmann, Sredni & Nitzan 2012]]; [[#Molina-Quiroz|Molina-Quiroz et al. 2012]]</ref>


Of the elements less often recognised as metalloids, beryllium and lead are noted for their toxicity, with [[lead arsenate]], in particular, having been extensively used as an insecticide.<ref>[[#Peryea|Peryea 1998]]</ref> Sulfur, too, is one of the oldest fungicides and pesticides. Phosphorus, sulfur, zinc and iodine are essential nutrients, as are possibly aluminium, tin and lead.<ref name=Neilsen/> Sulfur, gallium, iodine and bismuth, have medicinal applications. Sulfur, in the form of [[Sulfonamide (medicine)|sulfonamide drugs]], is still widely used for conditions such as acne and urinary tract infections.<ref>[[#Hager|Hager 2006, p.&nbsp;299]]</ref> [[Gallium nitrate]] is used to treat the side effects of cancer;<ref>[[#Apseloff|Apseloff 1999]]</ref> and gallium citrate, a [[radiopharmaceutical]], is used to facilitate body imaging in areas of inflammation, such as infection, and areas of rapid cell division.<ref>[[#Trivedi|Trivedi, Yung & Katz 2013, p.&nbsp;209]]</ref> Iodine is used as a disinfectant in various forms. Bismuth is an ingredient in some antibacterial pharmaceuticals.<ref>[[#Thomas|Thomas, Bialek & Hensel 2013, p.&nbsp;1]]</ref>
Of the elements less often recognised as metalloids, beryllium and lead are noted for their toxicity, with [[lead arsenate]] having been extensively used as an insecticide.<ref>[[#Peryea|Peryea 1998]]</ref> Sulfur is one of the oldest fungicides and pesticides. Phosphorus, sulfur, zinc and iodine are essential nutrients, as are possibly aluminium, tin and lead.<ref name=Neilsen/> Sulfur, gallium, iodine and bismuth have medicinal applications. Sulfur is a constituent of [[Sulfonamide (medicine)|sulfonamide drugs]], still widely used for conditions such as acne and urinary tract infections.<ref>[[#Hager|Hager 2006, p.&nbsp;299]]</ref> [[Gallium nitrate]] is used to treat the side effects of cancer;<ref>[[#Apseloff|Apseloff 1999]]</ref> and gallium citrate, a [[radiopharmaceutical]], is used to facilitate body imaging in areas of inflammation, such as infection, and areas of rapid cell division.<ref>[[#Trivedi|Trivedi, Yung & Katz 2013, p.&nbsp;209]]</ref> Iodine is used as a disinfectant in various forms. Bismuth is an ingredient in some antibacterial pharmaceuticals.<ref>[[#Thomas|Thomas, Bialek & Hensel 2013, p.&nbsp;1]]</ref>


===Flame retardants===
===Flame retardants===
Compounds of boron, silicon, arsenic and antimony have found or continue to find uses as flame retardants. Boron, in the form of [[borax]], has been used as a textile flame retardant since at least the 18th century.<ref>[[#LeBras|Le Bras, Wilkie & Bourbigot 2005, p.&nbsp;v]]</ref> Silicon compound additives such as [[silicone]]s, [[silane]]s, [[silsesquioxane]], [[silica]] and [[silicate]]s, some of which were developed as alternatives to more toxic [[halogenation|halogenated]] products, can considerably improve the flame retardancy of plastic materials.<ref>[[#Wilkie|Wilkie & Morgan 2009, p.&nbsp;187]]</ref>
Compounds of boron, silicon, arsenic and antimony have been used as flame retardants. Boron, in the form of [[borax]], has been used as a textile flame retardant since at least the 18th century.<ref>[[#LeBras|Le Bras, Wilkie & Bourbigot 2005, p.&nbsp;v]]</ref> Silicon compounds such as silicones, [[silane]]s, [[silsesquioxane]], [[silica]] and [[silicate]]s, some of which were developed as alternatives to more toxic [[halogenation|halogenated]] products, can considerably improve the flame retardancy of plastic materials.<ref>[[#Wilkie|Wilkie & Morgan 2009, p.&nbsp;187]]</ref>
Arsenic compounds in the form of [[sodium arsenite]] or [[sodium arsenate]] are effective flame retardants for wood but were less frequently used due to their toxicity.<ref>[[#Locke1956|Locke et al. 1956, p.&nbsp;88]]</ref> Antimony, as antimony trioxide, finds its greatest use as a flame retardant additive.<ref>[[#Carlin|Carlin 2011, p.&nbsp;6.2]]</ref> Aluminium, in the form of its [[aluminium hydroxide|hydroxide]], has been used as a wood-fibre, rubber, plastic and textile flame retardant since the 1890s.<ref>[[#Evans|Evans 1993, pp.&nbsp; 257–8]]</ref> Barring aluminium hydroxide, use of phosphorus based flame-retardants in the form of, for example, [[organophosphates]], now exceeds that of any of the other main retardant types, that is, those employing boron, antimony or halogenated hydrocarbon compounds.<ref>[[#Corbridge|Corbridge 2013, p.&nbsp;1149]]</ref>
Arsenic compounds in the form of [[sodium arsenite]] or [[sodium arsenate]] are effective flame retardants for wood but were less frequently used due to their toxicity.<ref>[[#Locke1956|Locke et al. 1956, p.&nbsp;88]]</ref> Antimony, as antimony trioxide, is a flame retardant.<ref>[[#Carlin|Carlin 2011, p.&nbsp;6.2]]</ref> [[Aluminium hydroxide]], has been used as a wood-fibre, rubber, plastic and textile flame retardant since the 1890s.<ref>[[#Evans|Evans 1993, pp.&nbsp; 257–8]]</ref> Barring aluminium hydroxide, use of phosphorus based flame-retardants in the form of, for example, [[organophosphates]], now exceeds that of any of the other main retardant types, which employ boron, antimony or halogenated hydrocarbon compounds.<ref>[[#Corbridge|Corbridge 2013, p.&nbsp;1149]]</ref>


===Glass formation===
===Glass formation===
[[File:Fibreoptic4.jpg|thumb|right|[[Optical fiber|Fibre-optic strands]], usually made of pure silicon dioxide glass, but with additives such as boron trioxide or germanium dioxide for increased sensitivity|alt=A bunch of pale yellow semi-transparent thin strands, with bright points of white light at their tips to one end.]]
[[File:Fibreoptic4.jpg|thumb|right|[[Optical fiber|Optical fibres]], usually made of pure silicon dioxide glass, with additives such as boron trioxide or germanium dioxide for increased sensitivity|alt=A bunch of pale yellow semi-transparent thin strands, with bright points of white light at their tips.]]
The oxides [[Boron trioxide|B<sub>2</sub>O<sub>3</sub>]], [[Silicon dioxide|SiO<sub>2</sub>]], [[Germanium dioxide|GeO<sub>2</sub>]], [[Arsenic trioxide|As<sub>2</sub>O<sub>3</sub>]] and [[Antimony trioxide|Sb<sub>2</sub>O<sub>3</sub>]] readily form glasses. [[Tellurium dioxide|TeO<sub>2</sub>]] forms a glass but this requires a 'heroic quench rate' or the addition of an impurity; otherwise the crystalline form results.<ref>[[#Kaminow2002|Kaminow & Li 2002, p.&nbsp;118]]</ref> These compounds have found or continue to find practical uses in chemical, domestic and industrial glassware<ref>[[#Deming1925|Deming 1925]], pp.&nbsp;330 (As<sub>2</sub>O<sub>3</sub>), 418 (B<sub>2</sub>O<sub>3</sub>; SiO<sub>2</sub>; Sb<sub>2</sub>O<sub>3</sub>); [[#Witt1968|Witt & Gatos 1968, p.&nbsp;242]] (GeO<sub>2</sub>)</ref> and optics.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;421]] (GeO<sub>2</sub>); [[#Rothenberg1976|Rothenberg 1976, 56, 118–19]] (TeO<sub>2</sub>)</ref> Boron trioxide is used as a glass fibre additive;<ref>[[#Geckeler1987|Geckeler 1987, p.&nbsp;20]]</ref> it is also a component of [[borosilicate glass]], which is widely used for laboratory glassware, as well as in home ovenware.<ref>[[#Kreith2005|Kreith & Goswami 2005, p.&nbsp;12–109]]</ref> Silicon dioxide forms the basis of ordinary domestic glassware.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;397]]</ref> Germanium dioxide is used as glass fibre additive, as well as in infrared optical systems.<ref>[[#Butterman2005|Butterman & Jorgenson 2005, pp.&nbsp;9–10]]</ref> Arsenic trioxide is used in the glass industry as a decolourizing and fining agent, as is antimony trioxide.<ref>[[#Butterman2004|Butterman & Carlin 2004, p.&nbsp;22]]; [[#Russell2005|Russell & Lee 2005, p.&nbsp;422]]</ref> Tellurium dioxide finds application in laser and [[nonlinear optics]].<ref>[[#Träger2007|Träger 2007, pp.&nbsp;438, 958]]; [[#Eranna2011|Eranna 2011, p.&nbsp;98]]</ref>
The oxides [[Boron trioxide|B<sub>2</sub>O<sub>3</sub>]], [[Silicon dioxide|SiO<sub>2</sub>]], [[Germanium dioxide|GeO<sub>2</sub>]], [[Arsenic trioxide|As<sub>2</sub>O<sub>3</sub>]] and [[Antimony trioxide|Sb<sub>2</sub>O<sub>3</sub>]] readily form glasses. [[Tellurium dioxide|TeO<sub>2</sub>]] forms a glass but this requires a 'heroic quench rate' or the addition of an impurity; otherwise the crystalline form results.<ref>[[#Kaminow2002|Kaminow & Li 2002, p.&nbsp;118]]</ref> These compounds are used in chemical, domestic and industrial glassware<ref>[[#Deming1925|Deming 1925]], pp.&nbsp;330 (As<sub>2</sub>O<sub>3</sub>), 418 (B<sub>2</sub>O<sub>3</sub>; SiO<sub>2</sub>; Sb<sub>2</sub>O<sub>3</sub>); [[#Witt1968|Witt & Gatos 1968, p.&nbsp;242]] (GeO<sub>2</sub>)</ref> and optics.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;421]] (GeO<sub>2</sub>); [[#Rothenberg1976|Rothenberg 1976, 56, 118–19]] (TeO<sub>2</sub>)</ref> Boron trioxide is used as a glass fibre additive;<ref>[[#Geckeler1987|Geckeler 1987, p.&nbsp;20]]</ref> it is also a component of [[borosilicate glass]], which is widely used for laboratory glassware and domestic ovenware.<ref>[[#Kreith2005|Kreith & Goswami 2005, p.&nbsp;12–109]]</ref> Silicon dioxide forms most ordinary glassware.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;397]]</ref> Germanium dioxide is used as a glass fibre additive, as well as in infrared optical systems.<ref>[[#Butterman2005|Butterman & Jorgenson 2005, pp.&nbsp;9–10]]</ref> Arsenic trioxide is used in the glass industry as a decolourizing and fining agent, as is antimony trioxide.<ref>[[#Butterman2004|Butterman & Carlin 2004, p.&nbsp;22]]; [[#Russell2005|Russell & Lee 2005, p.&nbsp;422]]</ref> Tellurium dioxide finds application in laser and [[nonlinear optics]].<ref>[[#Träger2007|Träger 2007, pp.&nbsp;438, 958]]; [[#Eranna2011|Eranna 2011, p.&nbsp;98]]</ref>


[[Amorphous]] [[metallic glass]]es are generally most easily prepared if one of the components is a metalloid or 'near metalloid' such as boron, carbon, silicon, phosphorus or germanium.<ref>[[#Rao2002|Rao 2002, p.&nbsp;552]]; [[#Loffler|Löffler, Kündig & Dalla Torre 2007, p.&nbsp;17-11]]</ref>{{#tag:ref|Research published in 2012 suggests that metal-metalloid glasses can be characterised by an interconnected atomic packing scheme in which metallic and covalent bonding structures coexist.<ref>[[#Guan|Guan et al. 2012]]; [[#World|WPI-AIM 2012]]</ref>|group=n}} Aside from thin films deposited at very low temperatures, the first known metallic glass was a metal-metalloid alloy of composition Au<sub>75</sub>Si<sub>25</sub> reported in 1960.<ref>[[#Klement|Klement, Willens & Duwez 1960]]; [[#Wanga|Wanga, Dongb & Shek 2004, p.&nbsp;45]]</ref> A metallic glass having a strength and toughness not previously seen in any other material, of composition Pd<sub>82.5</sub>P<sub>6</sub>Si<sub>9.5</sub>Ge<sub>2</sub>, was reported in 2011.<ref>[[#Demetriou|Demetriou et al 2011]]; [[#Oliwenstein|Oliwenstein 2011]]</ref>
[[Amorphous]] [[metallic glass]]es are generally most easily prepared if one of the components is a metalloid or 'near metalloid' such as boron, carbon, silicon, phosphorus or germanium.<ref>[[#Rao2002|Rao 2002, p.&nbsp;552]]; [[#Loffler|Löffler, Kündig & Dalla Torre 2007, p.&nbsp;17-11]]</ref>{{#tag:ref|Research published in 2012 suggests that metal-metalloid glasses can be characterised by an interconnected atomic packing scheme in which metallic and covalent bonding structures coexist.<ref>[[#Guan|Guan et al. 2012]]; [[#World|WPI-AIM 2012]]</ref>|group=n}} Aside from thin films deposited at very low temperatures, the first known metallic glass was a metal-metalloid alloy of composition Au<sub>75</sub>Si<sub>25</sub> reported in 1960.<ref>[[#Klement|Klement, Willens & Duwez 1960]]; [[#Wanga|Wanga, Dongb & Shek 2004, p.&nbsp;45]]</ref> A metallic glass having a strength and toughness not previously seen, of composition Pd<sub>82.5</sub>P<sub>6</sub>Si<sub>9.5</sub>Ge<sub>2</sub>, was reported in 2011.<ref>[[#Demetriou|Demetriou et al 2011]]; [[#Oliwenstein|Oliwenstein 2011]]</ref>


Phosphorus, selenium and lead, which are elements less often recognised as metalloids, are also used in glasses. [[Phosphate glass]] has a substrate of phosphorus pentoxide (P<sub>2</sub>O<sub>5</sub>), rather than the silica (SiO<sub>2</sub>) of convention silicate glasses and is used, for example, to make [[sodium lamp]]s.<ref>[[#Karabulut|Karabulut et al. 2001, p.&nbsp;15]]; [[#Haynes|Haynes 2012, p.&nbsp;4-26]]</ref> Selenium compounds can be used both as decolourising agents and to add a red colour to glass.<ref>[[#Schwartz2002|Schwartz 2002, pp.&nbsp;679–680]]</ref> 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.<ref>[[#Carter|Carter & Norton 2013, p.&nbsp;403]]</ref> Lead-based glasses have also been extensively used in or as electronics components; enameling; 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.<ref>[[#Maeder|Maeder 2013, pp.&nbsp;3, 9–11]]</ref>
Phosphorus, selenium and lead, which are less often recognised as metalloids, are also used in glasses. [[Phosphate glass]] has a substrate of phosphorus pentoxide (P<sub>2</sub>O<sub>5</sub>), rather than the silica (SiO<sub>2</sub>) of convention silicate glasses and is used, for example, to make [[sodium lamp]]s.<ref>[[#Karabulut|Karabulut et al. 2001, p.&nbsp;15]]; [[#Haynes|Haynes 2012, p.&nbsp;4-26]]</ref> Selenium compounds can be used both as decolourising agents and to add a red colour to glass.<ref>[[#Schwartz2002|Schwartz 2002, pp.&nbsp;679–680]]</ref> 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.<ref>[[#Carter|Carter & Norton 2013, p.&nbsp;403]]</ref> Lead-based glasses have also been extensively used in electronics 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.<ref>[[#Maeder|Maeder 2013, pp.&nbsp;3, 9–11]]</ref>


===Optical storage===
===Optical storage===
Varying compositions of GeSbTe ("[[GeSbTe|GST alloys]]") and [[AgInSbTe|Ag- and In- doped Sb<sub>2</sub>Te]] ("AIST alloys"), being examples of [[phase-change material]]s, are widely used in rewritable [[optical disc|optical disks]] and [[phase-change memory]] devices. By applying heat, they can be switched between amorphous (glassy) and [[crystalline]] states, thereby changing their optical and electrical properties and allowing the storage of information.<ref>[[#Tominaga2006|Tominaga 2006, p.&nbsp;327–8]]; [[#Chung2010|Chung 2010, p.&nbsp;285–6]]; [[#Kolobov 2012|Kolobov & Tominaga 2012, p.&nbsp;149]]</ref>
Varying compositions of GeSbTe ("[[GeSbTe|GST alloys]]") and [[AgInSbTe|Ag- and In- doped Sb<sub>2</sub>Te]] ("AIST alloys"), being examples of [[phase-change material]]s, are widely used in rewritable [[optical disc|optical disks]] and [[phase-change memory]] devices. By applying heat, they can be switched between amorphous (glassy) and [[crystalline]] states, changing their optical and electrical properties and allowing the storage of information.<ref>[[#Tominaga2006|Tominaga 2006, p.&nbsp;327–8]]; [[#Chung2010|Chung 2010, p.&nbsp;285–6]]; [[#Kolobov 2012|Kolobov & Tominaga 2012, p.&nbsp;149]]</ref>


===Semiconductors and electronics===
===Semiconductors and electronics===
[[File:Semiconductor-1.jpg|thumb|left|Semiconductor-based electronic components. From left to right: a [[transistor]], an [[integrated circuit]] and an [[LED]]. The elements commonly recognised as metalloids find widespread use in such devices, as elemental or [[compound semiconductor|compound]] semiconductor constituents (Si, Ge or [[GaAs]], for example) or as [[doping (semiconductor)|doping agents]] (B, Sb, Te, for example).|alt=A small square plastic piece with three parallel wire protrusions on one side; a larger rectangular plastic chip with multiple plastic and metal pin-like legs; and a small red light globe with two long wires coming out of its base.]]
[[File:Semiconductor-1.jpg|thumb|left|Semiconductor-based electronic components. From left to right: a [[transistor]], an [[integrated circuit]] and an [[LED]]. The elements commonly recognised as metalloids find widespread use in such devices, as elemental or [[compound semiconductor|compound]] semiconductor constituents (Si, Ge or [[GaAs]], for example) or as [[doping (semiconductor)|doping agents]] (B, Sb, Te, for example).|alt=A small square plastic piece with three parallel wire protrusions on one side; a larger rectangular plastic chip with multiple plastic and metal pin-like legs; and a small red light globe with two long wires coming out of its base.]]
All the elements commonly recognised as metalloids (or their compounds) have found application in the semiconductor or solid-state electronic industries.<ref>[[#Berger1997|Berger 1997, p.&nbsp;91]]; [[#Hampel1968|Hampel 1968, passim]]</ref> Some properties of boron have retarded 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.<ref>[[#Rochow1966|Rochow 1966, p.&nbsp;41]]; [[#Berger1997|Berger 1997, pp.&nbsp;42–3]]</ref> Silicon is the foremost commercial semiconductor; it forms the basis of modern electronics (including standard solar cells)<ref name=Bom>[[#Bomgardner|Bomgardner 2013, p.&nbsp;20]]</ref> and information and communication technologies.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;395]]; [[#Brown2009|Brown et al. 2009, p.&nbsp;489]]</ref> This occurred despite the study of semiconductors, early in the 20th century, being regarded as the 'physics of dirt' and not deserving of close attention.<ref>[[#Haller 2006|Haller 2006, p.&nbsp;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. [[Wolfgang Pauli|Pauli]] and I. [[Isidor Isaac Rabi|Rabi]], to comment derogatorily on the 'Physics of Dirt' '; [[#Hoddeson2007|Hoddeson 2007, pp.&nbsp;25–34 (29)]]</ref> Silicon has largely replaced germanium in semiconducting devices, being cheaper, more resilient at higher operating temperatures, and easier to work during the microelectronic fabrication process.<ref name=Russell2005401>[[#Russell2005|Russell & Lee 2005, p.&nbsp;401]]; [[#Büchel2003|Büchel, Moretto & Woditsch 2003, p.&nbsp;278]]</ref> Semiconducting [[silicon-germanium]] 'alloys' have however been growing in use, particularly for wireless communication devices; these alloys exploit the higher carrier mobility of germanium.<ref name=Russell2005401/> The synthesis of gram-scale quantities of semiconducting [[germanane]] was also reported in 2013. This comprises one-atom thick sheets of hydrogen-terminated germanium atoms. It conducts electrons more than ten times faster than silicon and five times faster than conventional germanium and is thought to have potential for a wide range of optoelectronic and sensing applications.<ref>[[#Bianco2013|Bianco et. al. 2013]]</ref>
All the elements commonly recognised as metalloids (or their compounds) have found application in the semiconductor or solid-state electronic industries.<ref>[[#Berger1997|Berger 1997, p.&nbsp;91]]; [[#Hampel1968|Hampel 1968, passim]]</ref> Some properties of boron have retarded 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.<ref>[[#Rochow1966|Rochow 1966, p.&nbsp;41]]; [[#Berger1997|Berger 1997, pp.&nbsp;42–3]]</ref> Silicon is the leading commercial semiconductor; it forms the basis of modern electronics (including standard solar cells)<ref name=Bom>[[#Bomgardner|Bomgardner 2013, p.&nbsp;20]]</ref> and information and communication technologies.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;395]]; [[#Brown2009|Brown et al. 2009, p.&nbsp;489]]</ref> This occurred despite the study of semiconductors, early in the 20th century, being regarded as the 'physics of dirt' and not deserving of close attention.<ref>[[#Haller 2006|Haller 2006, p.&nbsp;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. [[Wolfgang Pauli|Pauli]] and I. [[Isidor Isaac Rabi|Rabi]], to comment derogatorily on the 'Physics of Dirt' '; [[#Hoddeson2007|Hoddeson 2007, pp.&nbsp;25–34 (29)]]</ref> Silicon has largely replaced germanium in semiconducting devices, being cheaper, more resilient at higher operating temperatures, and easier to work during the microelectronic fabrication process.<ref name=Russell2005401>[[#Russell2005|Russell & Lee 2005, p.&nbsp;401]]; [[#Büchel2003|Büchel, Moretto & Woditsch 2003, p.&nbsp;278]]</ref> [[Silicon-germanium]] 'alloys' have been growing in use, particularly for wireless communication devices; these exploit the higher carrier mobility of germanium.<ref name=Russell2005401/> The synthesis of gram-scale quantities of semiconducting [[germanane]] was also reported in 2013. This comprises one-atom thick sheets of hydrogen-terminated germanium atoms. It conducts electrons more than ten times faster than silicon and five times faster than conventional germanium, and is thought to have potential for optoelectronic and sensing applications.<ref>[[#Bianco2013|Bianco et. al. 2013]]</ref>


Arsenic and antimony are not semiconductors in their [[standard state]]s. On the other hand, both form [[Compound semiconductor|type III-V semiconductors]] (such as [[GaAs]], [[AlSb]] or GaInAsSb) in which the average number of valence electrons per atom is the same as that of [[Group 14]] elements; these compounds are preferred for some special applications.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;421–2, 424]]</ref> Tellurium, which is a semiconductor in its standard state, is used mainly as a component in a very large group of type [[List of semiconductor materials|II/VI]] semiconducting-[[chalcogenide]]s; these compounds have applications in electro-optics and electronics.<ref>[[#Berger1997|Berger 1997, p.&nbsp;91]]</ref> [[Cadmium telluride]] (CdTe), in particular, finds application in solar modules due its high conversion efficiency, low manufacturing costs, and large [[band gap]] of 1.44 eV, meaning it absorbs a wide range of solar spectrum wavelengths.<ref name=Bom/> In the form of [[bismuth telluride]] (Bi<sub>2</sub>Te<sub>3</sub>) alloyed with selenium and antimony, tellurium is also a component of [[thermoelectric materials|thermoelectric devices]] used for refrigeration or portable power generation.<ref>[[#ScienceDaily|ScienceDaily 2012]]</ref> More ubiquitously, five of the preceding metalloids—boron, silicon, germanium, arsenic and antimony—can be found in cell phones (along with at least 39 other metals and nonmetals).<ref>[[#Reardon2005|Reardon 2005]]; [[#Meskers|Meskers, Hagelüken & Van Damme 2009, p.&nbsp;1131]]</ref> Tellurium, the last of the commonly recognised metalloids, is a component of [[phase change memory]] (see also above) and, as such, either has achieved cell phone incorporation, or is expected to find such use.<ref>[[#The Economist|The Economist 2012]]</ref> Of the elements less often recognised as metalloids, phosphorus, gallium (in particular) and selenium have semiconductor applications. Phosphorus is used in trace amounts as a [[dopant]] for [[n-type semiconductor]]s.<ref>[[#Whitten2007|Whitten 2007, p.&nbsp;488]]</ref> 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.<ref>[[#Jaskula|Jaskula 2013]]</ref> Selenium is used in the production of [[solar cell]]s<ref>[[#GES|German Energy Society 2008, p.&nbsp;43–44]]</ref> and in high-energy [[surge protector]]s.<ref>[[#Patel|Patel 2012, p.&nbsp;248]]</ref>
Arsenic and antimony are not semiconductors in their [[standard state]]s; both form [[Compound semiconductor|type III-V semiconductors]] (such as GaAs, [[AlSb]] or GaInAsSb) in which the average number of valence electrons per atom is the same as that of [[Group 14]] elements. These compounds are preferred for some special applications.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;421–2, 424]]</ref> Tellurium, which is a semiconductor in its standard state, is used mainly as a component in a large group of type [[List of semiconductor materials|II/VI]] semiconducting-[[chalcogenide]]s; these compounds have applications in electro-optics and electronics.<ref>[[#Berger1997|Berger 1997, p.&nbsp;91]]</ref> [[Cadmium telluride]] (CdTe) finds application in solar modules due its high conversion efficiency, low manufacturing costs, and large [[band gap]] of 1.44 eV, meaning it absorbs a wide range of solar wavelengths.<ref name=Bom/> [[Bismuth telluride]] (Bi<sub>2</sub>Te<sub>3</sub>) alloyed with selenium and antimony, is a component of [[thermoelectric materials|thermoelectric devices]] used for refrigeration or portable power generation.<ref>[[#ScienceDaily|ScienceDaily 2012]]</ref> Five metalloids—boron, silicon, germanium, arsenic and antimony—can be found in cell phones (along with at least 39 other metals and nonmetals).<ref>[[#Reardon2005|Reardon 2005]]; [[#Meskers|Meskers, Hagelüken & Van Damme 2009, p.&nbsp;1131]]</ref> Tellurium is a component of [[phase change memory]] and is expected to find such use.<ref>[[#The Economist|The Economist 2012]]</ref> Of the elements less often recognised as metalloids, phosphorus, gallium (in particular) and selenium have semiconductor applications. Phosphorus is used in trace amounts as a [[dopant]] for [[n-type semiconductor]]s.<ref>[[#Whitten2007|Whitten 2007, p.&nbsp;488]]</ref> 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.<ref>[[#Jaskula|Jaskula 2013]]</ref> Selenium is used in the production of [[solar cell]]s<ref>[[#GES|German Energy Society 2008, p.&nbsp;43–44]]</ref> and in high-energy [[surge protector]]s.<ref>[[#Patel|Patel 2012, p.&nbsp;248]]</ref>
{{clear}}
{{clear}}


==Nomenclature and history==
==Nomenclature and history==

===Derivation and other names===
===Derivation and other names===
The word metalloid comes from the [[Latin language|Latin]] ''metallum'' = "metal" and the [[Greek language|Greek]] ''oeides'' = "resembling in form or appearance".<ref>[[#OED1989|''Oxford English Dictionary'' 1989, 'metalloid']]; [[#GGH2003|Gordh, Gordh & Headrick 2003, p.&nbsp;753]]</ref> Although the terms '''amphoteric element,'''<ref>[[#Foster1936|Foster 1936, pp.&nbsp;212–13]]; [[#Brownlee1936|Brownlee et al. 1943, p.&nbsp;293]]</ref> '''boundary element,'''<ref>[[#Calderazzo|Calderazzo, Ercoli & Natta 1968, p. 257]]</ref> '''half-metal,'''<ref name=Klemm>[[#Klemm1950|Klemm 1950, pp.&nbsp;133–42]]; [[#Reilly2004|Reilly 2004, p.&nbsp;4]]</ref> '''half-way element,'''<ref>[[#Walters1982|Walters 1982, pp.&nbsp;32–3]]</ref> '''near metal,'''<ref name=tyler>[[#Tyler1948|Tyler 1948, p.&nbsp;105]]</ref> '''meta-metal,'''<ref>[[#Foster1958|Foster & Wrigley 1958, p.&nbsp;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''.'</ref> semiconductor,<ref>[[#Slade2006|Slade 2006, p.&nbsp;16]]</ref> semimetal<ref>[[#Corwin2005|Corwin 2005, p.&nbsp;80]]</ref> and '''submetal'''<ref>[[#Barsanov1974|Barsanov & Ginzburg 1974, p. 330]]; [[#Prokhorov|Prokhorov 1983, p.&nbsp;329]]</ref> are sometimes used synonymously, some of these have other meanings that may not be interchangeable. 'Amphoteric element' is sometimes used more broadly to include transition metals capable of forming [[oxyanion]]s, such as chromium and [[manganese]].<ref>[[#Bradbury1957|Bradbury et al. 1957, pp.&nbsp;157, 659]]</ref> As well, some elements referred to as metalloids do not show marked [[amphoterism|amphoteric]] behaviour or semiconductivity in their most stable forms. 'Half-metal' [[half-metal|is used in physics]] to refer to a compound (such as [[chromium dioxide]]) or alloy that can act as a conductor and an insulator. 'Meta-metal' is sometimes used instead to refer to certain metals ([[beryllium|Be]], [[zinc|Zn]], [[cadmium|Cd]], [[mercury (element)|Hg]], [[indium|In]], [[thallium|Tl]], [[Tin#Physical properties|β-Sn]], [[lead|Pb]]) located just to the left of the metalloids on standard periodic table layouts.<ref name=Klemm /> These metals are mostly diamagnetic<ref>[[#Miller2002|Miller, Lee & Choe 2002, p.&nbsp;21]]</ref> and tend to have distorted crystalline structures, electrical conductivity values at the lower end of those of metals, and amphoteric (weakly basic) oxides.<ref>[[#King2004|King 2004, pp.&nbsp;196–8]]; [[#Ferro2008|Ferro & Saccone 2008, p.&nbsp;233]]</ref> 'Semimetal' sometimes refers, loosely or explicitly, to metals with incomplete metallic character in crystalline structure, electrical conductivity or electronic structure. Examples include gallium,<ref>[[#Pashaey1973|Pashaey & Seleznev 1973, p.&nbsp;565]]; [[#Gladyshev1998|Gladyshev & Kovaleva 1998, p.&nbsp;1445]]; [[#Eason2007|Eason 2007, p.&nbsp;294]]</ref> [[ytterbium]],<ref>[[#Johansen1970|Johansen & Mackintosh 1970, pp.&nbsp;121–4]]; [[#Divakar1984|Divakar, Mohan & Singh 1984, p.&nbsp;2337]]; [[#Dávila2002|Dávila et al. 2002, p.&nbsp;035411-3]]</ref> bismuth<ref>[[#Jezequel1997|Jezequel & Thomas 1997, pp.&nbsp;6620–6]]</ref> and [[neptunium]].<ref>[[#Hindman1968|Hindman 1968, p.&nbsp;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.'; [[#Dunlap1970|Dunlap et al. 1970, pp.&nbsp;44, 46]]: '<span style="white-space: nowrap">...</span>α-Np is a semimetal, in which covalency effects are believed to also be of importance<span style="white-space: nowrap">...</span>For a semimetal having strong covalent bonding, like α-Np<span style="white-space: nowrap">...</span>'</ref>
The word metalloid comes from the [[Latin language|Latin]] ''metallum'' = "metal" and the [[Greek language|Greek]] ''oeides'' = "resembling in form or appearance".<ref>[[#OED1989|''Oxford English Dictionary'' 1989, 'metalloid']]; [[#GGH2003|Gordh, Gordh & Headrick 2003, p.&nbsp;753]]</ref> Although the terms '''amphoteric element,'''<ref>[[#Foster1936|Foster 1936, pp.&nbsp;212–13]]; [[#Brownlee1936|Brownlee et al. 1943, p.&nbsp;293]]</ref> '''boundary element,'''<ref>[[#Calderazzo|Calderazzo, Ercoli & Natta 1968, p. 257]]</ref> '''half-metal,'''<ref name=Klemm>[[#Klemm1950|Klemm 1950, pp.&nbsp;133–42]]; [[#Reilly2004|Reilly 2004, p.&nbsp;4]]</ref> '''half-way element,'''<ref>[[#Walters1982|Walters 1982, pp.&nbsp;32–3]]</ref> '''near metal,'''<ref name=tyler>[[#Tyler1948|Tyler 1948, p.&nbsp;105]]</ref> '''meta-metal,'''<ref>[[#Foster1958|Foster & Wrigley 1958, p.&nbsp;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''.'</ref> semiconductor,<ref>[[#Slade2006|Slade 2006, p.&nbsp;16]]</ref> semimetal<ref>[[#Corwin2005|Corwin 2005, p.&nbsp;80]]</ref> and '''submetal'''<ref>[[#Barsanov1974|Barsanov & Ginzburg 1974, p. 330]]; [[#Prokhorov|Prokhorov 1983, p.&nbsp;329]]</ref> are sometimes used synonymously, some of these have other meanings that may not be interchangeable. 'Amphoteric element' is sometimes used more broadly to include transition metals capable of forming [[oxyanion]]s, such as chromium and [[manganese]].<ref>[[#Bradbury1957|Bradbury et al. 1957, pp.&nbsp;157, 659]]</ref> Some elements referred to as metalloids do not show marked amphoteric behaviour or semiconductivity in their most stable forms. '[[Half-metal]]' is used in physics to refer to a compound (such as [[chromium dioxide]]) or alloy that can act as a conductor and an insulator. 'Meta-metal' is sometimes used instead to refer to certain metals ([[beryllium|Be]], [[zinc|Zn]], [[cadmium|Cd]], [[mercury (element)|Hg]], [[indium|In]], [[thallium|Tl]], [[Tin#Physical properties|β-Sn]], [[lead|Pb]]) located just to the left of the metalloids on standard periodic tables.<ref name=Klemm /> These metals are mostly diamagnetic<ref>[[#Miller2002|Miller, Lee & Choe 2002, p.&nbsp;21]]</ref> and tend to have distorted crystalline structures, electrical conductivity values at the lower end of those of metals, and amphoteric (weakly basic) oxides.<ref>[[#King2004|King 2004, pp.&nbsp;196–8]]; [[#Ferro2008|Ferro & Saccone 2008, p.&nbsp;233]]</ref> 'Semimetal' sometimes refers, loosely or explicitly, to metals with incomplete metallic character in crystalline structure, electrical conductivity or electronic structure. Examples include gallium,<ref>[[#Pashaey1973|Pashaey & Seleznev 1973, p.&nbsp;565]]; [[#Gladyshev1998|Gladyshev & Kovaleva 1998, p.&nbsp;1445]]; [[#Eason2007|Eason 2007, p.&nbsp;294]]</ref> [[ytterbium]],<ref>[[#Johansen1970|Johansen & Mackintosh 1970, pp.&nbsp;121–4]]; [[#Divakar1984|Divakar, Mohan & Singh 1984, p.&nbsp;2337]]; [[#Dávila2002|Dávila et al. 2002, p.&nbsp;035411-3]]</ref> bismuth<ref>[[#Jezequel1997|Jezequel & Thomas 1997, pp.&nbsp;6620–6]]</ref> and [[neptunium]].<ref>[[#Hindman1968|Hindman 1968, p.&nbsp;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.'; [[#Dunlap1970|Dunlap et al. 1970, pp.&nbsp;44, 46]]: '<span style="white-space: nowrap">...</span>α-Np is a semimetal, in which covalency effects are believed to also be of importance<span style="white-space: nowrap">...</span>For a semimetal having strong covalent bonding, like α-Np<span style="white-space: nowrap">...</span>'</ref>


===Origin and usage===
===Origin and usage===
{{Main|Metalloid (nomenclature origin and usage)}}
{{Main|Metalloid (nomenclature 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 at least as early as 1800.<ref>[[#Pinkerton1800|Pinkerton 1800, p.&nbsp;81]]</ref> Only recently, since the mid-20th century, has it been widely used to refer to intermediate or borderline chemical elements.<ref name="ReferenceA"/> The [[International Union of Pure and Applied Chemistry]] (IUPAC) has previously recommended abandoning the term metalloid, and suggested using the term semimetal instead.<ref>[[#Friend1953|Friend 1953, p.&nbsp;68]]; [[#IUPAC1959|IUPAC 1959, p.&nbsp;10]]; [[#IUPAC1971|IUPAC 1971, p.&nbsp;11]]</ref> However, use of this latter term has recently been discouraged<ref name=Atkins2010p20>[[#Atkins2010|Atkins et al. 2010, p.&nbsp;20]]</ref> as it has a distinct and different meaning in physics—one that more specifically refers to the [[electronic band structure]] of a substance rather than the overall classification of a chemical element. The most recent IUPAC publications on nomenclature and terminology do not include any recommendations on the usage or non-usage of the terms metalloid or semimetal.<ref>[[#IUPAC2005|IUPAC 2005]]; [[#IUPAC2006|IUPAC 2006–]]</ref>
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 as early as 1800.<ref>[[#Pinkerton1800|Pinkerton 1800, p.&nbsp;81]]</ref> Only since the mid-20th century has it been widely used to refer to intermediate or borderline chemical elements.<ref name="ReferenceA"/> The [[International Union of Pure and Applied Chemistry]] (IUPAC) has previously recommended abandoning the term metalloid, and suggested using the term semimetal instead.<ref>[[#Friend1953|Friend 1953, p.&nbsp;68]]; [[#IUPAC1959|IUPAC 1959, p.&nbsp;10]]; [[#IUPAC1971|IUPAC 1971, p.&nbsp;11]]</ref> Use of this latter term has more recently been discouraged<ref name=Atkins2010p20>[[#Atkins2010|Atkins et al. 2010, p.&nbsp;20]]</ref> 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 or non-usage of the terms metalloid or semimetal.<ref>[[#IUPAC2005|IUPAC 2005]]; [[#IUPAC2006|IUPAC 2006–]]</ref>


==Elements commonly recognised as metalloids==
==Elements commonly recognised as metalloids==
:''This section presents brief sketches of the physical and chemical properties of the applicable elements—in their most thermodynamically stable forms under ambient conditions. For complete profiles, including history, production, and specific uses, see the main article for each element.''


===Boron===
===Boron===
{{main|Boron}}
{{main|Boron}}
[[File:Boron R105.jpg|thumb|right|Boron, shown here in the form of its β-rhombohedral phase (its most thermodynamically stable allotrope)<ref>[[#VanSetten2007|Van Setten et al. 2007, pp.&nbsp;2460–1]]; [[#Oganov2009|Oganov et al. 2009, pp.&nbsp;863–4]]</ref>|alt=Several dozen small angular stone like shapes, grey in colour with scattered silver flecks and highlights.]] Pure boron appears as a shiny, silver-grey crystalline solid.<ref>[[#Housecroft2008|Housecroft & Sharpe 2008, p.&nbsp;331]]; [[#Oganov2010|Oganov 2010, p.&nbsp;212]]</ref> It is somewhat less dense than aluminium, hard and brittle. It is barely reactive under normal conditions, except for attack by fluorine,<ref>[[#Housecroft2008|Housecroft & Sharpe 2008, p.&nbsp;333]]</ref> and has a melting point several hundred degrees higher than that of steel. Boron is a semiconductor,<ref>[[#Berger1997|Berger 1997, p.&nbsp;37]]</ref> with a room temperature electrical conductivity of 1.5 × 10<sup>−6</sup> [[Siemens (unit)|S]]•cm<sup>−1</sup><ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;144]]</ref> (about 200 times less than that of tap water)<ref>[[#Kopp|Kopp, Lipták & Eren 2003, p.&nbsp221]]</ref> and a band gap of about 1.56&nbsp;eV.<ref>[[#Prudenziati1977|Prudenziati 1977, p.&nbsp;242]]</ref>{{#tag:ref|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).<ref>[[#Berger1997|Berger 1997, pp.&nbsp;87, 84]]</ref>|group=n}}
[[File:Boron R105.jpg|thumb|right|Boron, shown here in the form of its β-rhombohedral phase (its most thermodynamically stable allotrope)<ref>[[#VanSetten2007|Van Setten et al. 2007, pp.&nbsp;2460–1]]; [[#Oganov2009|Oganov et al. 2009, pp.&nbsp;863–4]]</ref>|alt=Several dozen small angular stone like shapes, grey with scattered silver flecks and highlights.]] Pure boron is a shiny, silver-grey crystalline solid.<ref>[[#Housecroft2008|Housecroft & Sharpe 2008, p.&nbsp;331]]; [[#Oganov2010|Oganov 2010, p.&nbsp;212]]</ref> It is less dense than aluminium, and is hard and brittle. It is barely reactive under normal conditions, except for attack by fluorine,<ref>[[#Housecroft2008|Housecroft & Sharpe 2008, p.&nbsp;333]]</ref> and has a melting point several hundred degrees higher than that of steel. Boron is a semiconductor,<ref>[[#Berger1997|Berger 1997, p.&nbsp;37]]</ref> with a room temperature electrical conductivity of 1.5 × 10<sup>−6</sup> [[Siemens (unit)|S]]•cm<sup>−1</sup><ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;144]]</ref> (about 200 times less than that of tap water)<ref>[[#Kopp|Kopp, Lipták & Eren 2003, p.&nbsp221]]</ref> and a band gap of about 1.56&nbsp;eV.<ref>[[#Prudenziati1977|Prudenziati 1977, p.&nbsp;242]]</ref>{{#tag:ref|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).<ref>[[#Berger1997|Berger 1997, pp.&nbsp;87, 84]]</ref>|group=n}}


The chemistry of boron is dominated by its small size, relatively high ionization energy, and having fewer valence electrons (three) than atomic orbitals (four) available for bonding. With only three valence electrons, simple covalent bonding is electron deficient with respect to the [[octet rule]].<ref>[[#Rayner2006|Rayner-Canham & Overton 2006, p.&nbsp;291]]</ref> Elements in this situation usually adopt metallic bonding. However, the small size and high ionization energies of boron tends to result in delocalized covalent bonding,<ref>[[#Bowser1993|Bowser 1993, p.&nbsp;393]]; [[#Grimes2011|Grimes 2011, pp.&nbsp;17–18]]</ref> in which three atoms share two electrons, rather than metallic bonding. The associated structural component that pervades the various allotropes of boron is the [[icosahedron|icosahedral]] B<sub>12</sub> unit. This likewise occurs, as do [[deltahedron|deltahedral]] variants or fragments, in several metal borides, certain hydrides, and some halides.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;141]]; [[#Henderson2000|Henderson 2000, p.&nbsp;58]]; [[#Housecroft2008|Housecroft & Sharpe 2008, pp.&nbsp;360–72]]</ref> The bonding in boron has been described as being characteristic of behaviour intermediate between metals and nonmetallic covalent network solids (a classic example of the latter being [[diamond]]).<ref>[[#Parry1970|Parry et al. 1970, pp.&nbsp;438, 448–51]]</ref> 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 how close boron is to the metal-nonmetal borderline.<ref name=Fehlner1990>[[#Fehlner1990|Fehlner 1990, p.&nbsp;202]]</ref>
The chemistry of boron is dominated by its small atomic size, relatively high ionization energy, and having fewer valence electrons (three) than atomic orbitals (four) available for bonding. With only three valence electrons, simple covalent bonding is electron deficient with respect to the [[octet rule]].<ref>[[#Rayner2006|Rayner-Canham & Overton 2006, p.&nbsp;291]]</ref> Elements in this situation usually adopt metallic bonding, but the small size and high ionization energies of boron tends to result in delocalized covalent bonding,<ref>[[#Bowser1993|Bowser 1993, p.&nbsp;393]]; [[#Grimes2011|Grimes 2011, pp.&nbsp;17–18]]</ref> in which three atoms share two electrons, rather than metallic bonding. The associated structural component that pervades the various allotropes of boron is the [[icosahedron|icosahedral]] B<sub>12</sub> unit. This also occurs, as do [[deltahedron|deltahedral]] variants or fragments, in several metal borides, certain hydrides, and some halides.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;141]]; [[#Henderson2000|Henderson 2000, p.&nbsp;58]]; [[#Housecroft2008|Housecroft & Sharpe 2008, pp.&nbsp;360–72]]</ref> The bonding in boron has been described as being characteristic of behaviour intermediate between metals and nonmetallic covalent network solids (such as [[diamond]]).<ref>[[#Parry1970|Parry et al. 1970, pp.&nbsp;438, 448–51]]</ref> 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 how close boron is to the metal-nonmetal borderline.<ref name=Fehlner1990>[[#Fehlner1990|Fehlner 1990, p.&nbsp;202]]</ref>


Most of the chemistry of boron is nonmetallic in nature.<ref name=Fehlner1990/> The small size of the boron atom, however, enables the preparation of many interstitial alloy-type borides.<ref name=Greenwood145>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;145]]</ref> Analogies between boron and transition metals have additionally been noted in the formation of complexes,<ref>[[#Houghton1979|Houghton 1979, p.&nbsp;59]]</ref> and [[adduct]]s (for example, BH<sub>3</sub> + [[Carbon monoxide|CO]] →BH<sub>3</sub>CO and, similarly Fe(CO)<sub>4</sub> + CO →Fe(CO)<sub>5</sub>), as well in the geometric and electronic structures of [[cluster compound|cluster species]] such as [B<sub>6</sub>H<sub>6</sub>]<sup>2–</sup> and [Ru<sub>6</sub>(CO)<sub>18</sub>]<sup>2–</sup>.<ref>[[#Fehlner1990|Fehlner 1990, pp.&nbsp;204, 207]]</ref>{{#tag:ref|On the analogy between boron and metals, Greenwood<ref>[[#Greenwood2001|Greenwood 2001, p. 2057]]</ref> 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...'.|group=n}} The aqueous chemistry of boron, more conventionally, is characterised by the formation of many different polyborate anions.<ref>[[#Salentine1987|Salentine 1987, pp.&nbsp;128–32]]; [[#MacKay2002|MacKay, MacKay & Henderson 2002, pp.&nbsp;439–40]]; [[#Kneen1972|Kneen, Rogers & Simpson 1972, p.&nbsp;394]]; [[#Hiller1960|Hiller & Herber 1960, inside front cover; p.&nbsp;225]]</ref> Given its high charge-to-size ratio nearly all compounds of boron are covalent, barring some complexed anionic and cationic species.<ref>[[#Watt1958|Watt 1958, p.&nbsp;387]]; [[#Sharp1983|Sharp 1983]]</ref> Boron has a strong affinity for oxygen, a characteristic manifested in the extensive chemistry of the [[borate]]s.<ref name=Greenwood145/> The oxide B<sub>2</sub>O<sub>3</sub> is polymeric in structure,<ref name=Pudd59>[[#Puddephatt1989|Puddephatt & Monaghan 1989, p.&nbsp;59]]</ref> weakly acidic,<ref>[[#Mahan1965|Mahan 1965, p.&nbsp;485]]</ref> and a glass former.<ref name=Rao22>[[#Rao2002|Rao 2002, p.&nbsp;22]]</ref> [[Organometallic chemistry|Organometallic compounds]] of boron have been known since the 19th century (see [[organoboron chemistry]]).<ref>[[#Haiduc1985|Haiduc & Zuckerman 1985, p.&nbsp;82]]</ref>
Most of the chemistry of boron is nonmetallic in nature.<ref name=Fehlner1990/> The small size of the boron atom enables the preparation of many interstitial alloy-type borides.<ref name=Greenwood145>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;145]]</ref> Analogies between boron and transition metals have been noted in the formation of complexes,<ref>[[#Houghton1979|Houghton 1979, p.&nbsp;59]]</ref> and [[adduct]]s (for example, BH<sub>3</sub> + [[Carbon monoxide|CO]] →BH<sub>3</sub>CO and, similarly Fe(CO)<sub>4</sub> + CO →Fe(CO)<sub>5</sub>), as well as in the geometric and electronic structures of [[cluster compound|cluster species]] such as [B<sub>6</sub>H<sub>6</sub>]<sup>2–</sup> and [Ru<sub>6</sub>(CO)<sub>18</sub>]<sup>2–</sup>.<ref>[[#Fehlner1990|Fehlner 1990, pp.&nbsp;204, 207]]</ref>{{#tag:ref|On the analogy between boron and metals, Greenwood<ref>[[#Greenwood2001|Greenwood 2001, p. 2057]]</ref> 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...'.|group=n}} The aqueous chemistry of boron is characterised by the formation of many different polyborate anions.<ref>[[#Salentine1987|Salentine 1987, pp.&nbsp;128–32]]; [[#MacKay2002|MacKay, MacKay & Henderson 2002, pp.&nbsp;439–40]]; [[#Kneen1972|Kneen, Rogers & Simpson 1972, p.&nbsp;394]]; [[#Hiller1960|Hiller & Herber 1960, inside front cover; p.&nbsp;225]]</ref> Given its high charge-to-size ratio, nearly all compounds of boron are covalent, with a few complexed anionic and cationic species.<ref>[[#Watt1958|Watt 1958, p.&nbsp;387]]; [[#Sharp1983|Sharp 1983]]</ref> Boron has a strong affinity for oxygen, and has an extensive chemistry as [[borate]]s.<ref name=Greenwood145/> The oxide B<sub>2</sub>O<sub>3</sub> is polymeric in structure,<ref name=Pudd59>[[#Puddephatt1989|Puddephatt & Monaghan 1989, p.&nbsp;59]]</ref> weakly acidic,<ref>[[#Mahan1965|Mahan 1965, p.&nbsp;485]]</ref> and a glass former.<ref name=Rao22>[[#Rao2002|Rao 2002, p.&nbsp;22]]</ref> [[Organometallic chemistry|Organometallic compounds]] of boron have been known since the 19th century (see [[organoboron chemistry]]).<ref>[[#Haiduc1985|Haiduc & Zuckerman 1985, p.&nbsp;82]]</ref>


===Silicon===
===Silicon===
{{main|Silicon}}
{{main|Silicon}}
[[File:SiliconCroda.jpg|thumb|left|Silicon has a shiny blue-grey metallic lustre.|alt=A lustrous blue gray potato shaped lump with an uneven and irregular corrugated surface.]]
[[File:SiliconCroda.jpg|thumb|left|Silicon has a shiny blue-grey metallic lustre.|alt=A lustrous blue grey potato shaped lump with an irregular corrugated surface.]]
Silicon appears as a shiny crystalline solid, with a blue-grey metallic lustre.<ref name=Greenwood331>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;331]]</ref> As with boron, it is somewhat less dense than aluminium, hard and brittle.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;824]]</ref> It is a relatively unreactive element,<ref name=Greenwood331/> the massive, crystalline form (especially if pure) being 'remarkably inert to all acids, including hydrofluoric'.<ref>[[#Rochow1973|Rochow 1973, p.&nbsp;1337‒38]]</ref> Less pure silicon, and the powdered form, are variously susceptible to attack by strong or heated acids, as well as by steam and fluorine.<ref>[[#Rochow1973|Rochow 1973, p.&nbsp;1337, 1340]]</ref> Silicon also dissolves in hot aqueous alkalis with the evolution of hydrogen, behaving in this way like metals<ref>[[#Allen1968|Allen & Ordway 1968, p.&nbsp;152]]</ref> such as beryllium, aluminium, zinc, gallium and indium.<ref>[[#Eagleson1994|Eagleson 1994, pp.&nbsp;48, 127, 438, 1194]]; [[#Massey2000|Massey 2000, p.&nbsp;191]]</ref> It melts at about the same temperature as steel. Silicon is a semiconductor with an electrical conductivity of 10<sup>−4</sup>&nbsp;S•cm<sup>−1</sup><ref>[[#Orton2004|Orton 2004, p.&nbsp;7.]] The listed figure is a typical value for high-purity silicon.</ref> and a band gap of about 1.11&nbsp;eV.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;393]]</ref> When it melts, silicon becomes a reasonable metal<ref>[[#Coles1976|Coles & Caplin 1976, p.&nbsp;106]]</ref> with an electrical conductivity of 1.0–1.3 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup>, a value similar to that of liquid mercury.<ref>[[#Glazov1969|Glazov, Chizhevskaya & Glagoleva 1969, pp.&nbsp;59–63]]; [[#Allen1987|Allen & Broughton 1987, p.&nbsp;4967]]</ref>
Silicon is a shiny crystalline solid, with a blue-grey metallic lustre.<ref name=Greenwood331>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;331]]</ref> Like boron, it is less dense than aluminium, and is hard and brittle.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;824]]</ref> It is a relatively unreactive element,<ref name=Greenwood331/> the massive, crystalline form (especially if pure) being 'remarkably inert to all acids, including hydrofluoric'.<ref>[[#Rochow1973|Rochow 1973, p.&nbsp;1337‒38]]</ref> Less pure silicon, and the powdered form, are variously susceptible to attack by strong or heated acids, as well as by steam and fluorine.<ref>[[#Rochow1973|Rochow 1973, p.&nbsp;1337, 1340]]</ref> Silicon reacts with hot aqueous alkalis with the evolution of hydrogen, like a metal<ref>[[#Allen1968|Allen & Ordway 1968, p.&nbsp;152]]</ref> such as beryllium, aluminium, zinc, gallium and indium.<ref>[[#Eagleson1994|Eagleson 1994, pp.&nbsp;48, 127, 438, 1194]]; [[#Massey2000|Massey 2000, p.&nbsp;191]]</ref> It melts at about the same temperature as steel. Silicon is a semiconductor with an electrical conductivity of 10<sup>−4</sup>&nbsp;S•cm<sup>−1</sup><ref>[[#Orton2004|Orton 2004, p.&nbsp;7.]] The listed figure is a typical value for high-purity silicon.</ref> and a band gap of about 1.11&nbsp;eV.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;393]]</ref> When it melts, silicon becomes a reasonable metal<ref>[[#Coles1976|Coles & Caplin 1976, p.&nbsp;106]]</ref> with an electrical conductivity of 1.0–1.3 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup>, similar to that of liquid mercury.<ref>[[#Glazov1969|Glazov, Chizhevskaya & Glagoleva 1969, pp.&nbsp;59–63]]; [[#Allen1987|Allen & Broughton 1987, p.&nbsp;4967]]</ref>


The chemistry of silicon is generally nonmetallic (covalent) in nature.<ref>[[#Cotton1995|Cotton, Wilkinson & Gaus 1995, p.&nbsp;393]]</ref> It does, however, form alloys with metals such as iron and copper.<ref>[[#Partington1944|Partington 1944, p.&nbsp;723]]</ref> Silicon shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox>[[#Cox2004|Cox 2004, p.&nbsp;27]]</ref> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225>[[#Hiller1960|Hiller & Herber 1960, inside front cover; p.&nbsp;225]]</ref> The high bond strength of the silicon-oxygen bond dominates the chemical behaviour of silicon.<ref>[[#Kneen1972|Kneen, Rogers and Simpson 1972, p.&nbsp;384]]</ref> Polymeric silicates, built up by tetrahedral SiO<sub>4</sub> units sharing their oxygen atoms, represent the most abundant and important compounds of silicon.<ref name="Bailar513"/> The polymeric borates, comprising linked trigonal and tetrahedral BO<sub>3</sub> or BO<sub>4</sub> units, are built on similar structural principles.<ref>[[#Cotton1995|Cotton, Wilkinson & Gaus 1995, pp.&nbsp;319, 321]]</ref> The oxide SiO<sub>2</sub> is polymeric in structure,<ref name=Pudd59/> weakly acidic,<ref>[[#Smith1990|Smith 1990, p.&nbsp;175]]</ref>{{#tag:ref|Although SiO<sub>2</sub> is classified as an acidic oxide, and hence reacts with alkalis to give silicates, its reaction with phosphoric acid yields a silicon oxide orthophosphate Si<sub>5</sub>O(PO<sub>4</sub>)<sub>6</sub>,<ref>[[#Poojary1993|Poojary, Borade & Clearfield 1993]]</ref> and with hydrofluoric acid to give [[hexafluorosilicic acid]] H<sub>2</sub>SiF<sub>6</sub>.<ref>[[#Wiberg2001|Wiberg 2001, pp.&nbsp;851, 858]]</ref>|group=n}} and a glass former.<ref name=Rao22/> Traditional organometallic chemistry includes the carbon compounds of silicon (see [[organosilicon]]).<ref>[[#Powell1988|Powell 1988, p.&nbsp;1]]</ref>
The chemistry of silicon is generally nonmetallic (covalent) in nature.<ref>[[#Cotton1995|Cotton, Wilkinson & Gaus 1995, p.&nbsp;393]]</ref> It can form alloys with metals such as iron and copper.<ref>[[#Partington1944|Partington 1944, p.&nbsp;723]]</ref> Silicon shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox>[[#Cox2004|Cox 2004, p.&nbsp;27]]</ref> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225>[[#Hiller1960|Hiller & Herber 1960, inside front cover; p.&nbsp;225]]</ref> The high strength of the silicon-oxygen bond dominates the chemical behaviour of silicon.<ref>[[#Kneen1972|Kneen, Rogers and Simpson 1972, p.&nbsp;384]]</ref> Polymeric silicates, built up by tetrahedral SiO<sub>4</sub> units sharing their oxygen atoms, are the most abundant and important compounds of silicon.<ref name="Bailar513"/> The polymeric borates, comprising linked trigonal and tetrahedral BO<sub>3</sub> or BO<sub>4</sub> units, are built on similar structural principles.<ref>[[#Cotton1995|Cotton, Wilkinson & Gaus 1995, pp.&nbsp;319, 321]]</ref> The oxide SiO<sub>2</sub> is polymeric in structure,<ref name=Pudd59/> weakly acidic,<ref>[[#Smith1990|Smith 1990, p.&nbsp;175]]</ref>{{#tag:ref|Although SiO<sub>2</sub> is classified as an acidic oxide, and hence reacts with alkalis to give silicates, its reaction with phosphoric acid yields a silicon oxide orthophosphate Si<sub>5</sub>O(PO<sub>4</sub>)<sub>6</sub>,<ref>[[#Poojary1993|Poojary, Borade & Clearfield 1993]]</ref> and with hydrofluoric acid to give [[hexafluorosilicic acid]] H<sub>2</sub>SiF<sub>6</sub>.<ref>[[#Wiberg2001|Wiberg 2001, pp.&nbsp;851, 858]]</ref>|group=n}} and a glass former.<ref name=Rao22/> Traditional organometallic chemistry includes the carbon compounds of silicon (see [[organosilicon]]).<ref>[[#Powell1988|Powell 1988, p.&nbsp;1]]</ref>


===Germanium===
===Germanium===
Line 290: Line 290:
Germanium appears as a shiny grey-white solid.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;371]]</ref> It is about one-third less dense than iron, hard and brittle.<ref>[[#Cusack1967|Cusack 1967, p.&nbsp;193]]</ref> It is mostly unreactive at room temperature{{#tag:ref|Temperatures above 400&nbsp;°C are required to form a noticeable surface oxide layer.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;399–400]]</ref>|group=n}} but is slowly attacked by hot concentrated sulfuric or nitric acid.<ref name=Greenwood373>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;373]]</ref> Germanium also reacts with molten [[sodium hydroxide|caustic soda]] to yield sodium germanate Na<sub>2</sub>GeO<sub>3</sub>, together with the evolution of hydrogen.<ref>[[#Moody1991|Moody 1991, p.&nbsp;273]]</ref> It melts at a lower temperature than steel, 938 °C vs. ~1400 °C. Germanium is a semiconductor with an electrical conductivity of around 2 × 10<sup>−2</sup>&nbsp;S•cm<sup>−1</sup><ref name=Greenwood373/> and a band gap of 0.67&nbsp;eV.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;399]]</ref> Liquid germanium is a metallic conductor, with an electrical conductivity on par with that of liquid mercury.<ref>[[#Berger1997|Berger 1997, pp.&nbsp;71–2]]</ref>
Germanium appears as a shiny grey-white solid.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;371]]</ref> It is about one-third less dense than iron, hard and brittle.<ref>[[#Cusack1967|Cusack 1967, p.&nbsp;193]]</ref> It is mostly unreactive at room temperature{{#tag:ref|Temperatures above 400&nbsp;°C are required to form a noticeable surface oxide layer.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;399–400]]</ref>|group=n}} but is slowly attacked by hot concentrated sulfuric or nitric acid.<ref name=Greenwood373>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;373]]</ref> Germanium also reacts with molten [[sodium hydroxide|caustic soda]] to yield sodium germanate Na<sub>2</sub>GeO<sub>3</sub>, together with the evolution of hydrogen.<ref>[[#Moody1991|Moody 1991, p.&nbsp;273]]</ref> It melts at a lower temperature than steel, 938 °C vs. ~1400 °C. Germanium is a semiconductor with an electrical conductivity of around 2 × 10<sup>−2</sup>&nbsp;S•cm<sup>−1</sup><ref name=Greenwood373/> and a band gap of 0.67&nbsp;eV.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;399]]</ref> Liquid germanium is a metallic conductor, with an electrical conductivity on par with that of liquid mercury.<ref>[[#Berger1997|Berger 1997, pp.&nbsp;71–2]]</ref>


Most of the chemistry of germanium is characteristic of a nonmetal.<ref>[[#Jolly1966|Jolly 1966, pp.&nbsp;125–6]]</ref> It does however form alloys with, for example, aluminium and gold.<ref>[[#Schwartz2002|Schwartz 2002, p.&nbsp;269]]</ref> Germanium shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox/> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Germanium generally forms tetravalent (IV) compounds, although it can form a smaller number of less stable divalent (II) compounds, in which it behaves more like a metal.<ref name="ReferenceC">[[#Eggins1972|Eggins 1972, p.&nbsp;66]]; [[#Wiberg2001|Wiberg 2001, p.&nbsp;895]]</ref> Germanium analogues of all of the major types of silicates have been prepared.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;383]]</ref> The metallic character of germanium is also suggested by the formation of various oxoacid salts. A phosphate [(HPO<sub>4</sub>)<sub>2</sub>Ge·H<sub>2</sub>O] and highly stable trifluoroacetate Ge(OCOCF<sub>3</sub>)<sub>4</sub> have been described, as have Ge<sub>2</sub>(SO<sub>4</sub>)<sub>2</sub>, Ge(ClO<sub>4</sub>)<sub>4</sub> and GeH<sub>2</sub>(C<sub>2</sub>O<sub>4</sub>)<sub>3</sub>.<ref>[[#Glockling1969|Glockling 1969, p.&nbsp;38]]; [[#Wells1984|Wells 1984, p.&nbsp;1175]]</ref> The oxide GeO<sub>2</sub> is polymeric,<ref name=Pudd59/> amphoteric,<ref>[[#Cooper1968|Cooper 1968, pp.&nbsp;28–9]]</ref> and a glass former.<ref name=Rao22/> The fact that the dioxide is soluble in acidic solutions (as is the monoxide GeO, only more so), is sometimes used as a basis to classify germanium as a metal.<ref>[[#Steele1966|Steele 1966, pp.&nbsp;178, 188–9]]</ref> Indeed, up to at least the 1930s germanium was considered to be a poorly conducting metal rather than a nonmetal.<ref name="Haller EE 2006, p.&nbsp;3"/> As is the case with all the elements commonly recognised as metalloids, germanium has an established organometallic chemistry (see [[organogermanium chemistry]]).<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;742]]</ref>
Most of the chemistry of germanium is characteristic of a nonmetal.<ref>[[#Jolly1966|Jolly 1966, pp.&nbsp;125–6]]</ref> It forms alloys with aluminium and gold.<ref>[[#Schwartz2002|Schwartz 2002, p.&nbsp;269]]</ref> Germanium shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox/> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Germanium generally forms tetravalent (IV) compounds, and it can also form less stable divalent (II) compounds, in which it behaves more like a metal.<ref name="ReferenceC">[[#Eggins1972|Eggins 1972, p.&nbsp;66]]; [[#Wiberg2001|Wiberg 2001, p.&nbsp;895]]</ref> Germanium analogues of all of the major types of silicates have been prepared.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;383]]</ref> The metallic character of germanium is also suggested by the formation of various oxoacid salts. A phosphate [(HPO<sub>4</sub>)<sub>2</sub>Ge·H<sub>2</sub>O] and highly stable trifluoroacetate Ge(OCOCF<sub>3</sub>)<sub>4</sub> have been described, as have Ge<sub>2</sub>(SO<sub>4</sub>)<sub>2</sub>, Ge(ClO<sub>4</sub>)<sub>4</sub> and GeH<sub>2</sub>(C<sub>2</sub>O<sub>4</sub>)<sub>3</sub>.<ref>[[#Glockling1969|Glockling 1969, p.&nbsp;38]]; [[#Wells1984|Wells 1984, p.&nbsp;1175]]</ref> The oxide GeO<sub>2</sub> is polymeric,<ref name=Pudd59/> amphoteric,<ref>[[#Cooper1968|Cooper 1968, pp.&nbsp;28–9]]</ref> and a glass former.<ref name=Rao22/> The dioxide is soluble in acidic solutions (and the monoxide GeO, is even more so), and this is sometimes used as a basis to classify germanium as a metal.<ref>[[#Steele1966|Steele 1966, pp.&nbsp;178, 188–9]]</ref> Up to the 1930s germanium was considered to be a poorly conducting metal rather than a nonmetal.<ref name="Haller EE 2006, p.&nbsp;3"/> As is the case with all the elements commonly recognised as metalloids, germanium has an established organometallic chemistry (see [[organogermanium chemistry]]).<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;742]]</ref>


===Arsenic===
===Arsenic===
{{main|Arsenic}}
{{main|Arsenic}}
[[File:Arsen 1a.jpg|thumb|left|Arsenic, sealed in a container to prevent tarnishing|alt=Two dull silver clusters of crystalline shards.]]
[[File:Arsen 1a.jpg|thumb|left|Arsenic, sealed in a container to prevent tarnishing|alt=Two dull silver clusters of crystalline shards.]]
Arsenic is a grey, metallic looking solid. It is about one-third less dense than iron, brittle, and moderately hard (more than aluminium; less than iron).<ref name="GWM2011">[[#Gray2011|Gray, Whitby & Mann 2011]]</ref> It is stable in dry air but develops a golden bronze patina in moist air, which blackens on further exposure. Arsenic is attacked by nitric acid and concentrated sulfuric acid. It reacts with fused caustic soda to give the arsenate Na<sub>3</sub>AsO<sub>3</sub>, together with the evolution of hydrogen.<ref name="Greenwood 2002, p. 552">[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;552]]</ref> Arsenic sublimes, rather than melts, at a temperature of 615 °C, which is significantly lower than the melting point of steel, ca. 1400 °C.<ref name=steelMP/> The vapour is lemon-yellow and smells like garlic.<ref>[[#Parkes1943|Parkes & Mellor 1943, p.&nbsp;740]]</ref> Arsenic only melts under a pressure of 38.6 [[Atmosphere (unit)|atm]], at 817 °C.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;420]]</ref> Arsenic is a semimetal with an electrical conductivity of around 3.9 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup><ref name="Carapella1968p30">[[#Carapella1968|Carapella 1968, p.&nbsp;30]]</ref> and a band overlap of 0.5&nbsp;eV.<ref name="Barfuß 1981, p. 967">[[#Barfuß1981|Barfuß et al. 1981, p.&nbsp;967]]</ref>{{#tag:ref|Arsenic also exists as a naturally occurring (but rare) allotrope ''(arsenolamprite),'' this being a semiconductor with a [[band gap]] of around 0.3&nbsp;eV or 0.4&nbsp;eV. It can furthermore be prepared in a semiconducting [[amorphous solid|amorphous]] form, with a band gap of around 1.2–1.4&nbsp;eV.<ref>[[#Greaves1974|Greaves, Knights & Davis 1974, p.&nbsp;369]]; [[#Madelung2004|Madelung 2004, pp.&nbsp;405, 410]]</ref>|group=n}} Liquid arsenic is a semiconductor with a band gap of 0.15&nbsp;eV.<ref>[[#Bailar1973|Bailar & Trotman-Dickenson 1973, p.&nbsp;558]]; [[#Li1990|Li 1990]]</ref>
Arsenic is a grey, metallic looking solid. It is about one-third less dense than iron, brittle, and moderately hard (more than aluminium; less than iron).<ref name="GWM2011">[[#Gray2011|Gray, Whitby & Mann 2011]]</ref> It is stable in dry air but develops a golden bronze patina in moist air, which blackens on further exposure. Arsenic is attacked by nitric acid and concentrated sulfuric acid. It reacts with fused caustic soda to give the arsenate Na<sub>3</sub>AsO<sub>3</sub>, together with the evolution of hydrogen.<ref name="Greenwood 2002, p. 552">[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;552]]</ref> Arsenic sublimes at 615 °C. The vapour is lemon-yellow and smells like garlic.<ref>[[#Parkes1943|Parkes & Mellor 1943, p.&nbsp;740]]</ref> Arsenic only melts under a pressure of 38.6 [[Atmosphere (unit)|atm]], at 817 °C.<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;420]]</ref> It is a semimetal with an electrical conductivity of around 3.9 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup><ref name="Carapella1968p30">[[#Carapella1968|Carapella 1968, p.&nbsp;30]]</ref> and a band overlap of 0.5&nbsp;eV.<ref name="Barfuß 1981, p. 967">[[#Barfuß1981|Barfuß et al. 1981, p.&nbsp;967]]</ref>{{#tag:ref|Arsenic also exists as a naturally occurring (but rare) allotrope ''(arsenolamprite),'' a semiconductor with a band gap of around 0.3&nbsp;eV or 0.4&nbsp;eV. It can be prepared in a semiconducting [[amorphous solid|amorphous]] form, with a band gap of around 1.2–1.4&nbsp;eV.<ref>[[#Greaves1974|Greaves, Knights & Davis 1974, p.&nbsp;369]]; [[#Madelung2004|Madelung 2004, pp.&nbsp;405, 410]]</ref>|group=n}} Liquid arsenic is a semiconductor with a band gap of 0.15&nbsp;eV.<ref>[[#Bailar1973|Bailar & Trotman-Dickenson 1973, p.&nbsp;558]]; [[#Li1990|Li 1990]]</ref>


The chemistry of arsenic is predominately nonmetallic in character.<ref>[[#Bailar1965|Bailar, Moeller & Kleinberg 1965, p.&nbsp;477]]</ref> It does however form alloys with many metals, most of these being brittle.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;91]]</ref> Arsenic shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox/> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Arsenic generally forms compounds in which it has an oxidation state of +3 or +5.<ref name="Massey267">[[#Massey2000|Massey 2000, p.&nbsp;267]]</ref> The halides, and the oxides and their derivatives are illustrative examples.<ref name="Bailar513">[[#Bailar1965|Bailar, Moeller & Kleinberg 1965, p.&nbsp;513]]</ref> In the trivalent state, arsenic shows some incipient metallic properties.<ref>[[#Timm1944|Timm 1944, p.&nbsp;454]]</ref> Thus, the halides are hydrolysed by water but these reactions, particularly those of the chloride, are reversible with the addition of a [[hydrohalic acid]].<ref>[[#Partington1944|Partington 1944, p.&nbsp;641]]; [[#Kleinberg1960|Kleinberg, Argersinger & Griswold 1960, p.&nbsp;419]]</ref> As well, and as noted below, the oxide is acidic but weakly amphoteric. The higher, less stable, pentavalent state has strongly acidic (nonmetallic) properties.<ref>[[#Morgan1906|Morgan 1906, p.&nbsp;163]]; [[#Moeller1954|Moeller 1954, p.&nbsp;559]]</ref> More generally, and compared to phosphorus, the stronger metallic character of arsenic is indicated by the formation of oxoacid salts such as AsPO<sub>4</sub>, As<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub>{{#tag:ref|The formulae of AsPO<sub>4</sub> and As<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub> suggest straightforward ionic formulations, with As<sup>3+</sup>, however compounds in which arsenic is present as a cation are extremely rare.<ref>[[#Burford|Burford & Royan 1989, p.&nbsp;3746]]</ref> AsPO<sub>4</sub>, 'which is virtually a covalent oxide,' has been referred to as a double oxide, of the form As<sub>2</sub>O<sub>3</sub>·P<sub>2</sub>O<sub>5</sub>. It comprises AsO<sub>3</sub> pyramids and PO<sub>4</sub> tetrahedra, joined together by all their corner atoms to form a continuous polymeric network.<ref>[[#Corbridge|Corbridge 2012, pp.&nbsp;122, 215]]</ref> As<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub> has a structure in which each SO<sub>4</sub> tetrahedron is bridged by two AsO<sub>3</sub> trigonal pyramida.<ref>[[#Douglade|Douglade 1982]]</ref>|group=n}} and arsenic acetate As(CH<sub>3</sub>COO)<sub>3</sub>.<ref>[[#Zingaro1994|Zingaro 1994, p.&nbsp;197]]; [[#Emeleús1959|Emeleús & Sharpe 1959, p.&nbsp;418]]; [[#Addison1972|Addison & Sowerby 1972, p.&nbsp;209]]; [[#Mellor1964|Mellor 1964, p.&nbsp;337]]</ref> The oxide As<sub>2</sub>O<sub>3</sub> is polymeric,<ref name=Pudd59/> amphoteric,<ref>[[#Pourbaix1974|Pourbaix 1974, p.&nbsp;521]]; [[#Eagleson1994|Eagleson 1994, p.&nbsp;92]]; [[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;572]]</ref>{{#tag:ref|Whilst As<sub>2</sub>O<sub>3</sub> is usually regarded as being amphoteric a few sources instead say it is (weakly)<ref>[[#Wiberg2001|Wiberg 2001, pp.&nbsp;750, 975]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;314]]</ref> acidic. They describe its 'basic' properties (that is, its reaction with concentrated hydrochloric acid to form arsenic trichloride) as being alcoholic, by analogy with the formation of covalent alklyl chlorides by covalent alcohols (e.g., R-OH + HCl <big>→</big> RCl + H<sub>2</sub>O)<ref>[[#Sidgwick1950|Sidgwick 1950, p.&nbsp;784]]; [[#Moody1991|Moody 1991, pp.&nbsp;248–9, 319]]</ref>|group=n}} and a glass former.<ref name=Rao22/> Arsenic has an extensive organometallic chemistry (see [[organoarsenic chemistry]]).<ref>[[#Krannich2006|Krannich & Watkins 2006]]</ref>
The chemistry of arsenic is predominately nonmetallic,<ref>[[#Bailar1965|Bailar, Moeller & Kleinberg 1965, p.&nbsp;477]]</ref> and with many metals it forms alloys; most of these are brittle.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;91]]</ref> Arsenic shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox/> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Arsenic generally forms compounds in which it has an oxidation state of +3 or +5.<ref name="Massey267">[[#Massey2000|Massey 2000, p.&nbsp;267]]</ref> The halides, and the oxides and their derivatives are illustrative examples.<ref name="Bailar513">[[#Bailar1965|Bailar, Moeller & Kleinberg 1965, p.&nbsp;513]]</ref> In the trivalent state, arsenic shows some incipient metallic properties.<ref>[[#Timm1944|Timm 1944, p.&nbsp;454]]</ref> The halides are hydrolysed by water but these reactions, particularly those of the chloride, are reversible with the addition of a [[hydrohalic acid]].<ref>[[#Partington1944|Partington 1944, p.&nbsp;641]]; [[#Kleinberg1960|Kleinberg, Argersinger & Griswold 1960, p.&nbsp;419]]</ref> The oxide is acidic but weakly amphoteric. The higher, less stable, pentavalent state has strongly acidic (nonmetallic) properties.<ref>[[#Morgan1906|Morgan 1906, p.&nbsp;163]]; [[#Moeller1954|Moeller 1954, p.&nbsp;559]]</ref> Compared to phosphorus, the stronger metallic character of arsenic is indicated by the formation of oxoacid salts such as AsPO<sub>4</sub>, As<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub>{{#tag:ref|The formulae of AsPO<sub>4</sub> and As<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub> suggest straightforward ionic formulations, with As<sup>3+</sup>, but compounds in which arsenic is present as a cation are extremely rare.<ref>[[#Burford|Burford & Royan 1989, p.&nbsp;3746]]</ref> AsPO<sub>4</sub>, 'which is virtually a covalent oxide,' has been referred to as a double oxide, of the form As<sub>2</sub>O<sub>3</sub>·P<sub>2</sub>O<sub>5</sub>. It comprises AsO<sub>3</sub> pyramids and PO<sub>4</sub> tetrahedra, joined together by all their corner atoms to form a continuous polymeric network.<ref>[[#Corbridge|Corbridge 2012, pp.&nbsp;122, 215]]</ref> As<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub> has a structure in which each SO<sub>4</sub> tetrahedron is bridged by two AsO<sub>3</sub> trigonal pyramida.<ref>[[#Douglade|Douglade 1982]]</ref>|group=n}} and arsenic acetate As(CH<sub>3</sub>COO)<sub>3</sub>.<ref>[[#Zingaro1994|Zingaro 1994, p.&nbsp;197]]; [[#Emeleús1959|Emeleús & Sharpe 1959, p.&nbsp;418]]; [[#Addison1972|Addison & Sowerby 1972, p.&nbsp;209]]; [[#Mellor1964|Mellor 1964, p.&nbsp;337]]</ref> The oxide As<sub>2</sub>O<sub>3</sub> is polymeric,<ref name=Pudd59/> amphoteric,<ref>[[#Pourbaix1974|Pourbaix 1974, p.&nbsp;521]]; [[#Eagleson1994|Eagleson 1994, p.&nbsp;92]]; [[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;572]]</ref>{{#tag:ref|As<sub>2</sub>O<sub>3</sub> is usually regarded as being amphoteric but a few sources say it is (weakly)<ref>[[#Wiberg2001|Wiberg 2001, pp.&nbsp;750, 975]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;314]]</ref> acidic. They describe its 'basic' properties (that is, its reaction with concentrated hydrochloric acid to form arsenic trichloride) as being alcoholic, by analogy with the formation of covalent alklyl chlorides by covalent alcohols (e.g., R-OH + HCl <big>→</big> RCl + H<sub>2</sub>O)<ref>[[#Sidgwick1950|Sidgwick 1950, p.&nbsp;784]]; [[#Moody1991|Moody 1991, pp.&nbsp;248–9, 319]]</ref>|group=n}} and a glass former.<ref name=Rao22/> Arsenic has an extensive organometallic chemistry (see [[organoarsenic chemistry]]).<ref>[[#Krannich2006|Krannich & Watkins 2006]]</ref>


===Antimony===
===Antimony===
{{main|Antimony}}
{{main|Antimony}}
[[File:Antimony-4.jpg|thumb|right|Antimony, showing its brilliant lustre|alt=A glistening silver rock-like chunk, with a blue tint, and roughly parallel furrows.]]
[[File:Antimony-4.jpg|thumb|right|Antimony, showing its brilliant lustre|alt=A glistening silver rock-like chunk, with a blue tint, and roughly parallel furrows.]]
Antimony appears as a silver-white solid with a blue tint and a brilliant lustre.<ref name="Greenwood 2002, p. 552"/> It is about 15% less dense than iron, brittle, and moderately hard (more so than arsenic; less so than iron; about the same as copper).<ref name="GWM2011"/> It is stable in air, and moisture, at room temperature. It is attacked by: concentrated nitric acid, yielding the hydrated pentoxide Sb<sub>2</sub>O<sub>5</sub>; [[aqua regia]], giving the pentachloride SbCl<sub>5</sub>; and (hot) concentrated sulfuric acid, resulting in the sulfate Sb<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub>.<ref name="Greenwood 2002, p. 553">[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;553]]</ref> It is not affected by molten alkali.<ref>[[#Dunstan1968|Dunstan 1968, p.&nbsp;433]]</ref> Antimony is capable of displacing hydrogen from water, when heated: 2Sb + 3 H<sub>2</sub>O → Sb<sub>2</sub>O<sub>3</sub> + 3 H<sub>2</sub>.<ref>[[#Parise1996|Parise 1996, p.&nbsp;112]]</ref> It melts at a temperature of 630.63 °C, which is approximately half that of steel. Antimony is a semimetal with an electrical conductivity of around 3.1 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup><ref>[[#Carapella1968a|Carapella 1968a, p.&nbsp;23]]</ref> and a band overlap of 0.16&nbsp;eV.<ref name="Barfuß 1981, p. 967"/>{{#tag:ref|Antimony can also be prepared in an [[amorphous solid|amorphous]] semiconducting black form, with an estimated (temperature-dependent) [[band gap]] of 0.06–0.18&nbsp;eV.<ref>[[#Moss1952|Moss 1952, pp.&nbsp;174, 179]]</ref>|group=n}} Liquid antimony is a metallic conductor with an electrical conductivity of around 5.3 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup>.<ref>[[#Dupree1982|Dupree, Kirby & Freyland 1982, p.&nbsp;604]]; [[#Mhiaoui2003|Mhiaoui, Sar, & Gasser 2003]]</ref>
Antimony is a silver-white solid with a blue tint and a brilliant lustre.<ref name="Greenwood 2002, p. 552"/> It is about 15% less dense than iron, brittle, and moderately hard (more so than arsenic; less so than iron; about the same as copper).<ref name="GWM2011"/> It is stable in air and moisture at room temperature. It is attacked by concentrated nitric acid, yielding the hydrated pentoxide Sb<sub>2</sub>O<sub>5</sub>. [[Aqua regia]] gives the pentachloride SbCl<sub>5</sub> and (hot) concentrated sulfuric acid results in the sulfate Sb<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub>.<ref name="Greenwood 2002, p. 553">[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;553]]</ref> It is not affected by molten alkali.<ref>[[#Dunstan1968|Dunstan 1968, p.&nbsp;433]]</ref> Antimony is capable of displacing hydrogen from water, when heated: 2Sb + 3 H<sub>2</sub>O → Sb<sub>2</sub>O<sub>3</sub> + 3 H<sub>2</sub>.<ref>[[#Parise1996|Parise 1996, p.&nbsp;112]]</ref> It melts at 630.63 °C. Antimony is a semimetal with an electrical conductivity of around 3.1 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup><ref>[[#Carapella1968a|Carapella 1968a, p.&nbsp;23]]</ref> and a band overlap of 0.16&nbsp;eV.<ref name="Barfuß 1981, p. 967"/>{{#tag:ref|Antimony can also be prepared in an [[amorphous solid|amorphous]] semiconducting black form, with an estimated (temperature-dependent) band gap of 0.06–0.18&nbsp;eV.<ref>[[#Moss1952|Moss 1952, pp.&nbsp;174, 179]]</ref>|group=n}} Liquid antimony is a metallic conductor with an electrical conductivity of around 5.3 × 10<sup>4</sup>&nbsp;S•cm<sup>−1</sup>.<ref>[[#Dupree1982|Dupree, Kirby & Freyland 1982, p.&nbsp;604]]; [[#Mhiaoui2003|Mhiaoui, Sar, & Gasser 2003]]</ref>


Most of the chemistry of antimony is characteristic of a nonmetal.<ref>[[#Kotz2009|Kotz, Treichel & Weaver 2009, p.&nbsp;62]]</ref> It does however form alloys with one or more metals such as aluminium,<ref>[[#Friend1953|Friend 1953, p.&nbsp;87]]</ref> iron, nickel, copper, zinc, tin, lead and bismuth.<ref>[[#Fesquet1872|Fesquet 1872, pp.&nbsp;109–14]]</ref> Antimony shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox/> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Like arsenic, antimony generally forms compounds in which it has an oxidation state of +3 or +5.<ref name=Massey267/> The halides, and the oxides and their derivatives are illustrative examples.<ref name=Bailar513/> The +5 state is less stable than the +3, but relatively easier to attain than is the case with arsenic. This is on account of the poor shielding afforded the arsenic nucleus by its 3d<sup>10</sup> electrons. In comparison, the tendency of antimony to [[redox|oxidize]] more easily partially offsets the effect of its 4d<sup>10</sup> shell.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;553]]; [[#Massey2000|Massey 2000, p.&nbsp;269]]</ref> Tripositive antimony is amphoteric; [[penta-|pentapositive]] antimony is (predominately) acidic.<ref>[[#King1994|King 1994,&nbsp;p.171]]</ref> Consistent with an increase in metallic character down [[pnictogen|group 15]], antimony forms salts or salt-like compounds including a nitrate Sb(NO<sub>3</sub>)<sub>3</sub>, phosphate SbPO<sub>4</sub>, sulfate Sb<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub> and perchlorate Sb(ClO<sub>4</sub>)<sub>3</sub>.<ref>[[#Turova2011|Torova 2011, p.&nbsp;46]]</ref> The otherwise acidic pentoxide Sb<sub>2</sub>O<sub>5</sub> shows some basic (metallic) behaviour in that it can be dissolved in very acidic solutions, with the formation of the oxycation SbO{{su|b=2|p=+}}.<ref>[[#Pourbaix1974|Pourbaix 1974, p.&nbsp;530]]</ref> The oxide Sb<sub>2</sub>O<sub>3</sub> is a polymeric,<ref name=Pudd59/> amphoteric,<ref name="Wiberg2001p764">[[#Wiberg2001|Wiberg 2001, p.&nbsp;764]]</ref> and a glass former.<ref name=Rao22/> Antimony has an extensive organometallic chemistry (see [[organoantimony chemistry]]).<ref>[[#House2008|House 2008, p.&nbsp;497]]</ref>
Most of the chemistry of antimony is characteristic of a nonmetal.<ref>[[#Kotz2009|Kotz, Treichel & Weaver 2009, p.&nbsp;62]]</ref> It forms alloys with one or more metals such as aluminium,<ref>[[#Friend1953|Friend 1953, p.&nbsp;87]]</ref> iron, nickel, copper, zinc, tin, lead and bismuth.<ref>[[#Fesquet1872|Fesquet 1872, pp.&nbsp;109–14]]</ref> Antimony has fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox/> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Like arsenic, antimony generally forms compounds in which it has an oxidation state of +3 or +5.<ref name=Massey267/> The halides, and the oxides and their derivatives are illustrative examples.<ref name=Bailar513/> The +5 state is less stable than the +3, but relatively easier to attain than with arsenic. This is explained by the poor shielding afforded the arsenic nucleus by its 3d<sup>10</sup> electrons. In comparison, the tendency of antimony to [[redox|oxidize]] more easily partially offsets the effect of its 4d<sup>10</sup> shell.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;553]]; [[#Massey2000|Massey 2000, p.&nbsp;269]]</ref> Tripositive antimony is amphoteric; [[penta-|pentapositive]] antimony is (predominately) acidic.<ref>[[#King1994|King 1994,&nbsp;p.171]]</ref> Consistent with an increase in metallic character down [[pnictogen|group 15]], antimony forms salts or salt-like compounds including a nitrate Sb(NO<sub>3</sub>)<sub>3</sub>, phosphate SbPO<sub>4</sub>, sulfate Sb<sub>2</sub>(SO<sub>4</sub>)<sub>3</sub> and perchlorate Sb(ClO<sub>4</sub>)<sub>3</sub>.<ref>[[#Turova2011|Torova 2011, p.&nbsp;46]]</ref> The otherwise acidic pentoxide Sb<sub>2</sub>O<sub>5</sub> shows some basic (metallic) behaviour in that it can be dissolved in very acidic solutions, with the formation of the oxycation SbO{{su|b=2|p=+}}.<ref>[[#Pourbaix1974|Pourbaix 1974, p.&nbsp;530]]</ref> The oxide Sb<sub>2</sub>O<sub>3</sub> is a polymeric,<ref name=Pudd59/> amphoteric,<ref name="Wiberg2001p764">[[#Wiberg2001|Wiberg 2001, p.&nbsp;764]]</ref> and a glass former.<ref name=Rao22/> Antimony has an extensive organometallic chemistry (see [[organoantimony chemistry]]).<ref>[[#House2008|House 2008, p.&nbsp;497]]</ref>


===Tellurium===
===Tellurium===
{{main|Tellurium}}
{{main|Tellurium}}
[[File:Tellurium2.jpg|thumb|left|Tellurium, described by [[Mendeleev]] as forming a transition between metals and nonmetals<ref>[[#Mendeléeff1897a|Mendeléeff 1897, p.&nbsp;274]]</ref>|alt=A shiny silver-white medallion with a striated surface, irregular around the outside, with a square spiral-like pattern in the middle.]]
[[File:Tellurium2.jpg|thumb|left|Tellurium, described by [[Dmitri Mendeleev]] as forming a transition between metals and nonmetals<ref>[[#Mendeléeff1897a|Mendeléeff 1897, p.&nbsp;274]]</ref>|alt=A shiny silver-white medallion with a striated surface, irregular around the outside, with a square spiral-like pattern in the middle.]]
Tellurium appears as a silvery-white solid with a shiny lustre.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;428]]</ref> It is about 15% less dense than iron, brittle, and the softest of the commonly recognised metalloids, being marginally harder than sulfur.<ref name="GWM2011"/> Massive tellurium is 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&nbsp;°C, to give the dioxide and hydrogen: Te + 2 H<sub>2</sub>O → TeO<sub>2</sub> + 2 H<sub>2</sub>.<ref name=Kudryavtsev78>[[#Kudryavtsev1974|Kudryavtsev 1974, p.&nbsp;78]]</ref> It reacts (to varying degrees) with, or combinations of, nitric, sulfuric and hydrochloric acids to give compounds such as the sulfoxide TeSO<sub>3</sub> or tellurous acid H<sub>2</sub>TeO<sub>3</sub>,<ref>[[#Bagnall1966|Bagnall 1966, pp.&nbsp;32–3, 59, 137]]</ref> the basic nitrate (Te<sub>2</sub>O<sub>4</sub>H)<sup>+</sup>(NO<sub>3</sub>)<sup>–</sup>,<ref>[[#Swink1966|Swink et al. 1966]]; [[#Anderson1980|Anderson et al. 1980]]</ref> or the oxide sulfate Te<sub>2</sub>O<sub>3</sub>(SO<sub>4</sub>).<ref>[[#Ahmed2000|Ahmed, Fjellvåg & Kjekshus 2000]]</ref> It dissolves in boiling alkalis, with the formation of the tellurite and telluride: 3 Te + 6 KOH = K<sub>2</sub>TeO<sub>3</sub> + 2 K<sub>2</sub>Te +3 H<sub>2</sub>O, a reaction that proceeds or is reversible with increasing or decreasing temperature.<ref>[[#Chizhikov1970|Chizhikov & Shchastlivyi 1970, p.&nbsp;28]]</ref> At higher temperatures tellurium is sufficiently plastic to extrude.<ref>[[#Kudryavtsev1974|Kudryavtsev 1974, p.&nbsp;77]]</ref> It melts at much lower temperatures than steel – ca. 450 °C vs. ca. 1400 °C.<ref name=steelMP>[http://www.britannica.com/EBchecked/topic/564627/steel steel (metallurgy)]. Encyclopaedia Britannica.</ref> Crystalline tellurium has a structure consisting of parallel infinite spiral chains. Whereas the bonding between adjacent atoms in a chain is covalent, there is evidence of a weak metallic interaction between the neighbouring atoms of different chains.<ref name="Stuke1074p178">[[#Stuke1974|Stuke 1974, p.&nbsp;178]]; [[#Donohue1982|Donohue 1982, pp.&nbsp;386–7]]; [[#Cotton1999|Cotton et al. 1999, p.&nbsp;501]]</ref> Tellurium is a semiconductor with an (intrinsic) electrical conductivity of around 1.0&nbsp;S•cm<sup>−1</sup><ref>[[#Becker1971|Becker, Johnson & Nussbaum 1971, p.&nbsp;56]]</ref> and a band gap of 0.32 to 0.38&nbsp;eV.<ref name=Berger90>[[#Berger1997|Berger 1997, p.&nbsp;90]]</ref> Liquid tellurium is a semiconductor, with an electrical conductivity, on melting, of around 1.9 × 10<sup>3</sup>&nbsp;S•cm<sup>−1</sup><ref name=Berger90/> Superheated liquid tellurium is a metallic conductor.<ref>[[#Chizhikov1970|Chizhikov & Shchastlivyi 1970, p.&nbsp;16]]</ref>
Tellurium is a silvery-white shiny solid.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;428]]</ref> It is about 15% less dense than iron, brittle, and the softest of the commonly recognised metalloids, being marginally harder than sulfur.<ref name="GWM2011"/> 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&nbsp;°C, to give the dioxide and hydrogen: Te + 2 H<sub>2</sub>O → TeO<sub>2</sub> + 2 H<sub>2</sub>.<ref name=Kudryavtsev78>[[#Kudryavtsev1974|Kudryavtsev 1974, p.&nbsp;78]]</ref> It reacts (to varying degrees) with nitric, sulfuric and hydrochloric acids to give compounds such as the sulfoxide TeSO<sub>3</sub> or tellurous acid H<sub>2</sub>TeO<sub>3</sub>,<ref>[[#Bagnall1966|Bagnall 1966, pp.&nbsp;32–3, 59, 137]]</ref> the basic nitrate (Te<sub>2</sub>O<sub>4</sub>H)<sup>+</sup>(NO<sub>3</sub>)<sup>–</sup>,<ref>[[#Swink1966|Swink et al. 1966]]; [[#Anderson1980|Anderson et al. 1980]]</ref> or the oxide sulfate Te<sub>2</sub>O<sub>3</sub>(SO<sub>4</sub>).<ref>[[#Ahmed2000|Ahmed, Fjellvåg & Kjekshus 2000]]</ref> It dissolves in boiling alkalis, with the formation of the tellurite and telluride: 3 Te + 6 KOH = K<sub>2</sub>TeO<sub>3</sub> + 2 K<sub>2</sub>Te +3 H<sub>2</sub>O, a reaction that proceeds or is reversible with increasing or decreasing temperature.<ref>[[#Chizhikov1970|Chizhikov & Shchastlivyi 1970, p.&nbsp;28]]</ref> At higher temperatures tellurium is sufficiently plastic to extrude.<ref>[[#Kudryavtsev1974|Kudryavtsev 1974, p.&nbsp;77]]</ref> It melts at ca. 450 °C. Crystalline tellurium has a structure consisting of parallel infinite spiral chains. Whereas the bonding between adjacent atoms in a chain is covalent, there is evidence of a weak metallic interaction between the neighbouring atoms of different chains.<ref name="Stuke1074p178">[[#Stuke1974|Stuke 1974, p.&nbsp;178]]; [[#Donohue1982|Donohue 1982, pp.&nbsp;386–7]]; [[#Cotton1999|Cotton et al. 1999, p.&nbsp;501]]</ref> Tellurium is a semiconductor with an (intrinsic) electrical conductivity of around 1.0&nbsp;S•cm<sup>−1</sup><ref>[[#Becker1971|Becker, Johnson & Nussbaum 1971, p.&nbsp;56]]</ref> and a band gap of 0.32 to 0.38&nbsp;eV.<ref name=Berger90>[[#Berger1997|Berger 1997, p.&nbsp;90]]</ref> Liquid tellurium is a semiconductor, with an electrical conductivity, on melting, of around 1.9 × 10<sup>3</sup>&nbsp;S•cm<sup>−1</sup><ref name=Berger90/> Superheated liquid tellurium is a metallic conductor.<ref>[[#Chizhikov1970|Chizhikov & Shchastlivyi 1970, p.&nbsp;16]]</ref>


Most of the chemistry of tellurium is characteristic of a nonmetal.<ref>[[#Jolly1966|Jolly 1966, pp.&nbsp;66–7]]</ref> It does however form alloys with, for example, aluminium, silver and tin.<ref>[[#Mellor1964a|Mellor 1964, p.&nbsp; 30]]; [[#Wiberg2001|Wiberg 2001, p.&nbsp;589]]</ref> Tellurium shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox /> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Tellurium generally forms compounds in which it has an oxidation state of −2, +4 or +6, with the tetrapositive state being the most stable.<ref name=Kudryavtsev78/> It combines easily with most other elements to form binary tellurides X<sub>''x''</sub>Te<sub>''y''</sub> these representing the most common mineral form. Non-stoichiometry is frequently encountered. This is particularly so with the transition metals, where [[electronegativity]] differences are small and irregular valence is favoured. Many of the associated tellurides can be treated as metallic alloys.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;765–6]]</ref> The increase in metallic character evident in tellurium, as compared to the lighter [[chalcogen]]s, is further reflected in the reported formation of various other oxyacid salts, such as a basic selenate 2TeO<sub>2</sub>·SeO<sub>3</sub> and an analogous perchlorate and periodate 2TeO<sub>2</sub>·HXO<sub>4</sub>.<ref>[[#Bagnall1966|Bagnall 1966, p.&nbsp;134–51]]; [[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;786]]</ref> Tellurium forms a polymeric,<ref name=Pudd59/> amphoteric,<ref name="Wiberg2001p764"/> glass-forming oxide<ref name=Rao22/> TeO<sub>2</sub>. The latter is a 'conditional' glass-forming oxide—it forms a glass with a very small amount of additive.<ref name=Rao22/> Tellurium has an extensive organometallic chemistry (see [[organotellurium chemistry]]).<ref>[[#Detty1994|Detty & O'Regan 1994, pp.&nbsp;1–2]]</ref>
Most of the chemistry of tellurium is characteristic of a nonmetal.<ref>[[#Jolly1966|Jolly 1966, pp.&nbsp;66–7]]</ref> It forms alloys with aluminium, silver and tin.<ref>[[#Mellor1964a|Mellor 1964, p.&nbsp; 30]]; [[#Wiberg2001|Wiberg 2001, p.&nbsp;589]]</ref> Tellurium shows fewer tendencies to anionic behaviour than ordinary nonmetals.<ref name=Cox /> Its solution chemistry is characterised by the formation of oxyanions.<ref name=Hiller225/> Tellurium generally forms compounds in which it has an oxidation state of −2, +4 or +6, with the tetrapositive state being the most stable.<ref name=Kudryavtsev78/> It combines easily with most other elements to form binary tellurides X<sub>''x''</sub>Te<sub>''y''</sub> these representing the most common mineral form. Non-stoichiometry is frequently encountered. This is particularly so with the transition metals, where electronegativity differences are small and irregular valence is favoured. Many of the associated tellurides can be treated as metallic alloys.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;765–6]]</ref> The increase in metallic character evident in tellurium, as compared to the lighter [[chalcogen]]s, is further reflected in the reported formation of various other oxyacid salts, such as a basic selenate 2TeO<sub>2</sub>·SeO<sub>3</sub> and an analogous perchlorate and periodate 2TeO<sub>2</sub>·HXO<sub>4</sub>.<ref>[[#Bagnall1966|Bagnall 1966, p.&nbsp;134–51]]; [[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;786]]</ref> Tellurium forms a polymeric,<ref name=Pudd59/> amphoteric,<ref name="Wiberg2001p764"/> glass-forming oxide<ref name=Rao22/> TeO<sub>2</sub>. The latter is a 'conditional' glass-forming oxide—it forms a glass with a very small amount of additive.<ref name=Rao22/> Tellurium has an extensive organometallic chemistry (see [[organotellurium chemistry]]).<ref>[[#Detty1994|Detty & O'Regan 1994, pp.&nbsp;1–2]]</ref>


==Elements less commonly recognised as metalloids==
==Elements less commonly recognised as metalloids==
:''This section presents brief sketches of the physical and chemical properties of the applicable elements—in their most thermodynamically stable forms under ambient conditions. For complete profiles, including history, production, and specific uses, see the main article for each element.''


===Carbon===
===Carbon===
{{main|Carbon}}
{{main|Carbon}}
[[File:Graphite2.jpg|thumb|right|Carbon (as graphite). Delocalized valence electrons within the layers of graphite give it a metallic appearance.<ref>[[#Hill2000|Hill & Holman 2000, p.&nbsp;124]]</ref>|alt=A shiny grey-black cuboid nugget with a rough surface.]]
[[File:Graphite2.jpg|thumb|right|Carbon (as graphite). Delocalized valence electrons within the layers of graphite give it a metallic appearance.<ref>[[#Hill2000|Hill & Holman 2000, p.&nbsp;124]]</ref>|alt=A shiny grey-black cuboid nugget with a rough surface.]]
Carbon is ordinarily classified as a nonmetal<ref>[[#Chang2002|Chang 2002, p.&nbsp;314]]</ref> although it has some metallic properties and is occasionally classified as a metalloid.<ref>[[#Kent1950|Kent 1950, pp.&nbsp;1–2]]; [[#Clark1960|Clark 1960, p.&nbsp;588]]; [[#Warren1981|Warren & Geballe 1981]]</ref> As applicable, the properties summarised in the following paragraphs are for hexagonal graphitic [[Graphite|carbon]], the most thermodynamically stable form of carbon under ambient conditions.<ref>[[#Housecroft2008|Housecroft & Sharpe 2008, p.&nbsp;384]]; [[#IUPAC2006|IUPAC 2006–, rhombohedral graphite entry]]</ref>
Carbon is ordinarily classified as a nonmetal<ref>[[#Chang2002|Chang 2002, p.&nbsp;314]]</ref> but has some metallic properties and is occasionally classified as a metalloid.<ref>[[#Kent1950|Kent 1950, pp.&nbsp;1–2]]; [[#Clark1960|Clark 1960, p.&nbsp;588]]; [[#Warren1981|Warren & Geballe 1981]]</ref> The properties summarised in the following paragraphs are for hexagonal graphitic [[Graphite|carbon]], the most thermodynamically stable form of carbon under ambient conditions.<ref>[[#Housecroft2008|Housecroft & Sharpe 2008, p.&nbsp;384]]; [[#IUPAC2006|IUPAC 2006–, rhombohedral graphite entry]]</ref>


In terms of metallic character, carbon has a lustrous appearance<ref>[[#Mingos1998|Mingos 1998, p.&nbsp;171]]</ref> and is a fairly good electrical conductor.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;781]]</ref> Its conductivity in the direction of its planes decreases as the temperature is raised, behaving in this way as a metal;<ref name="Atkins320">[[#Atkins2006|Atkins et al. 2006, pp.&nbsp;320–1]]</ref>{{#tag:ref|Liquid carbon may<ref>[[#Savvatimskiy2005|Savvatimskiy 2005, p.&nbsp;1138]]</ref> or may not<ref>[[#Togaya2000|Togaya 2000]]</ref> be a metallic conductor, depending on pressure and temperature; see also.<ref>[[#Savvatimskiy2009|Savvatimskiy 2009]]</ref>|group=n}} it actually has the electronic band structure of a semimetal.<ref name=Atkins320/> The various allotropes of carbon, including graphite, are capable of accepting foreign atoms or compounds into their structures via substitution, [[Intercalation (chemistry)|intercalation]] or doping ([[Interstitial compound|interstitial]] or intrastitial) with the resulting materials being referred to as 'carbon alloys'.<ref>[[#Inagaki2000|Inagaki 2000, p.&nbsp;216]]; [[#Yasuda2003|Yasuda et al. 2003, pp.&nbsp;3–11]]</ref> Carbon can form ionic salts, including a sulfate, perchlorate, nitrate, hydrogen selenate, and hydrogen phosphate;<ref name=Wiberg795>[[#Wiberg2001|Wiberg 2001, p.&nbsp;795]]</ref>{{#tag:ref|For the sulfate, the method of preparation is (careful) direct oxidation of graphite in concentrated sulfuric acid by an 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]]. Analogous salts are formed in other strong acids, such as [[perchloric acid]].<ref name=Wiberg795/>|group=n}} in [[organic chemistry]], carbon can form complex cations—termed [[carbocation|''carbocations'']]—in which the positive charge is on the carbon atom; examples are [[Carbenium ion|{{chem|CH|3|+}}]] and [[Carbonium ion|{{chem|CH|5|+}}]], and their derivatives.<ref>[[#Traynham1989|Traynham 1989, pp.&nbsp;930–1]]; [[#Prakash1997|Prakash & Schleyer 1997]]</ref>
Carbon has a lustrous appearance<ref>[[#Mingos1998|Mingos 1998, p.&nbsp;171]]</ref> and is a fairly good electrical conductor.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;781]]</ref> Its conductivity in the direction of its planes decreases as the temperature is raised, like a metal;<ref name="Atkins320">[[#Atkins2006|Atkins et al. 2006, pp.&nbsp;320–1]]</ref>{{#tag:ref|Liquid carbon may<ref>[[#Savvatimskiy2005|Savvatimskiy 2005, p.&nbsp;1138]]</ref> or may not<ref>[[#Togaya2000|Togaya 2000]]</ref> be a metallic conductor, depending on pressure and temperature; see also.<ref>[[#Savvatimskiy2009|Savvatimskiy 2009]]</ref>|group=n}} it has the electronic band structure of a semimetal.<ref name=Atkins320/> The allotropes of carbon, including graphite, can accept foreign atoms or compounds into their structures via substitution, [[Intercalation (chemistry)|intercalation]] or doping ([[Interstitial compound|interstitial]] or intrastitial) with the resulting materials being referred to as 'carbon alloys'.<ref>[[#Inagaki2000|Inagaki 2000, p.&nbsp;216]]; [[#Yasuda2003|Yasuda et al. 2003, pp.&nbsp;3–11]]</ref> Carbon can form ionic salts, including a sulfate, perchlorate, nitrate, hydrogen selenate, and hydrogen phosphate;<ref name=Wiberg795>[[#Wiberg2001|Wiberg 2001, p.&nbsp;795]]</ref>{{#tag:ref|For the sulfate, the method of preparation is (careful) direct oxidation of graphite in concentrated sulfuric acid by an 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]]. Analogous salts are formed in other strong acids, such as [[perchloric acid]].<ref name=Wiberg795/>|group=n}} in [[organic chemistry]], carbon can form complex cations—termed [[carbocation|''carbocations'']]—in which the positive charge is on the carbon atom; examples are [[Carbenium ion|{{chem|CH|3|+}}]] and [[Carbonium ion|{{chem|CH|5|+}}]], and their derivatives.<ref>[[#Traynham1989|Traynham 1989, pp.&nbsp;930–1]]; [[#Prakash1997|Prakash & Schleyer 1997]]</ref>


In terms of nonmetallic character, carbon is brittle<ref>[[#Olmsted1997|Olmsted & Williams 1997, p.&nbsp;436]]</ref> and behaves as a semiconductor perpendicular to the direction of its planes.<ref name=Atkins320/> Most of its chemistry is nonmetallic;<ref>[[#Bailar1989|Bailar et al. 1989, p.&nbsp;743]]</ref> it has a relatively high ionization energy<ref>[[#Moore1985|Moore et al. 1985]]</ref> and, compared to most metals, a relatively high electronegativity.<ref>[[#House2010|House & House 2010, p.&nbsp;526]]</ref> Carbon can form anions such as C<sup>4–</sup> ([[methanide]]), C{{su|b=2|p=2–}} ([[acetylide]]) and C{{su|b=4|p=3–}} ([[Sesquicarbide|sesquicarbide or allylenide]]), in compounds with metals of main groups 1–3, and with the [[lanthanide]]s and [[actinide]]s.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;798]]</ref> Its oxide [[Carbon dioxide|CO<sub>2</sub>]] forms a medium-strength ''[[carbonic acid]]'' H<sub>2</sub>CO<sub>3</sub>.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;175]]</ref>{{#tag:ref|Only a very small fraction of dissolved CO<sub>2</sub> is present in water as carbonic acid so, even though H<sub>2</sub>CO<sub>3</sub> is actually a medium-strong acid, solutions of carbonic acid are only weakly acidic.<ref>[[#Atkins2006|Atkins et al. 2006, p.&nbsp;121]]</ref>|group=n}}
Carbon is brittle<ref>[[#Olmsted1997|Olmsted & Williams 1997, p.&nbsp;436]]</ref> like nonmetals and behaves as a semiconductor perpendicular to the direction of its planes.<ref name=Atkins320/> Most of its chemistry is nonmetallic;<ref>[[#Bailar1989|Bailar et al. 1989, p.&nbsp;743]]</ref> it has a relatively high ionization energy<ref>[[#Moore1985|Moore et al. 1985]]</ref> and, compared to most metals, a relatively high electronegativity.<ref>[[#House2010|House & House 2010, p.&nbsp;526]]</ref> Carbon can form anions such as C<sup>4–</sup> ([[methanide]]), C{{su|b=2|p=2–}} ([[acetylide]]) and C{{su|b=4|p=3–}} ([[Sesquicarbide|sesquicarbide or allylenide]]), in compounds with metals of main groups 1–3, and with the [[lanthanide]]s and [[actinide]]s.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;798]]</ref> Its oxide [[Carbon dioxide|CO<sub>2</sub>]] forms a medium-strength ''[[carbonic acid]]'' H<sub>2</sub>CO<sub>3</sub>.<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;175]]</ref>{{#tag:ref|Only a very small fraction of dissolved CO<sub>2</sub> is present in water as carbonic acid so, even though H<sub>2</sub>CO<sub>3</sub> is actually a medium-strong acid, solutions of carbonic acid are only weakly acidic.<ref>[[#Atkins2006|Atkins et al. 2006, p.&nbsp;121]]</ref>|group=n}}


===Aluminium===
===Aluminium===
{{main|Aluminium}}
{{main|Aluminium}}
[[File:Aluminium-4.jpg|thumb|left|Aluminium. People handling high purity Al for the first time often question if it really is aluminium since it is very much softer than familiar Al alloys.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;358–9]]</ref>|alt=A silvery white steam-iron shaped lump with semi-circular striations along the width of its top surface and rough furrows in the middle portion of its left edge.]]
[[File:Aluminium-4.jpg|thumb|left|People handling high purity aluminium for the first time often question if it really is aluminium since it is very much softer than its familiar alloys.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;358–9]]</ref>|alt=A silvery white steam-iron shaped lump with semi-circular striations along the width of its top surface and rough furrows in the middle portion of its left edge.]]
Aluminium is ordinarily classified as a metal. Features associated with this status include its lustre, malleability and ductility, high electrical and thermal conductivity and close-packed crystalline structure.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;358–60 et seq]]</ref>
Aluminium is classified as a metal. It is lustrous, malleable and ductile, has high electrical and thermal conductivity and a close-packed crystalline structure.<ref>[[#Russell2005|Russell & Lee 2005, pp.&nbsp;358–60 et seq]]</ref>


It does however have some properties that are unusual for a metal; taken together,<ref name="Metcalfe et al. 1974, p.,[object Object], ,[object Object],539">[[#Metcalfe1974|Metcalfe et al. 1974, p.&nbsp;539]]</ref> these properties are sometimes used as a basis to classify aluminium as a metalloid.<ref>[[#Cobb2005|Cobb & Fetterolf 2005, p.&nbsp;64]]; [[#Metcalfe1974|Metcalfe, Williams & Castka 1974, p.&nbsp;539]]</ref> Its crystalline structure shows some evidence of directional bonding.<ref>[[#Ogata2002|Ogata, Li & Yip 2002]]; [[#Boyer2004|Boyer et al. 2004, p.&nbsp;1023]]; [[#Russell2005|Russell & Lee 2005, p.&nbsp;359]]</ref> Although it forms an Al<sup>3+</sup> cation in some compounds, aluminium bonds covalently in most others.<ref>[[#Cooper1968|Cooper 1968, p.&nbsp;25]]; [[#Henderson2000|Henderson 2000, p.&nbsp;5]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;314]]</ref> Its oxide is amphoteric,<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;1014]]</ref> and a conditional glass-former.<ref name=Rao22/> Aluminium can form anionic [[aluminate]]s,<ref name="Metcalfe et al. 1974, p.,[object Object], ,[object Object],539" /> such behaviour being considered nonmetallic in character.<ref name="Hamm 1969, p.,[object Object], ,[object Object],653">[[#Hamm1969|Hamm 1969, p.&nbsp;653]]</ref>
It has some properties that are unusual for a metal; taken together,<ref name="Metcalfe et al. 1974, p.,[object Object], ,[object Object],539">[[#Metcalfe1974|Metcalfe et al. 1974, p.&nbsp;539]]</ref> these properties are sometimes used as a basis to classify aluminium as a metalloid.<ref>[[#Cobb2005|Cobb & Fetterolf 2005, p.&nbsp;64]]; [[#Metcalfe1974|Metcalfe, Williams & Castka 1974, p.&nbsp;539]]</ref> Its crystalline structure shows some evidence of directional bonding.<ref>[[#Ogata2002|Ogata, Li & Yip 2002]]; [[#Boyer2004|Boyer et al. 2004, p.&nbsp;1023]]; [[#Russell2005|Russell & Lee 2005, p.&nbsp;359]]</ref> Although it forms an Al<sup>3+</sup> cation in some compounds, aluminium bonds covalently in most others.<ref>[[#Cooper1968|Cooper 1968, p.&nbsp;25]]; [[#Henderson2000|Henderson 2000, p.&nbsp;5]]; [[#Silberberg2006|Silberberg 2006, p.&nbsp;314]]</ref> Its oxide is amphoteric,<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;1014]]</ref> and a conditional glass-former.<ref name=Rao22/> Aluminium can form anionic [[aluminate]]s,<ref name="Metcalfe et al. 1974, p.,[object Object], ,[object Object],539" /> such behaviour being considered nonmetallic in character.<ref name="Hamm 1969, p.,[object Object], ,[object Object],653">[[#Hamm1969|Hamm 1969, p.&nbsp;653]]</ref>


Stott<ref>[[#Stott1956|Stott 1956, p.&nbsp;100]]</ref> labels aluminium as weak metal. It has the physical properties of a good metal but some of the chemical properties of a nonmetal. Steele<ref>[[#Steele1966|Steele 1966, p.&nbsp;60]]</ref> notes the somewhat paradoxical chemical behaviour of aluminium. It resembles a weak metal with its amphoteric oxide and the covalent character of many of its compounds. Yet it is also a strongly [[Electronegativity#Electropositivity|electropositive]] metal, with a [[Table of standard electrode potentials|high negative]] [http://www.chemguide.co.uk/physical/redoxeqia/introduction.html electrode potential.]
Stott<ref>[[#Stott1956|Stott 1956, p.&nbsp;100]]</ref> labels aluminium as a weak metal. It has the physical properties of a metal but some of the chemical properties of a nonmetal. Steele<ref>[[#Steele1966|Steele 1966, p.&nbsp;60]]</ref> notes the paradoxical chemical behaviour of aluminium. It resembles a weak metal with its amphoteric oxide and the covalent character of many of its compounds. Yet it is also a strongly [[Electronegativity#Electropositivity|electropositive]] metal, with a [[Table of standard electrode potentials|high negative]] electrode potential.


The notion of aluminium as a metalloid is sometimes disputed<ref>[[#Daub1996|Daub & Seese 1996]], pp.&nbsp;70, 109: 'Aluminum is not a metalloid but a metal because it has mostly metallic properties.'; [[#Denniston2004|Denniston, Topping & Caret 2004, p.&nbsp;57]]: 'Note that aluminum (Al) is classified as a metal, not a metalloid.'; [[#Hasan2009|Hasan 2009, p.&nbsp;16]]: 'Aluminum does not have the characteristics of a metalloid but rather those of a metal.'</ref> given it has many metallic properties. Aluminium is therefore, arguably, an exception to the mnemonic that elements adjacent to the metal-nonmetal dividing line are metalloids.<ref>[[#Holt2007|Holt, Rinehart & Wilson c. 2007]]</ref>{{#tag:ref|A mnemonic that captures the elements commonly recognised as metalloids goes: ''Up, up-down, up-down, up<span style="white-space: nowrap">...</span>are the metalloids!''<ref>[[#Tuthill2011|Tuthill 2011]]</ref>|group=n}}
The notion of aluminium as a metalloid is sometimes disputed<ref>[[#Daub1996|Daub & Seese 1996]], pp.&nbsp;70, 109: 'Aluminum is not a metalloid but a metal because it has mostly metallic properties.'; [[#Denniston2004|Denniston, Topping & Caret 2004, p.&nbsp;57]]: 'Note that aluminum (Al) is classified as a metal, not a metalloid.'; [[#Hasan2009|Hasan 2009, p.&nbsp;16]]: 'Aluminum does not have the characteristics of a metalloid but rather those of a metal.'</ref> as it has many metallic properties. Aluminium is therefore, arguably, an exception to the mnemonic that elements adjacent to the metal-nonmetal dividing line are metalloids.<ref>[[#Holt2007|Holt, Rinehart & Wilson c. 2007]]</ref>{{#tag:ref|A mnemonic that captures the elements commonly recognised as metalloids goes: ''Up, up-down, up-down, up<span style="white-space: nowrap">...</span>are the metalloids!''<ref>[[#Tuthill2011|Tuthill 2011]]</ref>|group=n}}


===Selenium===
===Selenium===
{{main|Selenium}}
{{main|Selenium}}
[[File:Selenium black (cropped).jpg|thumb|right|Selenium. Being a [[photoconductivity|photoconductor]], grey selenium conducts electricity around 1,000 times better when light falls on it, a property used since the mid-1870s in various light-sensing applications.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;382]]</ref>|alt=A small glass jar filled with small dull gray concave shaped buttons. The pieces of selenium look like tiny mushrooms without their stems.]]
[[File:Selenium black (cropped).jpg|thumb|right|upright|Selenium. Being a [[photoconductivity|photoconductor]], grey selenium conducts electricity around 1,000 times better when light falls on it, a property used since the mid-1870s in various light-sensing applications.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;382]]</ref>|alt=A small glass jar filled with small dull grey concave buttons. The pieces of selenium look like tiny mushrooms without their stems.]]
Selenium shows borderline metalloid or nonmetal behaviour.<ref>[[#Young2010|Young et al. 2010, p.&nbsp;9]]; [[#Craig2003|Craig 2003, p.&nbsp;391]]. Selenium is included in this work on account of its 'near metalloidal' status.</ref>{{#tag:ref|[[Eugene G. Rochow|Rochow]],<ref>[[#Rochow1957|Rochow 1957]]</ref> who would later write his 1966 monograph ''The metalloids'',<ref>[[#Rochow1966|Rochow 1966]]</ref> commented that, 'In some respects selenium acts like a metalloid and tellurium certainly does'.|group=n}}
Selenium shows borderline metalloid or nonmetal behaviour.<ref>[[#Young2010|Young et al. 2010, p.&nbsp;9]]; [[#Craig2003|Craig 2003, p.&nbsp;391]]. Selenium is 'near metalloidal'.</ref>{{#tag:ref|[[Eugene G. Rochow|Rochow]],<ref>[[#Rochow1957|Rochow 1957]]</ref> who later wrote his 1966 monograph ''The metalloids'',<ref>[[#Rochow1966|Rochow 1966]]</ref> commented that, 'In some respects selenium acts like a metalloid and tellurium certainly does'.|group=n}}


Its most stable form, the grey trigonal allotrope, is sometimes called 'metallic' selenium. This is because its electrical conductivity is several orders of magnitude greater than that of the red monoclinic form.<ref>[[#Moss1952|Moss 1952, p.&nbsp;192]]</ref> The metallic character of selenium is further shown by its lustre<ref name="Glinka 1965, p.,[object Object], ,[object Object],356">[[#Glinka1965|Glinka 1965, p.&nbsp;356]]</ref> and its crystalline structure, the latter of which is thought to include weakly 'metallic' interchain bonding.<ref>[[#Evans1966|Evans 1966, pp.&nbsp;124–5]]</ref> Selenium can be drawn into thin threads, when molten.<ref>[[#Regnault1853|Regnault 1853, p.&nbsp;208]]</ref> It exhibits a reluctance to acquire 'the high positive oxidation numbers characteristic of nonmetals'.<ref>[[#Scott1962|Scott & Kanda 1962, p.&nbsp;311]]</ref> It can form cyclic polycations (such as Se{{su|b=8|p=2+}}) when dissolved in [[oleum]]s<ref>[[#Cotton1999|Cotton et al. 1999, pp.&nbsp;496, 503–4]]</ref> (an attribute it shares with sulfur and tellurium); and a hydrolysed cationic salt in the form of trihydroxoselenium (IV) perchlorate <span style="white-space: nowrap">[Se(OH)<sub>3</sub>]<sup>+</sup>·ClO{{su|b=4|p=–}}</sup>.<ref>[[#Arlman1939|Arlman 1939]]; [[#Bagnall1966|Bagnall 1966, pp.&nbsp;135, 142–3]]</ref></span>
Its most stable form, the grey trigonal allotrope, is sometimes called 'metallic' selenium. This is because its electrical conductivity is several orders of magnitude greater than that of the red monoclinic form.<ref>[[#Moss1952|Moss 1952, p.&nbsp;192]]</ref> The metallic character of selenium is further shown by its lustre<ref name="Glinka 1965, p.,[object Object], ,[object Object],356">[[#Glinka1965|Glinka 1965, p.&nbsp;356]]</ref> and its crystalline structure, the latter of which is thought to include weakly 'metallic' interchain bonding.<ref>[[#Evans1966|Evans 1966, pp.&nbsp;124–5]]</ref> Selenium can be drawn into thin threads when molten.<ref>[[#Regnault1853|Regnault 1853, p.&nbsp;208]]</ref> It shows reluctance to acquire 'the high positive oxidation numbers characteristic of nonmetals'.<ref>[[#Scott1962|Scott & Kanda 1962, p.&nbsp;311]]</ref> It can form cyclic polycations (such as Se{{su|b=8|p=2+}}) when dissolved in [[oleum]]s<ref>[[#Cotton1999|Cotton et al. 1999, pp.&nbsp;496, 503–4]]</ref> (an attribute it shares with sulfur and tellurium); and a hydrolysed cationic salt in the form of trihydroxoselenium (IV) perchlorate <span style="white-space: nowrap">[Se(OH)<sub>3</sub>]<sup>+</sup>·ClO{{su|b=4|p=–}}</sup>.<ref>[[#Arlman1939|Arlman 1939]]; [[#Bagnall1966|Bagnall 1966, pp.&nbsp;135, 142–3]]</ref></span>


The nonmetallic character of selenium is shown by its brittleness<ref name="Glinka 1965, p.,[object Object], ,[object Object],356" /> and the low electrical conductivity (~10<sup>−9</sup> to 10<sup>−12</sup>&nbsp;S•cm<sup>−1</sup>) of its highly purified form.<ref name="Kozyrev" /> This is comparable to or less than that of [[bromine]] (7.95{{e|–12}}&nbsp;S•cm<sup>−1</sup>),<ref>[[#Chao1964|Chao & Stenger 1964]]</ref> a nonmetal. Selenium has the electronic band structure of a semiconductor<ref name="Berger 1997, pp.,[object Object], ,[object Object],86–7">[[#Berger1997|Berger 1997, pp.&nbsp;86–7]]</ref> and retains its semiconducting properties in liquid form.<ref name="Berger 1997, pp.,[object Object], ,[object Object],86–7" /> It has a relatively high<ref>[[#Snyder1966|Snyder 1966, p.&nbsp;242]]</ref> electronegativity (2.55 revised [[Linus Pauling|Pauling]] scale). Its reaction chemistry is mainly that of its nonmetallic anionic forms Se<sup>2–</sup>, SeO{{su|b=3|p=2−}} and SeO{{su|b=4|p=2−}}.<ref>[[#Fritz2008|Fritz & Gjerde 2008, p.&nbsp;235]]</ref>
The nonmetallic character of selenium is shown by its brittleness<ref name="Glinka 1965, p.,[object Object], ,[object Object],356" /> and the low electrical conductivity (~10<sup>−9</sup> to 10<sup>−12</sup>&nbsp;S•cm<sup>−1</sup>) of its highly purified form.<ref name="Kozyrev" /> This is comparable to or less than that of [[bromine]] (7.95{{e|–12}}&nbsp;S•cm<sup>−1</sup>),<ref>[[#Chao1964|Chao & Stenger 1964]]</ref> a nonmetal. Selenium has the electronic band structure of a semiconductor<ref name="Berger 1997, pp.,[object Object], ,[object Object],86–7">[[#Berger1997|Berger 1997, pp.&nbsp;86–7]]</ref> and retains its semiconducting properties in liquid form.<ref name="Berger 1997, pp.,[object Object], ,[object Object],86–7" /> It has a relatively high<ref>[[#Snyder1966|Snyder 1966, p.&nbsp;242]]</ref> electronegativity (2.55 revised Pauling scale). Its reaction chemistry is mainly that of its nonmetallic anionic forms Se<sup>2–</sup>, SeO{{su|b=3|p=2−}} and SeO{{su|b=4|p=2−}}.<ref>[[#Fritz2008|Fritz & Gjerde 2008, p.&nbsp;235]]</ref>


Selenium is commonly described as a metalloid in the [[environmental chemistry]] literature.<ref>[[#Meyer2005|Meyer et al. 2005, p.&nbsp;284]]; [[#Manahan|Manahan 2001, p.&nbsp;911]]; [[#Szpunar|Szpunar et al. 2004, p.&nbsp;17]]</ref> Processes and reactions affecting its fate in the aquatic environment are similar to those found for arsenic and antimony.<ref>[[#USEPA1988|US Environmental Protection Agency 1988, p.&nbsp;1]]; [[#Uden|Uden 2005, pp.&nbsp;347‒8]]</ref> Moreover, while trace amounts of selenium are essential to human health, its water soluble salts (in higher concentrations) have a [[toxicology|toxicological profile]] similar to that of arsenic.<ref>[[#DeZuane|De Zuane 1997, p.&nbsp;93]]; [[#Dev|Dev 2008, pp.&nbsp;2‒3]]</ref>
Selenium is commonly described as a metalloid in the [[environmental chemistry]] literature.<ref>[[#Meyer2005|Meyer et al. 2005, p.&nbsp;284]]; [[#Manahan|Manahan 2001, p.&nbsp;911]]; [[#Szpunar|Szpunar et al. 2004, p.&nbsp;17]]</ref> Processes and reactions affecting its fate in the aquatic environment are similar to those found for arsenic and antimony.<ref>[[#USEPA1988|US Environmental Protection Agency 1988, p.&nbsp;1]]; [[#Uden|Uden 2005, pp.&nbsp;347‒8]]</ref> Trace amounts of selenium are essential to human health, and its water-soluble salts (in higher concentrations) have a [[toxicology|toxicological profile]] similar to that of arsenic.<ref>[[#DeZuane|De Zuane 1997, p.&nbsp;93]]; [[#Dev|Dev 2008, pp.&nbsp;2‒3]]</ref>


===Polonium===
===Polonium===
{{main|Polonium}}
{{main|Polonium}}
[[File:Polonium.jpg|thumb|left|Polonium, in the form of a thin film on a stainless-steel disc|alt=A thin film of a bluish-grey metal on a stainless steel disc.]]
[[File:Polonium.jpg|thumb|left|Polonium, in the form of a thin film on a stainless-steel disc|alt=A thin film of a bluish-grey metal on a stainless steel disc.]]
Polonium is 'distinctly metallic' in some ways.<ref name="Cotton FA 1999, p.,[object Object], ,[object Object],502">[[#Cotton1999|Cotton et al. 1999, p.&nbsp;502]]</ref> Both of its allotropic forms are metallic conductors.<ref name="Cotton FA 1999, p.,[object Object], ,[object Object],502" /> It is soluble in acids, thereby forming the rose-coloured Po<sup>2+</sup> cation and displacing hydrogen: Po + 2 H<sup>+</sup> → Po<sup>2+</sup> + H<sub>2</sub>.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;594]]</ref> Many polonium [[Salt (chemistry)|salts]] are known.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;786]]; [[#Schwietzer2010|Schwietzer & Pesterfield 2010, pp.&nbsp;242–3]]</ref> The oxide [[polonium dioxide|PoO<sub>2</sub>]] is predominantly basic in nature.<ref name=Bagnall1966p41>[[#Bagnall1966|Bagnall 1966, p.&nbsp;41]]; [[#Nickless1968|Nickless 1968, p.&nbsp;79]]</ref> Polonium is a reluctant oxidizing agent, unlike its lighter congener oxygen: highly reducing conditions are required for the formation of the Po<sup>2–</sup> anion in aqueous solution.<ref>[[#Bagnall1990|Bagnall 1990, pp.&nbsp;313–14]]; [[#Lehto2011|Lehto & Hou 2011, p.&nbsp;220]]; [[#Siekierski2002|Siekierski & Burgess 2002, p.&nbsp;117]]: 'The tendency to form X<sup>2–</sup> anions decreases down the Group [16 elements]<span style="white-space: nowrap">...</span>'</ref>
Polonium is 'distinctly metallic' in some ways.<ref name="Cotton FA 1999, p.,[object Object], ,[object Object],502">[[#Cotton1999|Cotton et al. 1999, p.&nbsp;502]]</ref> Both of its allotropic forms are metallic conductors.<ref name="Cotton FA 1999, p.,[object Object], ,[object Object],502" /> It is soluble in acids, forming the rose-coloured Po<sup>2+</sup> cation and displacing hydrogen: Po + 2 H<sup>+</sup> → Po<sup>2+</sup> + H<sub>2</sub>.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;594]]</ref> Many polonium [[Salt (chemistry)|salts]] are known.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;786]]; [[#Schwietzer2010|Schwietzer & Pesterfield 2010, pp.&nbsp;242–3]]</ref> The oxide [[polonium dioxide|PoO<sub>2</sub>]] is predominantly basic in nature.<ref name=Bagnall1966p41>[[#Bagnall1966|Bagnall 1966, p.&nbsp;41]]; [[#Nickless1968|Nickless 1968, p.&nbsp;79]]</ref> Polonium is a reluctant oxidizing agent, unlike its lighter congener oxygen: highly reducing conditions are required for the formation of the Po<sup>2–</sup> anion in aqueous solution.<ref>[[#Bagnall1990|Bagnall 1990, pp.&nbsp;313–14]]; [[#Lehto2011|Lehto & Hou 2011, p.&nbsp;220]]; [[#Siekierski2002|Siekierski & Burgess 2002, p.&nbsp;117]]: 'The tendency to form X<sup>2–</sup> anions decreases down the Group [16 elements]<span style="white-space: nowrap">...</span>'</ref>


Polonium also shows nonmetallic character in its halides, and by way of the existence of polonides. The halides have properties generally characteristic of nonmetal halides (being volatile, easily hydrolyzed, and soluble in organic solvents).<ref>[[#Bagnall1957|Bagnall 1957, p.&nbsp;62]]; [[#Fernelius1982|Fernelius 1982, p.&nbsp;741]]</ref> Many metal [[polonide]]s, obtained by heating the elements together at 500–1,000&nbsp;°C, and containing the Po<sup>2–</sup> anion, are also known.<ref>[[#Bagnall1966|Bagnall 1966, p.&nbsp;41]]; [[#Barrett2003|Barrett 2003, p.&nbsp;119]]</ref>
Polonium also shows nonmetallic character in its halides, and by way of the existence of polonides. The halides have properties generally characteristic of nonmetal halides (being volatile, easily hydrolyzed, and soluble in organic solvents).<ref>[[#Bagnall1957|Bagnall 1957, p.&nbsp;62]]; [[#Fernelius1982|Fernelius 1982, p.&nbsp;741]]</ref> Many metal [[polonide]]s, obtained by heating the elements together at 500–1,000&nbsp;°C, and containing the Po<sup>2–</sup> anion, are also known.<ref>[[#Bagnall1966|Bagnall 1966, p.&nbsp;41]]; [[#Barrett2003|Barrett 2003, p.&nbsp;119]]</ref>
Line 356: Line 355:
===Astatine===
===Astatine===
{{main|Astatine}}
{{main|Astatine}}
Astatine, which is ordinarily classified as a nonmetal<ref>[[#Hawkes2010|Hawkes 2010]]; [[#Holt2007|Holt, Rinehart & Wilson c. 2007]]; [[#Hawkes1999|Hawkes 1999, p.&nbsp;14]]; [[#Roza2009|Roza 2009, p.&nbsp;12]]</ref> has some marked metallic properties,<ref>[[#Keller1985|Keller 1985]]</ref> and may instead be either a metalloid,<ref>[[#Harding2002|Harding, Johnson & Janes 2002, p.&nbsp;61]]</ref> or a metal.{{#tag:ref|A third option is to include astatine both as a nonmetal and as a metalloid.<ref>[[#Long1986|Long & Hentz 1986, p.&nbsp;58]]</ref>|group=n}} Immediately following its production in 1940, early investigators considered it a metal.<ref>[[#Vasáros1985|Vasáros & Berei 1985, p.&nbsp;109]]</ref> In 1949 it was called the most noble (difficult to [[redox|reduce]]) nonmetal as well as being a relatively noble (difficult to [[redox|oxidize]]) metal.<ref>[[#Haissinsky1949|Haissinsky & Coche 1949, p.&nbsp;400]]</ref> In 1950 astatine was described as a [[halogen]] and (therefore) a [[reactivity (chemistry)|reactive]] nonmetal.<ref>[[#Brownlee1950|Brownlee et al. 1950, p.&nbsp;173]]</ref> In 2013, on the basis of relativitic modelling, astatine was predicted to be a monatomic metal, having a face-centred cubic crystalline structure.<ref>[[#Hermann|Hermann, Hoffmann & Ashcroft 2013]]</ref>
Astatine, which is ordinarily classified as a nonmetal<ref>[[#Hawkes2010|Hawkes 2010]]; [[#Holt2007|Holt, Rinehart & Wilson c. 2007]]; [[#Hawkes1999|Hawkes 1999, p.&nbsp;14]]; [[#Roza2009|Roza 2009, p.&nbsp;12]]</ref> has some marked metallic properties,<ref>[[#Keller1985|Keller 1985]]</ref> and may be either a metalloid,<ref>[[#Harding2002|Harding, Johnson & Janes 2002, p.&nbsp;61]]</ref> or a metal.{{#tag:ref|A third option is to include astatine both as a nonmetal and as a metalloid.<ref>[[#Long1986|Long & Hentz 1986, p.&nbsp;58]]</ref>|group=n}} Immediately following its production in 1940, early investigators considered it a metal.<ref>[[#Vasáros1985|Vasáros & Berei 1985, p.&nbsp;109]]</ref> In 1949 it was called the most noble (difficult to [[redox|reduce]]) nonmetal as well as being a relatively noble (difficult to oxidize) metal.<ref>[[#Haissinsky1949|Haissinsky & Coche 1949, p.&nbsp;400]]</ref> In 1950 astatine was described as a [[halogen]] and (therefore) a [[reactivity (chemistry)|reactive]] nonmetal.<ref>[[#Brownlee1950|Brownlee et al. 1950, p.&nbsp;173]]</ref> In 2013, on the basis of relativistic modelling, astatine was predicted to be a monatomic metal, having a face-centred cubic crystalline structure.<ref>[[#Hermann|Hermann, Hoffmann & Ashcroft 2013]]</ref>


Several authors have commented on the metallic nature of some of the properties of astatine. Since iodine is 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.<ref>[[#Siekierski2002|Siekierski & Burgess 2002, pp.&nbsp;65, 122]]</ref>{{#tag:ref|A visible piece of astatine would be immediately and completely vaporized because of the heat generated by its intense radioactivity.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;48]]</ref>|group=n}} Astatine may at least be metallic in the liquid state on the basis that elements with an [[enthalpy of vaporization]] (EoV) greater than ~42 kJ/mol are metallic when liquid.<ref name="Rao & Ganguly 1986">[[#Rao1986|Rao & Ganguly 1986]]</ref> Such elements include boron,{{#tag:ref|The literature is contradictory as to whether boron exhibits metallic conductivity in liquid form. Krishnan et al.<ref>[[#Krishnan1998|Krishnan et al. 1998]]</ref> found that liquid boron behaved like a metal. Glorieux et al.<ref>[[#Glorieux2001|Glorieux, Saboungi & Enderby 2001]]</ref> characterised liquid boron as a semiconductor, on the basis of its low electrical conductivity. Millot et al.<ref>[[#Millot2002|Millot et al. 2002]]</ref> reported that the emissivity of liquid boron was not consistent with that of a liquid metal.|group=n}} silicon, germanium, antimony, selenium and tellurium. Estimated values for the EoV of diatomic astatine are 50 kJ/mol or higher;<ref>[[#Vasáros1985|Vasáros & Berei 1985, p.&nbsp;117]]</ref> diatomic iodine, with an EoV of 41.71,<ref>[[#Kaye1973|Kaye & Laby 1973, p.&nbsp;228]]</ref> falls just short of the threshold figure. '[L]ike typical metals, it [astatine] is precipitated by hydrogen sulfide even from strongly acid solutions and is displaced in a free form from sulfate solutions; it is deposited on the cathode on electrolysis'.<ref>[[#Samsonov1968|Samsonov 1968, p.&nbsp;590]]</ref>{{#tag:ref|Korenman<ref>[[#Korenman1959|Korenman 1959, p.&nbsp;1368]]</ref> similarly noted that 'the ability precipitate with hydrogen sulfide distinguishes astatine from other halogens and brings it closer to bismuth and other heavy metals.'|group=n}} Further indications of a tendency for astatine to behave like a (heavy) metal are: '<span style="white-space: nowrap">...</span>the formation of [[pseudohalide]] compounds<span style="white-space: nowrap">...</span>complexes of astatine cations<span style="white-space: nowrap">...</span>complex anions of trivalent astatine<span style="white-space: nowrap">...</span>as well as complexes with a variety of organic solvents'.<ref>[[#Rossler1985|Rossler 1985, pp.&nbsp;143–4]]</ref> It has also been argued that it demonstrates cationic behaviour, by way of stable At<sup>+</sup> and AtO<sup>+</sup> forms, in strongly acidic aqueous solutions.<ref>[[#Champion2010|Champion et al. 2010]]</ref>
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.<ref>[[#Siekierski2002|Siekierski & Burgess 2002, pp.&nbsp;65, 122]]</ref>{{#tag:ref|A visible piece of astatine would be immediately and completely vaporized because of the heat generated by its intense radioactivity.<ref>[[#Emsley2001|Emsley 2001, p.&nbsp;48]]</ref>|group=n}} Astatine may be metallic in the liquid state on the basis that elements with an [[enthalpy of vaporization]] (EoV) greater than ~42 kJ/mol are metallic when liquid.<ref name="Rao & Ganguly 1986">[[#Rao1986|Rao & Ganguly 1986]]</ref> Such elements include boron,{{#tag:ref|The literature is contradictory as to whether boron exhibits metallic conductivity in liquid form. Krishnan et al.<ref>[[#Krishnan1998|Krishnan et al. 1998]]</ref> found that liquid boron behaved like a metal. Glorieux et al.<ref>[[#Glorieux2001|Glorieux, Saboungi & Enderby 2001]]</ref> characterised liquid boron as a semiconductor, on the basis of its low electrical conductivity. Millot et al.<ref>[[#Millot2002|Millot et al. 2002]]</ref> reported that the emissivity of liquid boron was not consistent with that of a liquid metal.|group=n}} silicon, germanium, antimony, selenium and tellurium. Estimated values for the EoV of diatomic astatine are 50 kJ/mol or higher;<ref>[[#Vasáros1985|Vasáros & Berei 1985, p.&nbsp;117]]</ref> diatomic iodine, with an EoV of 41.71,<ref>[[#Kaye1973|Kaye & Laby 1973, p.&nbsp;228]]</ref> falls just short of the threshold figure. '[L]ike typical metals, it [astatine] is precipitated by hydrogen sulfide even from strongly acid solutions and is displaced in a free form from sulfate solutions; it is deposited on the cathode on electrolysis'.<ref>[[#Samsonov1968|Samsonov 1968, p.&nbsp;590]]</ref>{{#tag:ref|Korenman<ref>[[#Korenman1959|Korenman 1959, p.&nbsp;1368]]</ref> similarly noted that 'the ability precipitate with hydrogen sulfide distinguishes astatine from other halogens and brings it closer to bismuth and other heavy metals.'|group=n}} Further indications of a tendency for astatine to behave like a (heavy) metal are: '<span style="white-space: nowrap">...</span>the formation of [[pseudohalide]] compounds<span style="white-space: nowrap">...</span>complexes of astatine cations<span style="white-space: nowrap">...</span>complex anions of trivalent astatine<span style="white-space: nowrap">...</span>as well as complexes with a variety of organic solvents'.<ref>[[#Rossler1985|Rossler 1985, pp.&nbsp;143–4]]</ref> It has also been argued that it demonstrates cationic behaviour, by way of stable At<sup>+</sup> and AtO<sup>+</sup> forms, in strongly acidic aqueous solutions.<ref>[[#Champion2010|Champion et al. 2010]]</ref>


On the other hand, some of the reported properties of astatine are nonmetallic in nature. It has the narrow liquid range ordinarily associated with nonmetals (mp 575&nbsp;K, bp 610).<ref>[[#Borst1982|Borst 1982, pp.&nbsp;465, 473]]</ref> Batsanov gives a calculated band gap energy for astatine of 0.7&nbsp;eV;<ref>[[#Batsanov1971|Batsanov 1971, p.&nbsp;811]]</ref> this is consistent with nonmetals (in physics) having separated valence and conduction bands and thereby being either semiconductors or insulators.<ref>[[#Swalin1962|Swalin 1962, p.&nbsp;216]]; [[#Feng2005|Feng & Lin 2005, p.&nbsp;157]]</ref> The chemistry of astatine in aqueous solution is predominately characterised by the formation of various anionic species.<ref>[[#Schwietzer2010|Schwietzer & Pesterfield 2010, pp.&nbsp;258–60]]</ref> Most of its known compounds resemble those of iodine,<ref>[[#Hawkes1999|Hawkes 1999, p.&nbsp;14]]</ref> which is halogen and a nonmetal.<ref>[[#Olmsted1997|Olmsted & Williams 1997, p.&nbsp;328]]; [[#Daintith2004|Daintith 2004, p.&nbsp;277]]</ref> Such compounds include astatides (XAt), astatates (XAtO<sub>3</sub>), and monovalent [[interhalogen compound]]s.<ref>[[#Eberle1985|Eberle1985, pp.&nbsp;213–16, 222–7]]</ref>
Some of astatine's reported properties are nonmetallic. It has the narrow liquid range ordinarily associated with nonmetals (mp 575&nbsp;K, bp 610).<ref>[[#Borst1982|Borst 1982, pp.&nbsp;465, 473]]</ref> Batsanov gives a calculated band gap energy for astatine of 0.7&nbsp;eV;<ref>[[#Batsanov1971|Batsanov 1971, p.&nbsp;811]]</ref> this is consistent with nonmetals (in physics) having separated valence and conduction bands and thereby being either semiconductors or insulators.<ref>[[#Swalin1962|Swalin 1962, p.&nbsp;216]]; [[#Feng2005|Feng & Lin 2005, p.&nbsp;157]]</ref> The chemistry of astatine in aqueous solution is predominately characterised by the formation of various anionic species.<ref>[[#Schwietzer2010|Schwietzer & Pesterfield 2010, pp.&nbsp;258–60]]</ref> Most of its known compounds resemble those of iodine,<ref>[[#Hawkes1999|Hawkes 1999, p.&nbsp;14]]</ref> which is a halogen and a nonmetal.<ref>[[#Olmsted1997|Olmsted & Williams 1997, p.&nbsp;328]]; [[#Daintith2004|Daintith 2004, p.&nbsp;277]]</ref> Such compounds include astatides (XAt), astatates (XAtO<sub>3</sub>), and monovalent [[interhalogen compound]]s.<ref>[[#Eberle1985|Eberle1985, pp.&nbsp;213–16, 222–7]]</ref>


Restrepo et al.<ref>[[#Restrepo2004|Restrepo et al. 2004, p.&nbsp;69]]; [[#Restrepo2006|Restrepo et al. 2006, p.&nbsp;411]]</ref> reported that astatine appeared to share more in common with polonium than it did with the established halogens. They did so on the basis of detailed comparative studies of the known and interpolated properties of 72 elements.
Restrepo et al.<ref>[[#Restrepo2004|Restrepo et al. 2004, p.&nbsp;69]]; [[#Restrepo2006|Restrepo et al. 2006, p.&nbsp;411]]</ref> reported that astatine appeared to be more like polonium than the established halogens. They did so on the basis of detailed comparative studies of the known and interpolated properties of 72 elements.
<span id="SCC"></span>
<span id="SCC"></span>


Line 368: Line 367:


===Elements near the metalloids===
===Elements near the metalloids===
[[File:Iodinecrystals.JPG|thumb|right|Iodine crystals, showing a metallic lustre. Iodine is a semiconductor in the direction of its planes, with a band gap of ~1.3&nbsp;eV, and an electrical conductivity of 1.7 × 10<sup>−8</sup>&nbsp;S•cm<sup>−1</sup> at room temperature.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p. 804]]</ref> The latter value is higher than that of selenium but lower than that of boron, the least electrically conducting of the recognised metalloids.{{#tag:ref|The separation between molecules in the layers of iodine (350&nbsp;pm) is much less than the separation between iodine layers (427&nbsp;pm; cf. twice the van der Waals radius of 430&nbsp;pm).<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;803]]</ref> This is thought due to significant electronic interactions between the molecules in each layer of iodine, which in turn give rise to its semiconducting properties and shiny appearance.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;416]]</ref>|group=n}}|alt=Shiny violet-black coloured crystalline shards.]]
[[File:Iodinecrystals.JPG|thumb|right|Iodine crystals, showing a metallic lustre. Iodine is a semiconductor in the direction of its planes, with a band gap of ~1.3&nbsp;eV. It has an electrical conductivity of 1.7 × 10<sup>−8</sup>&nbsp;S•cm<sup>−1</sup> at room temperature,<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p. 804]]</ref> higher than selenium but lower than boron, the least electrically conducting of the recognised metalloids.{{#tag:ref|The separation between molecules in the layers of iodine (350&nbsp;pm) is much less than the separation between iodine layers (427&nbsp;pm; cf. twice the van der Waals radius of 430&nbsp;pm).<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, p.&nbsp;803]]</ref> This is thought due to electronic interactions between the molecules in each layer of iodine, which in turn give rise to its semiconducting properties and shiny appearance.<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;416]]</ref>|group=n}}|alt=Shiny violet-black coloured crystalline shards.]]
Some of the elements that occupy periodic table positions adjacent to those of the commonly recognised metalloids, although usually classified as either metals or nonmetals, are occasionally referred to as ''near-metalloids''<ref>[[#Craig2003|Craig 2003, p.&nbsp;391]]; [[#Schroers2013|Schroers 2013, p.&nbsp;32]]; [[#Vernon|Vernon 2013, pp.&nbsp;1704–1705]]</ref> or noted for their metalloidal character. To the left of the metal-nonmetal dividing line, such elements include gallium,<ref>[[#Cotton1999|Cotton et al. 1999, p.&nbsp;42]]</ref> tin<ref>[[#Marezio|Marezio & Licci 2000, p.&nbsp;11]]</ref> and bismuth.<ref name=Vernon/> They show 'odd' packing structures,<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;5]]</ref> marked covalent chemistry (molecular or polymeric),<ref>[[#Parish1977|Parish 1977, pp.&nbsp;178, 192–3]]</ref> and amphoterism.<ref>[[#Eggins1972|Eggins 1972, p.&nbsp;66]]; [[#Rayner2006|Rayner-Canham & Overton 2006, pp.&nbsp;29–30]]</ref> To the right of the dividing line are carbon,<ref>[[#Atkins2006|Atkins et al. 2006, pp.&nbsp;320–1]]; [[#Bailar1989|Bailar et al. 1989, p.&nbsp;742–3]]</ref> phosphorus,<ref>[[#Rochow1966|Rochow 1966, p.&nbsp;7]]; [[#Taniguchi1984|Taniguchi et al. 1984, p.&nbsp;867]]: '<span style="white-space: nowrap">...</span>black phosphorus<span style="white-space: nowrap">...</span>[is] characterized by the wide valence bands with rather delocalized nature.'; [[#Morita1986|Morita 1986, p.&nbsp;230]]; [[#Carmalt1998|Carmalt & Norman 1998, pp.&nbsp;1–38]]: 'Phosphorus<span style="white-space: nowrap">...</span>should therefore be expected to have some metalloid properties.'; [[#Du2010|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&nbsp;eV; observed 0.3&nbsp;eV) as opposed to the larger band gap of a single layer (calculated ~0.75&nbsp;eV).</ref> selenium<ref>[[#Stuke1974|Stuke 1974, p.&nbsp;178]]; [[#Cotton1999|Cotton et al. 1999, p.&nbsp;501]]; [[#Craig2003|Craig 2003, p.&nbsp;391]]</ref> and iodine.<ref>[[#Steudel1977|Steudel 1977, p.&nbsp;240]]: '<span style="white-space: nowrap">...</span>considerable orbital overlap must exist, to form intermolecular, many-center<span style="white-space: nowrap">...</span>[sigma] bonds, spread through the layer and populated with delocalized electrons, reflected in the properties of iodine (lustre, color, moderate electrical conductivity).'; [[#Segal1989|Segal 1989, p.&nbsp;481]]: 'Iodine exhibits some metallic properties<span style="white-space: nowrap">...</span>'.</ref> They exhibit metallic lustre, semiconducting properties{{#tag:ref|For example: intermediate electrical conductivity;<ref name="Lutz 2011, p.&nbsp;16">[[#Lutz2011|Lutz 2011, p.&nbsp;16]]</ref> a relatively narrow band gap;<ref>[[#Yacobi1990|Yacobi & Holt 1990, p.&nbsp;10]]; [[#Wiberg2001|Wiberg 2001, p.&nbsp;160]]</ref> light sensitivity.<ref name="Lutz 2011, p.&nbsp;16"/>|group=n}} 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;{{#tag:ref|White phosphorus is the most common, industrially important,<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;820]]</ref> and easily reproducible allotrope. For those reasons it is the standard state of the element.<ref>[[#Oxtoby2008|Oxtoby, Gillis & Campion 2008, p.&nbsp;508]]</ref> Paradoxically, it is also thermodynamically the least stable, as well as the most volatile and reactive form.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, pp.&nbsp;479, 482]]</ref>|group=n}} and selenium as grey selenium.
Some of the elements adjacent in the periodic table to the commonly recognised metalloids, although usually classified as either metals or nonmetals, are occasionally referred to as ''near-metalloids''<ref>[[#Craig2003|Craig 2003, p.&nbsp;391]]; [[#Schroers2013|Schroers 2013, p.&nbsp;32]]; [[#Vernon|Vernon 2013, pp.&nbsp;1704–1705]]</ref> or noted for their metalloidal character. To the left of the metal-nonmetal dividing line, such elements include gallium,<ref>[[#Cotton1999|Cotton et al. 1999, p.&nbsp;42]]</ref> tin<ref>[[#Marezio|Marezio & Licci 2000, p.&nbsp;11]]</ref> and bismuth.<ref name=Vernon/> They show 'odd' packing structures,<ref>[[#Russell2005|Russell & Lee 2005, p.&nbsp;5]]</ref> marked covalent chemistry (molecular or polymeric),<ref>[[#Parish1977|Parish 1977, pp.&nbsp;178, 192–3]]</ref> and amphoterism.<ref>[[#Eggins1972|Eggins 1972, p.&nbsp;66]]; [[#Rayner2006|Rayner-Canham & Overton 2006, pp.&nbsp;29–30]]</ref> To the right of the dividing line are carbon,<ref>[[#Atkins2006|Atkins et al. 2006, pp.&nbsp;320–1]]; [[#Bailar1989|Bailar et al. 1989, p.&nbsp;742–3]]</ref> phosphorus,<ref>[[#Rochow1966|Rochow 1966, p.&nbsp;7]]; [[#Taniguchi1984|Taniguchi et al. 1984, p.&nbsp;867]]: '<span style="white-space: nowrap">...</span>black phosphorus<span style="white-space: nowrap">...</span>[is] characterized by the wide valence bands with rather delocalized nature.'; [[#Morita1986|Morita 1986, p.&nbsp;230]]; [[#Carmalt1998|Carmalt & Norman 1998, pp.&nbsp;1–38]]: 'Phosphorus<span style="white-space: nowrap">...</span>should therefore be expected to have some metalloid properties.'; [[#Du2010|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&nbsp;eV; observed 0.3&nbsp;eV) as opposed to the larger band gap of a single layer (calculated ~0.75&nbsp;eV).</ref> selenium<ref>[[#Stuke1974|Stuke 1974, p.&nbsp;178]]; [[#Cotton1999|Cotton et al. 1999, p.&nbsp;501]]; [[#Craig2003|Craig 2003, p.&nbsp;391]]</ref> and iodine.<ref>[[#Steudel1977|Steudel 1977, p.&nbsp;240]]: '<span style="white-space: nowrap">...</span>considerable orbital overlap must exist, to form intermolecular, many-center<span style="white-space: nowrap">...</span>[sigma] bonds, spread through the layer and populated with delocalized electrons, reflected in the properties of iodine (lustre, color, moderate electrical conductivity).'; [[#Segal1989|Segal 1989, p.&nbsp;481]]: 'Iodine exhibits some metallic properties<span style="white-space: nowrap">...</span>'.</ref> They exhibit metallic lustre, semiconducting properties{{#tag:ref|For example: intermediate electrical conductivity;<ref name="Lutz 2011, p.&nbsp;16">[[#Lutz2011|Lutz 2011, p.&nbsp;16]]</ref> a relatively narrow band gap;<ref>[[#Yacobi1990|Yacobi & Holt 1990, p.&nbsp;10]]; [[#Wiberg2001|Wiberg 2001, p.&nbsp;160]]</ref> light sensitivity.<ref name="Lutz 2011, p.&nbsp;16"/>|group=n}} 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;{{#tag:ref|White phosphorus is the most common, industrially important,<ref>[[#Eagleson1994|Eagleson 1994, p.&nbsp;820]]</ref> and easily reproducible allotrope. For those reasons it is the standard state of the element.<ref>[[#Oxtoby2008|Oxtoby, Gillis & Campion 2008, p.&nbsp;508]]</ref> Paradoxically, it is also thermodynamically the least stable, as well as the most volatile and reactive form.<ref>[[#Greenwood2002|Greenwood & Earnshaw 2002, pp.&nbsp;479, 482]]</ref>|group=n}} and selenium as grey selenium.


===Allotropes===
===Allotropes===
[[File:Sn-Alpha-Beta.jpg|thumb|left|White tin (left) and grey tin (right). Both forms have a metallic appearance.|alt=Many small, shiny, silver-coloured spheres on the left; many of the same sized spheres on the right are duller and darker than the ones of the left and have a subdued metallic shininess.]]
[[File:Sn-Alpha-Beta.jpg|thumb|left|White tin (left) and grey tin (right). Both forms have a metallic appearance.|alt=Many small, shiny, silver-coloured spheres on the left; many of the same sized spheres on the right are duller and darker than the ones of the left and have a subdued metallic shininess.]]
When an element exists in more than one crystalline form, the different forms 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.<ref>[[#Brescia1980|Brescia et al. 1980, pp.&nbsp;166–71]]</ref> The existence of such allotropes can complicate the classification of the elements involved.<ref>[[#Fine|Fine & Beall 1990, p.&nbsp;578]]</ref>
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.<ref>[[#Brescia1980|Brescia et al. 1980, pp.&nbsp;166–71]]</ref> The existence of such allotropes can complicate the classification of the elements involved.<ref>[[#Fine|Fine & Beall 1990, p.&nbsp;578]]</ref>


Tin, for example, has two allotropes: tetragonal 'white' β-tin and cubic 'grey' α-tin, as shown in the picture to the right. White tin is a silvery-white, very shiny, ductile and malleable metal. It is the stable form of tin at or above room temperature and has an electrical conductivity of 9.17×10<sup>4</sup>&nbsp;S·cm<sup>−1</sup> (~1/6th that of copper).<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;901]]</ref> Grey tin, in contrast, usually has the appearance of a grey micro-crystalline powder although it can also be prepared in ordinary crystalline or polycrystalline forms having a semi-lustrous appearance and a brittle comportment. It is the stable form of tin below 13.2&nbsp;°C (56&nbsp;°F) and has an electrical conductivity of between (2–5)×10<sup>2</sup>&nbsp;S·cm<sup>−1</sup> (~1/250th that of white tin).<ref>[[#Berger1997|Berger 1997, p.&nbsp;80]]</ref> Grey tin has the same crystalline structure as that of the diamond allotrope of carbon. It behaves as a semiconductor (with a band gap of 0.08&nbsp;eV), but has the electronic band structure of a semimetal.<ref>[[#Lovett1977|Lovett 1977, p.&nbsp;101]]</ref> It has been referred to as either a very poor metal,<ref>[[#Cohen|Cohen & Chelikowsky 1988, p.&nbsp;99]]</ref> a metalloid,<ref>[[#Taguena|Taguena-Martinez, Barrio & Chambouleyron 1991, p.&nbsp;141]]</ref> a nonmetal<ref>[[#Ebbing|Ebbing & Gammon 2010, p.&nbsp;891]]</ref> or a near metalloid.<ref name=Vernon>[[#Vernon|Vernon 2013, p.&nbsp;1705]]</ref>
Tin, for example, has two allotropes: tetragonal 'white' β-tin and cubic 'grey' α-tin, as shown in the picture. White tin is a silvery-white, very shiny, ductile and malleable metal. It is the stable form of tin at or above room temperature and has an electrical conductivity of 9.17×10<sup>4</sup>&nbsp;S·cm<sup>−1</sup> (~1/6th that of copper).<ref>[[#Wiberg2001|Wiberg 2001, p.&nbsp;901]]</ref> Grey tin usually has the appearance of a grey micro-crystalline powder although it can also be prepared in crystalline or polycrystalline forms having a semi-lustrous appearance and a brittle comportment. It is the stable form of tin below 13.2&nbsp;°C (56&nbsp;°F) and has an electrical conductivity of between (2–5)×10<sup>2</sup>&nbsp;S·cm<sup>−1</sup> (~1/250th that of white tin).<ref>[[#Berger1997|Berger 1997, p.&nbsp;80]]</ref> Grey tin has the same crystalline structure as that of the diamond allotrope of carbon. It behaves as a semiconductor (with a band gap of 0.08&nbsp;eV), but has the electronic band structure of a semimetal.<ref>[[#Lovett1977|Lovett 1977, p.&nbsp;101]]</ref> It has been referred to as either a very poor metal,<ref>[[#Cohen|Cohen & Chelikowsky 1988, p.&nbsp;99]]</ref> a metalloid,<ref>[[#Taguena|Taguena-Martinez, Barrio & Chambouleyron 1991, p.&nbsp;141]]</ref> a nonmetal<ref>[[#Ebbing|Ebbing & Gammon 2010, p.&nbsp;891]]</ref> or a near metalloid.<ref name=Vernon>[[#Vernon|Vernon 2013, p.&nbsp;1705]]</ref>


The diamond allotrope of carbon, as another example, is clearly nonmetallic, being translucent and having a relatively poor electrical conductivity of 10<sup>−14</sup> to 10<sup>−16</sup>&nbsp;S·cm<sup>−1</sup>.<ref>[[#Asmussen|Asmussen & Reinhard 2002, p.&nbsp;7]]</ref> The semi-lustrous and more stable graphite allotrope, in contrast, has an electrical conductivity of 3×10<sup>4</sup>&nbsp;S·cm<sup>−1</sup>,<ref>[[#Deprez1988|Deprez & McLachan 1988]]</ref> a figure more characteristic of a metal. Phosphorus, sulfur, arsenic, selenium, antimony and bismuth also have less stable allotropes that display borderline or either more or less metallic or nonmetallic behaviour.<ref>[[#Addison1964|Addison 1964 (P, Se, Sn)]]; [[#Marko1998|Marković, Christiansen & Goldman 1998 (Bi)]]; [[#Nagao2004|Nagao et al. 2004]]</ref>
The diamond allotrope of carbon is clearly nonmetallic, being translucent and having a relatively poor electrical conductivity of 10<sup>−14</sup> to 10<sup>−16</sup>&nbsp;S·cm<sup>−1</sup>.<ref>[[#Asmussen|Asmussen & Reinhard 2002, p.&nbsp;7]]</ref> Graphite has an electrical conductivity of 3×10<sup>4</sup>&nbsp;S·cm<sup>−1</sup>,<ref>[[#Deprez1988|Deprez & McLachan 1988]]</ref> a figure more characteristic of a metal. Phosphorus, sulfur, arsenic, selenium, antimony and bismuth also have less stable allotropes that display borderline or either more or less metallic or nonmetallic behaviour.<ref>[[#Addison1964|Addison 1964 (P, Se, Sn)]]; [[#Marko1998|Marković, Christiansen & Goldman 1998 (Bi)]]; [[#Nagao2004|Nagao et al. 2004]]</ref>


==Notes==
==Notes==

Revision as of 16:35, 19 February 2014

  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 chemical element that has properties that are in between those of metals and nonmetals. There is no standard definition of a metalloid, nor is there agreement as to which elements are appropriately classified as such. Despite this lack of specificity the term remains in use in chemistry literature.

The six elements commonly recognised as metalloids are boron, silicon, germanium, arsenic, antimony and tellurium. Other elements less commonly recognised as metalloids include carbon, aluminium, selenium, polonium and astatine. On a standard periodic table all of these elements can be found in or near a diagonal region of the p-block, having its main axis anchored by boron at one end and astatine at the other. Some periodic tables include a dividing line between metals and nonmetals and it is generally the elements adjacent to this line or, less often, one or more of the elements adjacent to those elements, which are identified as metalloids.

Physically, metalloids usually have a metallic appearance but they are brittle and only fair conductors of electricity; chemically, they mostly behave as (weak) nonmetals. They can form alloys with metals. Ordinarily, most of the other physical and chemical properties of metalloids are intermediate in nature.

Metalloids and their compounds are too brittle to have any structural uses. They are used in alloys, biological agents, flame retardants, glasses, optical storage and semiconductors. 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 onward.[1]

The term metalloid originally referred to nonmetals. Its more recent meaning, as a category of elements with intermediate or hybrid properties, did not become widespread until 1940–1960. Metalloids are sometimes called semimetals, a practice that has been discouraged.[2] This is because the term semimetal 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 a chemical element.

Definition

Metalloids are usually regarded as a third category of chemical elements occupying a fuzzy 'buffer zone' between metals and nonmetals.[3][n 2] There is no universally agreed, rigorous definition of a metalloid.[8] The feasibility of establishing a specific definition has also been questioned, as anomalies can be found in such attempted constructs.[9] Classifying an element as a metalloid has been described as 'arbitrary'.[10]

Generic

A metalloid is a chemical element that has properties in between or a mixture of those of metals and nonmetals and is difficult to classify as a metal or a nonmetal. 'In chemistry a metalloid is an element with properties intermediate between those of metals and nonmetals.'[11] These elements are characterised by their mixed properties, and categorisation difficulties: '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.'[12] These difficulties can be accommodated by recognising another category of elements: 'Chemists sometimes use the name metalloid...for these elements difficult to classify one way or the other.'[13] A few authors are more explicit: '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...'.[14] Metalloids have also been referred to as 'elements that...are somewhat of a cross between metals and nonmetals'[15] or 'weird in-between elements.'[16]

The criterion that metalloids be difficult to unambiguously classify one way or the other is a key tenet. In contrast, elements such as sodium and potassium 'have metallic properties to a high degree' and fluorine, chlorine and oxygen 'are almost exclusively nonmetallic.'[17] Although most elements have a mixture of metallic and nonmetallic properties,[17] they can be classified as either metals or nonmetals according to which set of properties is more pronounced.[18][n 3] Only the elements at or near the margins, regarded as lacking a sufficiently clear preponderance of metallic or nonmetallic properties, are classified as metalloids.[22]

Boron, silicon, germanium, arsenic, antimony and tellurium are commonly recognised as metalloids.[23][n 4] Depending on the author, one or more from selenium, polonium or astatine are sometimes added to this list.[25] Boron is sometimes excluded, by itself or together with silicon.[26] Tellurium is sometimes not regarded as a metalloid.[27] The inclusion of antimony, polonium and astatine as metalloids has also been questioned.[28]

Some other elements are occasionally classified as metalloids. These elements include[29] hydrogen,[30] beryllium,[31] nitrogen,[32] phosphorus,[33] sulfur,[34] zinc,[35] gallium,[36] tin, iodine,[37] lead,[38] bismuth[27] and radon.[39] The term metalloid has also been used to refer to elements that exhibit metallic lustre and electrical conductivity, and that are amphoteric; examples include arsenic, antimony, vanadium, chromium, molybdenum, tungsten, tin, lead and aluminium.[40] The poor metals,[41] and nonmetals (such as carbon or nitrogen) that can form alloys with,[42] or modify the properties of,[43] metals have also occasionally been considered as metalloids.

Specific

    Element
IE 
EN
 Band structure   
Boron  191    2.04   semiconductor 
  Silicon  187    1.90   semiconductor   
Germanium   182    2.01   semiconductor 
  Arsenic  225    2.18   semimetal  
Antimony  198    2.05   semimetal
  Tellurium  207    2.10   semiconductor   
average   198    2.05 
The elements commonly recognised as metalloids, and their ionization energies (kcal/mol);[44] electronegativities (revised Pauling scale); and electronic band structures[45] (most thermodynamically stable forms under ambient conditions).

Metalloids tend to be characterised in terms of generalities or a few broadly indicative physical or chemical properties.[46] A single quantitative criterion, such as electronegativity, is also occasionally mentioned.[n 5] Jones[55] (writing on the role of classification in science) observed that, 'Classes are usually defined by more than two attributes.'

Masterton and Slowinski[56] offer a more specific treatment. They write that metalloids have ionization energies around 200 kcal/mol, and electronegativity values close to 2.0. They also say that metalloids are typically semiconductors, though antimony and arsenic (semimetals in the physics sense) have electrical conductivities that approach those of metals. Their description, using these three more or less clearly defined properties, encompasses the six elements commonly recognised as metalloids (see table, upper right). Selenium and polonium are probably excluded from this scheme; astatine may or may not be included.[n 6]

The commonly recognised metalloids can also be quantitatively described in terms of their intermediate packing efficiencies (between 34% to 41%) and Goldhammer-Herzfeld criterion metallization ratios (between ~0.85 to 1.1; average 1.0).[59][n 7] The packing efficiency of boron is 38%; silicon and germanium 34; arsenic 38.5; antimony 41; and tellurium 36.4.[61] These values are lower than the values of most metals (at least 80% of which have a packing efficiency of at least 68%)[62][n 8] but higher than those of elements usually classified as nonmetals. Packing efficiencies for nonmetals are: graphite 17%,[66] sulfur 19.2,[67] iodine 23.9,[67] selenium 24.2,[67] and black phosphorus 28.5.[64] The Goldhammer-Herzfeld criterion ratio values of the recognised metalloids are lower than those of representative and transition metals and higher than those of nonmetals.[n 9]

Periodic table territory

Location

Template:Periodic table (metalloid border) 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.[71] 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.[72]

The diagonal positioning of the metalloids represents an exception to the observation that elements with similar properties tend to occur in vertical columns.[73] Going along a period, the nuclear charge increases with atomic number as there is as an increase in electrons. The additional 'pull' on outer electrons with increasing nuclear charge generally outweighs the screening efficacy 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.[74] 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. With some irregularities, atoms become larger, ionization energy falls, and metallic character increases.[75] The combined effect of these competing horizontal and vertical trends is that the location of the metal-nonmetal transition zone shifts to the right in going down a group.[73] A related effect can be seen in other diagonal similarities between some elements and their lower right neighbours, such as lithium-magnesium, beryllium-aluminium, carbon-phosphorus, and nitrogen-sulfur.[76]

Number, composition and alternative treatments

How many and which elements are metalloids depends on the classification criteria being used. Emsley,[77] for example, recognised only four: germanium, arsenic, antimony and tellurium; James et al.,[78] listed twelve: boron, carbon, silicon, germanium, arsenic, selenium, antimony, tellurium, bismuth, polonium, ununpentium and livermorium. On average, seven elements are included in such lists.

With no standardized division of the elements into metals, metalloids and nonmetals, a specified subset of the continuum from the metallic to the nonmetallic can potentially serve its particular purpose as well as any other.[79] Individual metalloid classification arrangements tend to share common ground, with most variations occurring around the indistinct[80] margins.[n 10]

Some authors do not classify elements bordering the metal-nonmetal dividing line as metalloids, noting that a binary classification can facilitate the establishment of some simple rules for determining bond types between metals and nonmetals.[3] Metalloids are grouped instead with metals,[82] regarded as nonmetals[83] or treated as a sub-category of them.[84][n 11] Other authors have suggested that classifying some elements as metalloids 'emphasizes that properties change gradually rather than abruptly as one moves across or down the periodic table'.[86] Some periodic tables distinguish elements that are metalloids in the absence of any formal dividing line between metals and nonmetals. Metalloids are shown as occurring in a diagonal band[87] or diffuse region.[88]

Properties of metalloids

Physical and chemical

Metalloids are usually characterised as metallic-looking 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 have amphoteric or weakly acidic oxides. They can form alloys with metals. Most of the other physical and chemical properties of metalloids are intermediate in nature.

Distinctive

Brittleness,[89] semiconductivity[90] or both[91] have been used as distinguishing indicators of metalloid status. Metallic lustre along with dualistic chemical behaviour—for example, amphoterism—has also been cited as a benchmark.[92]

A metalloid is a chemistry-based concept describing the physical (including electronic) and chemical properties of elements. Semiconductor is a physics-based concept referring to the electronic properties of materials (including both elements and compounds).[93] Not all elements classified in the literature as metalloids display semiconductivity, although most do.[94]

Metalloids are all solid,[95] and have metallic lustre, but their other properties vary.[96] Given that metallic character (for example) is a combination of several properties, it has been suggested that metalloid status be judged separately for each element. This could be done based on the extent to which an element exhibits properties relevant to such status.[97]

Compared to those of metals and nonmetals

Characteristic properties of metals, metalloids and nonmetals are summarized in the table.[98] 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)[99][n 12] solid[101] mostly gases[102]
Appearance lustrous (at least when freshly fractured) lustrous[101] several colourless; others coloured, or metallic grey to black
Elasticity typically elastic, ductile, malleable (when solid) brittle[103] brittle, if solid
Electrical conductivity good to high[n 13] intermediate[105] to good[n 14] poor to good[n 15]
Band structure metallic (Bi = semimetallic) are semiconductors or, if not (As, Sb = semimetallic), exist in semiconducting forms[109] semiconductor or insulator[110]
Chemical property Metals Metalloids Nonmetals
General chemical behaviour metallic nonmetallic[111] nonmetallic
Ionization energy relatively low intermediate ionization energies,[112] usually falling between those of metals and nonmetals[113] relatively high
Electronegativity usually low have electronegativity values close to 2[114] (revised Pauling scale) or within the range of 1.9–2.2 (Allen scale)[24][n 16] high
When mixed
with metals
give alloys can form alloys[117] ionic or interstitial compounds formed
Oxides lower oxides basic; higher oxides increasingly acidic amphoteric or weakly acidic[118] acidic

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.

Typical or shared applications

Metalloids are too brittle to have any structural uses in their pure forms.[119] Theyand their compounds are used as alloying components, biological agents (toxicological, nutritional and medicinal), flame retardants, glasses (oxide and metallic), optical storage media and electronics, and semiconductors.[n 17]

Alloys

Several dozen metallic pellets, reddish-brown with a highly polished appearance, as if they had a cellophane coating.
Copper-germanium alloy pellets, likely ~84% Cu; 16% Ge.[121] When combined with silver the result is a tarnish resistant sterling silver. Also present are two silver pellets.

Writing early in the history of intermetallic compounds, the British metallurgist Cecil Desch observed that 'certain non-metallic elements are capable of forming compounds of distinctly metallic character with metals, and these elements may therefore enter into the composition of alloys'. He associated silicon, arsenic and tellurium—in particular—with the alloy-forming elements.[122] Compounds of silicon, germanium, arsenic and antimony with the poor metals, it has been suggested, 'are probably best classed as alloys.'[123]

Alloys with transition metals are well-represented. Boron can form intermetallic compounds and alloys with such metals, of the composition MnB, if n > 2.[124] Ferroboron (15% boron) is used to introduce boron into steel; nickel-boron alloys are ingredients in welding alloys and face-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.[125] Arsenic can form alloys with metals, including platinum and copper;[126] it is also added to copper and its alloys to improve corrosion resistance[127] and appears to confer the same benefit when added to magnesium.[128] Antimony is well known as an alloy former, including with the coinage metals. Its alloys are exemplified by pewter (a tin alloy with up to 20% antimony) and type metal (a lead alloy with up to 25% antimony).[129] Tellurium readily alloys with iron, as ferrotellurium (50–58% tellurium), and with copper, in the form of copper tellurium (40–50% tellurium).[130] Ferrotellurium is used as a stabilizer for carbon in steel casting.[131] Of the non-metallic elements that are less often recognised as metalloids, selenium—in the form of ferroselenium (50–58% selenium)—is used to improve the workability of stainless steels.[132]

Biological agents

A clear glass dish on which is a small mound of a white crystalline powder.
Arsenic trioxide or white arsenic, one of the most toxic and prevalent forms of arsenic. The antileukaemic properties of white arsenic were first reported in 1878.[133]

All six of the elements commonly recognised as metalloids have toxic, dietary or medicinal properties[134] to varying degrees. Arsenic and antimony are toxic; boron, silicon, and possibly arsenic, are essential trace elements. Boron, silicon, arsenic and antimony have medical applications (with germanium and tellurium being thought to have potential).

Boron is used in insecticides[135] and herbicides.[136] It is an essential trace element.[137] As boric acid, it has antiseptic, antifungal, and antiviral properties.

Silicon is not toxic. It is present in silatrane, a highly toxic rodenticide.[138] Long-term inhalation of silica dust causes silicosis, a fatal disease of the lungs. Silicon is an essential trace element.[137] Silicone gel can be applied to badly burned patients to reduce scarring.[139]

Salts of germanium are potentially harmful to humans and animals if ingested on a prolonged basis.[140] It is not an essential trace element. Although interest in the pharmacological actions of germanium compounds is ongoing there is (as yet) no licensed medicine.[141]

Arsenic is notoriously poisonous and may possibly also be an essential element in ultratrace amounts.[142] It has been used as a pharmaceutical agent since antiquity for the treatment of syphilis prior to the development of antibiotics.[143] Arsenic is also a component of melarsoprol, a medicinal drug used in the treatment of human African trypanosomiasis or sleeping sickness. In 2003, arsenic trioxide (under the trade name Trisenox) was re-introduced for the treatment of acute promyelocytic leukaemia, a cancer of the blood and bone marrow.[143]

Metallic antimony is relatively non-toxic; most antimony compounds are poisonous.[144] It is not an essential element. Compounds of antimony are used as antiprotozoan drugs, and in some veterinary preparations.

Tellurium is not considered particularly toxic although as little as two grams of sodium tellurate, if administered, can be lethal.[145] People exposed to small amounts of airborne tellurium exude a foul and persistent garlic-like odour.[146] It is not an essential element. Tellurium dioxide has been used to treat seborrhoeic dermatitis; other tellurium compounds were used as antimicrobial agents before the development of antibiotics.[147] Such compounds may have the potential to act as substitutes for antibiotics that have become ineffective due to increasing bacterial resistance.[148]

Of the elements less often recognised as metalloids, beryllium and lead are noted for their toxicity, with lead arsenate having been extensively used as an insecticide.[149] Sulfur is one of the oldest fungicides and pesticides. Phosphorus, sulfur, zinc and iodine are essential nutrients, as are possibly aluminium, tin and lead.[142] Sulfur, gallium, iodine and bismuth have medicinal applications. Sulfur is a constituent of sulfonamide drugs, still widely used for conditions such as acne and urinary tract infections.[150] Gallium nitrate is used to treat the side effects of cancer;[151] and gallium citrate, a radiopharmaceutical, is used to facilitate body imaging in areas of inflammation, such as infection, and areas of rapid cell division.[152] Iodine is used as a disinfectant in various forms. Bismuth is an ingredient in some antibacterial pharmaceuticals.[153]

Flame retardants

Compounds of boron, silicon, arsenic and antimony have been used as flame retardants. Boron, in the form of borax, has been used as a textile flame retardant since at least the 18th century.[154] Silicon compounds such as silicones, silanes, silsesquioxane, silica and silicates, some of which were developed as alternatives to more toxic halogenated products, can considerably improve the flame retardancy of plastic materials.[155] Arsenic compounds in the form of sodium arsenite or sodium arsenate are effective flame retardants for wood but were less frequently used due to their toxicity.[156] Antimony, as antimony trioxide, is a flame retardant.[157] Aluminium hydroxide, has been used as a wood-fibre, rubber, plastic and textile flame retardant since the 1890s.[158] Barring aluminium hydroxide, use of phosphorus based flame-retardants in the form of, for example, organophosphates, now exceeds that of any of the other main retardant types, which employ boron, antimony or halogenated hydrocarbon compounds.[159]

Glass formation

A bunch of pale yellow semi-transparent thin strands, with bright points of white light at their tips.
Optical fibres, usually made of pure silicon dioxide glass, with additives such as boron trioxide or germanium dioxide for increased sensitivity

The oxides B2O3, SiO2, GeO2, As2O3 and Sb2O3 readily form glasses. TeO2 forms a glass but this requires a 'heroic quench rate' or the addition of an impurity; otherwise the crystalline form results.[160] These compounds are used in chemical, domestic and industrial glassware[161] and optics.[162] Boron trioxide is used as a glass fibre additive;[163] it is also a component of borosilicate glass, which is widely used for laboratory glassware and domestic ovenware.[164] Silicon dioxide forms most ordinary glassware.[165] Germanium dioxide is used as a glass fibre additive, as well as in infrared optical systems.[166] Arsenic trioxide is used in the glass industry as a decolourizing and fining agent, as is antimony trioxide.[167] Tellurium dioxide finds application in laser and nonlinear optics.[168]

Amorphous 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.[169][n 18] Aside from thin films deposited at very low temperatures, the first known metallic glass was a metal-metalloid alloy of composition Au75Si25 reported in 1960.[171] A metallic glass having a strength and toughness not previously seen, of composition Pd82.5P6Si9.5Ge2, was reported in 2011.[172]

Phosphorus, selenium and lead, which are less often recognised as metalloids, are also used in glasses. Phosphate glass has a substrate of phosphorus pentoxide (P2O5), rather than the silica (SiO2) of convention silicate glasses and is used, for example, to make sodium lamps.[173] Selenium compounds can be used both as decolourising agents and to add a red colour to glass.[174] 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.[175] Lead-based glasses have also been extensively used in electronics 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.[176]

Optical storage

Varying compositions of GeSbTe ("GST alloys") and Ag- and In- doped Sb2Te ("AIST alloys"), being examples of phase-change materials, are widely used in rewritable optical disks and phase-change memory devices. By applying heat, they can be switched between amorphous (glassy) and crystalline states, changing their optical and electrical properties and allowing the storage of information.[177]

Semiconductors and electronics

A small square plastic piece with three parallel wire protrusions on one side; a larger rectangular plastic chip with multiple plastic and metal pin-like legs; and a small red light globe with two long wires coming out of its base.
Semiconductor-based electronic components. From left to right: a transistor, an integrated circuit and an LED. The elements commonly recognised as metalloids find widespread use in such devices, as elemental or compound semiconductor constituents (Si, Ge or GaAs, for example) or as doping agents (B, Sb, Te, for example).

All the elements commonly recognised as metalloids (or their compounds) have found application in the semiconductor or solid-state electronic industries.[178] Some properties of boron have retarded 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.[179] Silicon is the leading commercial semiconductor; it forms the basis of modern electronics (including standard solar cells)[180] and information and communication technologies.[181] This occurred despite the study of semiconductors, early in the 20th century, being regarded as the 'physics of dirt' and not deserving of close attention.[182] Silicon has largely replaced germanium in semiconducting devices, being cheaper, more resilient at higher operating temperatures, and easier to work during the microelectronic fabrication process.[121] Silicon-germanium 'alloys' have been growing in use, particularly for wireless communication devices; these exploit the higher carrier mobility of germanium.[121] The synthesis of gram-scale quantities of semiconducting germanane was also reported in 2013. This comprises one-atom thick sheets of hydrogen-terminated germanium atoms. It conducts electrons more than ten times faster than silicon and five times faster than conventional germanium, and is thought to have potential for optoelectronic and sensing applications.[183]

Arsenic and antimony are not semiconductors in their standard states; both form type III-V semiconductors (such as GaAs, AlSb or GaInAsSb) in which the average number of valence electrons per atom is the same as that of Group 14 elements. These compounds are preferred for some special applications.[184] Tellurium, which is a semiconductor in its standard state, is used mainly as a component in a large group of type II/VI semiconducting-chalcogenides; these compounds have applications in electro-optics and electronics.[185] Cadmium telluride (CdTe) finds application in solar modules due its high conversion efficiency, low manufacturing costs, and large band gap of 1.44 eV, meaning it absorbs a wide range of solar wavelengths.[180] Bismuth telluride (Bi2Te3) alloyed with selenium and antimony, is a component of thermoelectric devices used for refrigeration or portable power generation.[186] Five metalloids—boron, silicon, germanium, arsenic and antimony—can be found in cell phones (along with at least 39 other metals and nonmetals).[187] Tellurium is a component of phase change memory and is expected to find such use.[188] Of the elements less often recognised as metalloids, phosphorus, gallium (in particular) and selenium have semiconductor applications. Phosphorus is used in trace amounts as a dopant for n-type semiconductors.[189] 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.[190] Selenium is used in the production of solar cells[191] and in high-energy surge protectors.[192]

Nomenclature and history

Derivation and other names

The word metalloid comes from the Latin metallum = "metal" and the Greek oeides = "resembling in form or appearance".[193] Although the terms amphoteric element,[194] boundary element,[195] half-metal,[196] half-way element,[197] near metal,[198] meta-metal,[199] semiconductor,[200] semimetal[201] and submetal[202] are sometimes used synonymously, some of these have other meanings that may not be interchangeable. 'Amphoteric element' is sometimes used more broadly to include transition metals capable of forming oxyanions, such as chromium and manganese.[203] Some elements referred to as metalloids do not show marked amphoteric behaviour or semiconductivity in their most stable forms. 'Half-metal' is used in physics to refer to a compound (such as 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.[196] These metals are mostly diamagnetic[204] and tend to have distorted crystalline structures, electrical conductivity values at the lower end of those of metals, and amphoteric (weakly basic) oxides.[205] 'Semimetal' sometimes refers, loosely or explicitly, to metals with incomplete metallic character in crystalline structure, electrical conductivity or electronic structure. Examples include gallium,[206] ytterbium,[207] bismuth[208] and neptunium.[209]

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 as early as 1800.[210] Only since the mid-20th century has it been widely used to refer to intermediate or borderline chemical elements.[46] The International Union of Pure and Applied Chemistry (IUPAC) has previously recommended abandoning the term metalloid, and suggested using the term semimetal instead.[211] Use of this latter term has more recently been discouraged[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 or non-usage of the terms metalloid or semimetal.[212]

Elements commonly recognised as metalloids

Boron

Several dozen small angular stone like shapes, grey with scattered silver flecks and highlights.
Boron, shown here in the form of its β-rhombohedral phase (its most thermodynamically stable allotrope)[213]

Pure boron is a shiny, silver-grey crystalline solid.[214] It is less dense than aluminium, and is hard and brittle. It is barely reactive under normal conditions, except for attack by fluorine,[215] and has a melting point several hundred degrees higher than that of steel. Boron is a semiconductor,[216] with a room temperature electrical conductivity of 1.5 × 10−6 S•cm−1[217] (about 200 times less than that of tap water)[218] and a band gap of about 1.56 eV.[219][n 19]

The chemistry of boron is dominated by its small atomic size, relatively high ionization energy, and having fewer valence electrons (three) than atomic orbitals (four) available for bonding. With only three valence electrons, simple covalent bonding is electron deficient with respect to the octet rule.[221] Elements in this situation usually adopt metallic bonding, but the small size and high ionization energies of boron tends to result in delocalized covalent bonding,[222] in which three atoms share two electrons, rather than metallic bonding. The associated structural component that pervades the various allotropes of boron is the icosahedral B12 unit. This also occurs, as do deltahedral variants or fragments, in several metal borides, certain hydrides, and some halides.[223] The bonding in boron has been described as being characteristic of behaviour intermediate between metals and nonmetallic covalent network solids (such as diamond).[224] 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 how close boron is to the metal-nonmetal borderline.[225]

Most of the chemistry of boron is nonmetallic in nature.[225] The small size of the boron atom enables the preparation of many interstitial alloy-type borides.[226] Analogies between boron and transition metals have been noted in the formation of complexes,[227] and adducts (for example, BH3 + CO →BH3CO and, similarly Fe(CO)4 + CO →Fe(CO)5), as well as in the geometric and electronic structures of cluster species such as [B6H6]2– and [Ru6(CO)18]2–.[228][n 20] The aqueous chemistry of boron is characterised by the formation of many different polyborate anions.[230] Given its high charge-to-size ratio, nearly all compounds of boron are covalent, with a few complexed anionic and cationic species.[231] Boron has a strong affinity for oxygen, and has an extensive chemistry as borates.[226] The oxide B2O3 is polymeric in structure,[232] weakly acidic,[233] and a glass former.[234] Organometallic compounds of boron have been known since the 19th century (see organoboron chemistry).[235]

Silicon

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

Silicon is a shiny crystalline solid, with a blue-grey metallic lustre.[236] Like boron, it is less dense than aluminium, and is hard and brittle.[237] It is a relatively unreactive element,[236] the massive, crystalline form (especially if pure) being 'remarkably inert to all acids, including hydrofluoric'.[238] Less pure silicon, and the powdered form, are variously susceptible to attack by strong or heated acids, as well as by steam and fluorine.[239] Silicon reacts with hot aqueous alkalis with the evolution of hydrogen, like a metal[240] such as beryllium, aluminium, zinc, gallium and indium.[241] It melts at about the same temperature as steel. Silicon is a semiconductor with an electrical conductivity of 10−4 S•cm−1[242] and a band gap of about 1.11 eV.[243] When it melts, silicon becomes a reasonable metal[244] with an electrical conductivity of 1.0–1.3 × 104 S•cm−1, similar to that of liquid mercury.[245]

The chemistry of silicon is generally nonmetallic (covalent) in nature.[246] It can form alloys with metals such as iron and copper.[247] Silicon shows fewer tendencies to anionic behaviour than ordinary nonmetals.[248] Its solution chemistry is characterised by the formation of oxyanions.[249] The high strength of the silicon-oxygen bond dominates the chemical behaviour of silicon.[250] Polymeric silicates, built up by tetrahedral SiO4 units sharing their oxygen atoms, are the most abundant and important compounds of silicon.[251] The polymeric borates, comprising linked trigonal and tetrahedral BO3 or BO4 units, are built on similar structural principles.[252] The oxide SiO2 is polymeric in structure,[232] weakly acidic,[253][n 21] and a glass former.[234] Traditional organometallic chemistry includes the carbon compounds of silicon (see organosilicon).[256]

Germanium

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

Germanium appears as a shiny grey-white solid.[257] It is about one-third less dense than iron, hard and brittle.[258] It is mostly unreactive at room temperature[n 22] but is slowly attacked by hot concentrated sulfuric or nitric acid.[260] Germanium also reacts with molten caustic soda to yield sodium germanate Na2GeO3, together with the evolution of hydrogen.[261] It melts at a lower temperature than steel, 938 °C vs. ~1400 °C. Germanium is a semiconductor with an electrical conductivity of around 2 × 10−2 S•cm−1[260] and a band gap of 0.67 eV.[262] Liquid germanium is a metallic conductor, with an electrical conductivity on par with that of liquid mercury.[263]

Most of the chemistry of germanium is characteristic of a nonmetal.[264] It forms alloys with aluminium and gold.[265] Germanium shows fewer tendencies to anionic behaviour than ordinary nonmetals.[248] Its solution chemistry is characterised by the formation of oxyanions.[249] Germanium generally forms tetravalent (IV) compounds, and it can also form less stable divalent (II) compounds, in which it behaves more like a metal.[266] Germanium analogues of all of the major types of silicates have been prepared.[267] The metallic character of germanium is also suggested by the formation of various 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.[268] The oxide GeO2 is polymeric,[232] amphoteric,[269] and a glass former.[234] The dioxide is soluble in acidic solutions (and the monoxide GeO, is even more so), and this is sometimes used as a basis to classify germanium as a metal.[270] Up to the 1930s germanium was considered to be a poorly conducting metal rather than a nonmetal.[271] As is the case with all the elements commonly recognised as metalloids, germanium has an established organometallic chemistry (see organogermanium chemistry).[272]

Arsenic

Two dull silver clusters of crystalline shards.
Arsenic, sealed in a container to prevent tarnishing

Arsenic is a grey, metallic looking solid. It is about one-third less dense than iron, brittle, and moderately hard (more than aluminium; less than iron).[273] It is stable in dry air but develops a golden bronze patina in moist air, which blackens on further exposure. Arsenic is attacked by nitric acid and concentrated sulfuric acid. It reacts with fused caustic soda to give the arsenate Na3AsO3, together with the evolution of hydrogen.[274] Arsenic sublimes at 615 °C. The vapour is lemon-yellow and smells like garlic.[275] Arsenic only melts under a pressure of 38.6 atm, at 817 °C.[276] It is a semimetal with an electrical conductivity of around 3.9 × 104 S•cm−1[277] and a band overlap of 0.5 eV.[278][n 23] Liquid arsenic is a semiconductor with a band gap of 0.15 eV.[280]

The chemistry of arsenic is predominately nonmetallic,[281] and with many metals it forms alloys; most of these are brittle.[282] Arsenic shows fewer tendencies to anionic behaviour than ordinary nonmetals.[248] Its solution chemistry is characterised by the formation of oxyanions.[249] Arsenic generally forms compounds in which it has an oxidation state of +3 or +5.[283] The halides, and the oxides and their derivatives are illustrative examples.[251] In the trivalent state, arsenic shows some incipient metallic properties.[284] The halides are hydrolysed by water but these reactions, particularly those of the chloride, are reversible with the addition of a hydrohalic acid.[285] The oxide is acidic but weakly amphoteric. The higher, less stable, pentavalent state has strongly acidic (nonmetallic) properties.[286] Compared to phosphorus, the stronger metallic character of arsenic is indicated by the formation of oxoacid salts such as AsPO4, As2(SO4)3[n 24] and arsenic acetate As(CH3COO)3.[290] The oxide As2O3 is polymeric,[232] amphoteric,[291][n 25] and a glass former.[234] Arsenic has an extensive organometallic chemistry (see organoarsenic chemistry).[294]

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.[274] It is about 15% less dense than iron, brittle, and moderately hard (more so than arsenic; less so than iron; about the same as copper).[273] It is stable in air and moisture at room temperature. It is attacked by concentrated nitric acid, yielding the hydrated pentoxide Sb2O5. Aqua regia gives the pentachloride SbCl5 and (hot) concentrated sulfuric acid results in the sulfate Sb2(SO4)3.[295] It is not affected by molten alkali.[296] Antimony is capable of displacing hydrogen from water, when heated: 2Sb + 3 H2O → Sb2O3 + 3 H2.[297] It melts at 630.63 °C. Antimony is a semimetal with an electrical conductivity of around 3.1 × 104 S•cm−1[298] and a band overlap of 0.16 eV.[278][n 26] Liquid antimony is a metallic conductor with an electrical conductivity of around 5.3 × 104 S•cm−1.[300]

Most of the chemistry of antimony is characteristic of a nonmetal.[301] It forms alloys with one or more metals such as aluminium,[302] iron, nickel, copper, zinc, tin, lead and bismuth.[303] Antimony has fewer tendencies to anionic behaviour than ordinary nonmetals.[248] Its solution chemistry is characterised by the formation of oxyanions.[249] Like arsenic, antimony generally forms compounds in which it has an oxidation state of +3 or +5.[283] The halides, and the oxides and their derivatives are illustrative examples.[251] The +5 state is less stable than the +3, but relatively easier to attain than with arsenic. This is explained by the poor shielding afforded the arsenic nucleus by its 3d10 electrons. In comparison, the tendency of antimony to oxidize more easily partially offsets the effect of its 4d10 shell.[304] Tripositive antimony is amphoteric; pentapositive antimony is (predominately) acidic.[305] Consistent with an increase in metallic character down group 15, antimony forms salts or salt-like compounds including a nitrate Sb(NO3)3, phosphate SbPO4, sulfate Sb2(SO4)3 and perchlorate Sb(ClO4)3.[306] 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
.[307] The oxide Sb2O3 is a polymeric,[232] amphoteric,[308] and a glass former.[234] Antimony has an extensive organometallic chemistry (see organoantimony chemistry).[309]

Tellurium

A shiny silver-white medallion with a striated surface, irregular around the outside, with a square spiral-like pattern in the middle.
Tellurium, described by Dmitri Mendeleev as forming a transition between metals and nonmetals[310]

Tellurium is a silvery-white shiny solid.[311] It is about 15% less dense than iron, brittle, and the softest of the commonly recognised metalloids, being marginally harder than sulfur.[273] 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.[312] It reacts (to varying degrees) with nitric, sulfuric and hydrochloric acids to give compounds such as the sulfoxide TeSO3 or tellurous acid H2TeO3,[313] the basic nitrate (Te2O4H)+(NO3),[314] or the oxide sulfate Te2O3(SO4).[315] It dissolves in boiling alkalis, with the formation of 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.[316] At higher temperatures tellurium is sufficiently plastic to extrude.[317] It melts at ca. 450 °C. Crystalline tellurium has a structure consisting of parallel infinite spiral chains. Whereas the bonding between adjacent atoms in a chain is covalent, there is evidence of a weak metallic interaction between the neighbouring atoms of different chains.[318] Tellurium is a semiconductor with an (intrinsic) electrical conductivity of around 1.0 S•cm−1[319] and a band gap of 0.32 to 0.38 eV.[320] Liquid tellurium is a semiconductor, with an electrical conductivity, on melting, of around 1.9 × 103 S•cm−1[320] Superheated liquid tellurium is a metallic conductor.[321]

Most of the chemistry of tellurium is characteristic of a nonmetal.[322] It forms alloys with aluminium, silver and tin.[323] Tellurium shows fewer tendencies to anionic behaviour than ordinary nonmetals.[248] Its solution chemistry is characterised by the formation of oxyanions.[249] Tellurium generally forms compounds in which it has an oxidation state of −2, +4 or +6, with the tetrapositive state being the most stable.[312] It combines easily with most other elements to form binary tellurides XxTey these representing the most common mineral form. Non-stoichiometry is frequently encountered. This is particularly so with the transition metals, where electronegativity differences are small and irregular valence is favoured. Many of the associated tellurides can be treated as metallic alloys.[324] The increase in metallic character evident in tellurium, as compared to the lighter chalcogens, is further reflected in the reported formation of various other oxyacid salts, such as a basic selenate 2TeO2·SeO3 and an analogous perchlorate and periodate 2TeO2·HXO4.[325] Tellurium forms a polymeric,[232] amphoteric,[308] glass-forming oxide[234] TeO2. The latter is a 'conditional' glass-forming oxide—it forms a glass with a very small amount of additive.[234] Tellurium has an extensive organometallic chemistry (see organotellurium chemistry).[326]

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.[327]

Carbon is ordinarily classified as a nonmetal[328] but has some metallic properties and is occasionally classified as a metalloid.[329] The properties summarised in the following paragraphs are for hexagonal graphitic carbon, the most thermodynamically stable form of carbon under ambient conditions.[330]

Carbon has a lustrous appearance[331] and is a fairly good electrical conductor.[332] Its conductivity in the direction of its planes decreases as the temperature is raised, like a metal;[333][n 27] it has the electronic band structure of a semimetal.[333] The allotropes of carbon, including graphite, can accept foreign atoms or compounds into their structures via substitution, intercalation or doping (interstitial or intrastitial) with the resulting materials being referred to as 'carbon alloys'.[337] Carbon can form ionic salts, including a sulfate, perchlorate, nitrate, hydrogen selenate, and hydrogen phosphate;[338][n 28] 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.[339]

Carbon is brittle[340] like nonmetals and behaves as a semiconductor perpendicular to the direction of its planes.[333] Most of its chemistry is nonmetallic;[341] it has a relatively high ionization energy[342] and, compared to most metals, a relatively high electronegativity.[343] Carbon can form anions such as C4– (methanide), C2–
2
(acetylide) and C3–
4
(sesquicarbide or allylenide), in compounds with metals of main groups 1–3, and with the lanthanides and actinides.[344] Its oxide CO2 forms a medium-strength carbonic acid H2CO3.[345][n 29]

Aluminium

A silvery white steam-iron shaped lump with semi-circular striations along the width of its top surface and rough furrows in the middle portion of its left edge.
People handling high purity aluminium for the first time often question if it really is aluminium since it is very much softer than its familiar alloys.[347]

Aluminium is classified as a metal. It is lustrous, malleable and ductile, has high electrical and thermal conductivity and a close-packed crystalline structure.[348]

It has some properties that are unusual for a metal; taken together,[349] these properties are sometimes used as a basis to classify aluminium as a metalloid.[350] Its crystalline structure shows some evidence of directional bonding.[351] Although it forms an Al3+ cation in some compounds, aluminium bonds covalently in most others.[352] Its oxide is amphoteric,[353] and a conditional glass-former.[234] Aluminium can form anionic aluminates,[349] such behaviour being considered nonmetallic in character.[71]

Stott[354] labels aluminium as a weak metal. It has the physical properties of a metal but some of the chemical properties of a nonmetal. Steele[355] notes the paradoxical chemical behaviour of aluminium. It resembles a weak metal with its amphoteric oxide and the covalent character of many of its compounds. Yet it is also a strongly electropositive metal, with a high negative electrode potential.

The notion of aluminium as a metalloid is sometimes disputed[356] as it has many metallic properties. Aluminium is therefore, arguably, an exception to the mnemonic that elements adjacent to the metal-nonmetal dividing line are metalloids.[357][n 30]

Selenium

A small glass jar filled with small dull grey concave buttons. The pieces of selenium look like tiny mushrooms without their stems.
Selenium. Being a photoconductor, grey selenium conducts electricity around 1,000 times better when light falls on it, a property used since the mid-1870s in various light-sensing applications.[359]

Selenium shows borderline metalloid or nonmetal behaviour.[360][n 31]

Its most stable form, the grey trigonal allotrope, is sometimes called 'metallic' selenium. This is because its electrical conductivity is several orders of magnitude greater than that of the red monoclinic form.[363] The metallic character of selenium is further shown by its lustre[364] and its crystalline structure, the latter of which is thought to include weakly 'metallic' interchain bonding.[365] Selenium can be drawn into thin threads when molten.[366] It shows reluctance to acquire 'the high positive oxidation numbers characteristic of nonmetals'.[367] It can form cyclic polycations (such as Se2+
8
) when dissolved in oleums[368] (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
.[369]

The nonmetallic character of selenium is shown by its brittleness[364] and the low electrical conductivity (~10−9 to 10−12 S•cm−1) of its highly purified form.[107] This is comparable to or less than that of bromine (7.95×10–12 S•cm−1),[370] a nonmetal. Selenium has the electronic band structure of a semiconductor[371] and retains its semiconducting properties in liquid form.[371] It has a relatively high[372] electronegativity (2.55 revised Pauling scale). Its reaction chemistry is mainly that of its nonmetallic anionic forms Se2–, SeO2−
3
and SeO2−
4
.[373]

Selenium is commonly described as a metalloid in the environmental chemistry literature.[374] Processes and reactions affecting its fate in the aquatic environment are similar to those found for arsenic and antimony.[375] Trace amounts of selenium are essential to human health, and its water-soluble salts (in higher concentrations) have a toxicological profile similar to that of arsenic.[376]

Polonium

A thin film of a bluish-grey metal on a stainless steel disc.
Polonium, in the form of a thin film on a stainless-steel disc

Polonium is 'distinctly metallic' in some ways.[377] Both of its allotropic forms are metallic conductors.[377] It is soluble in acids, forming the rose-coloured Po2+ cation and displacing hydrogen: Po + 2 H+ → Po2+ + H2.[378] Many polonium salts are known.[379] The oxide PoO2 is predominantly basic in nature.[380] Polonium is a reluctant oxidizing agent, unlike its lighter congener oxygen: highly reducing conditions are required for the formation of the Po2– anion in aqueous solution.[381]

Polonium also shows nonmetallic character in its halides, and by way of the existence of polonides. The halides have properties generally characteristic of nonmetal halides (being volatile, easily hydrolyzed, and soluble in organic solvents).[382] Many metal polonides, obtained by heating the elements together at 500–1,000 °C, and containing the Po2– anion, are also known.[383]

Astatine

Astatine, which is ordinarily classified as a nonmetal[384] has some marked metallic properties,[385] and may be either a metalloid,[386] or a metal.[n 32] Immediately following its production in 1940, early investigators considered it a metal.[388] In 1949 it was called the most noble (difficult to reduce) nonmetal as well as being a relatively noble (difficult to oxidize) metal.[389] In 1950 astatine was described as a halogen and (therefore) a reactive nonmetal.[390] In 2013, on the basis of relativistic modelling, astatine was predicted to be a monatomic metal, having a face-centred cubic crystalline structure.[391]

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.[392][n 33] Astatine may be metallic in the liquid state on the basis that elements with an enthalpy of vaporization (EoV) greater than ~42 kJ/mol are metallic when liquid.[394] Such elements include boron,[n 34] silicon, germanium, antimony, selenium and tellurium. Estimated values for the EoV of diatomic astatine are 50 kJ/mol or higher;[398] diatomic iodine, with an EoV of 41.71,[399] falls just short of the threshold figure. '[L]ike typical metals, it [astatine] is precipitated by hydrogen sulfide even from strongly acid solutions and is displaced in a free form from sulfate solutions; it is deposited on the cathode on electrolysis'.[400][n 35] Further indications of a tendency for astatine to behave like a (heavy) metal are: '...the formation of pseudohalide compounds...complexes of astatine cations...complex anions of trivalent astatine...as well as complexes with a variety of organic solvents'.[402] It has also been argued that it demonstrates cationic behaviour, by way of stable At+ and AtO+ forms, in strongly acidic aqueous solutions.[403]

Some of astatine's reported properties are nonmetallic. It has the narrow liquid range ordinarily associated with nonmetals (mp 575 K, bp 610).[404] Batsanov gives a calculated band gap energy for astatine of 0.7 eV;[405] this is consistent with nonmetals (in physics) having separated valence and conduction bands and thereby being either semiconductors or insulators.[406] The chemistry of astatine in aqueous solution is predominately characterised by the formation of various anionic species.[407] Most of its known compounds resemble those of iodine,[408] which is a halogen and a nonmetal.[409] Such compounds include astatides (XAt), astatates (XAtO3), and monovalent interhalogen compounds.[410]

Restrepo et al.[411] reported that astatine appeared to be more like polonium than the established halogens. They did so on the basis of detailed comparative studies of the known and interpolated properties of 72 elements.

Honourable mentions

Elements near the metalloids

Shiny violet-black coloured crystalline shards.
Iodine crystals, showing a metallic lustre. Iodine is a semiconductor in the direction of its planes, with a band gap of ~1.3 eV. It has an electrical conductivity of 1.7 × 10−8 S•cm−1 at room temperature,[412] higher than selenium but lower than boron, the least electrically conducting of the recognised metalloids.[n 36]

Some of the elements adjacent in the periodic table to the commonly recognised metalloids, although usually classified as either metals or nonmetals, are occasionally referred to as near-metalloids[415] or noted for their metalloidal character. To the left of the metal-nonmetal dividing line, such elements include gallium,[416] tin[417] and bismuth.[418] They show 'odd' packing structures,[419] marked covalent chemistry (molecular or polymeric),[420] and amphoterism.[421] To the right of the dividing line are carbon,[422] phosphorus,[423] selenium[424] and iodine.[425] They exhibit metallic lustre, semiconducting properties[n 37] 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 38] and selenium as grey selenium.

Allotropes

Many small, shiny, silver-coloured spheres on the left; many of the same sized spheres on the right are duller and darker than the ones of the left and have a subdued metallic shininess.
White tin (left) and 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.[431] The existence of such allotropes can complicate the classification of the elements involved.[432]

Tin, for example, has two allotropes: tetragonal 'white' β-tin and cubic 'grey' α-tin, as shown in the picture. White tin is a silvery-white, very shiny, ductile and malleable metal. It is the stable form of tin at or above room temperature and has an electrical conductivity of 9.17×104 S·cm−1 (~1/6th that of copper).[433] Grey tin usually has the appearance of a grey micro-crystalline powder although it can also be prepared in crystalline or polycrystalline forms having a semi-lustrous appearance and a brittle comportment. It is the stable form of tin below 13.2 °C (56 °F) and has an electrical conductivity of between (2–5)×102 S·cm−1 (~1/250th that of white tin).[434] Grey tin has the same crystalline structure as that of the diamond allotrope of carbon. It behaves as a semiconductor (with a band gap of 0.08 eV), but has the electronic band structure of a semimetal.[435] It has been referred to as either a very poor metal,[436] a metalloid,[437] a nonmetal[438] or a near metalloid.[418]

The diamond allotrope of carbon is clearly nonmetallic, being translucent and having a relatively poor electrical conductivity of 10−14 to 10−16 S·cm−1.[439] Graphite has an electrical conductivity of 3×104 S·cm−1,[440] a figure more characteristic of a metal. Phosphorus, sulfur, arsenic, selenium, antimony and bismuth also have less stable allotropes that display borderline or either more or less metallic or nonmetallic behaviour.[441]

Notes

  1. ^ For a related commentary see also: Vernon RE 2013, 'Which Elements Are Metalloids?', Journal of Chemical Education, vol. 90, no. 12, pp. 1703–1707, doi:10.1021/ed3008457
  2. ^ On the fuzziness of metalloids see, for example: Rouvray;[4] Cobb & Fetterolf;[5] and Fellet.[6] For the 'buffer zone' terminology see Rochow.[7]
  3. ^ Gold, for example, has mixed properties but is still recognised as 'king of metals.' Besides metallic behaviour (such as high electrical conductivity, and cation formation), gold shows marked nonmetallic behaviour: On halogen character, see also Belpassi et al.[20] 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.[21]
  4. ^ Mann et al.[24] refer to these elements as 'the recognized metalloids'.
  5. ^ Rochow[47] concluded that no single measurement indicates exactly which elements are properly classified as metalloids, and that, therefore, present-day students and teachers usually agree to use electronegativity as a compromise criterion. He described metalloids as a collection of in between elements, of electronegativity 1.8 to 2.2 (classical Pauling scale), that "...resemble metals, yet are not completely metallic either in appearance or in properties," and "...are neither metals nor nonmetals." In mentioning 'appearance', Rochow referred to the intermediate reflectivity values of the commonly recognised metalloids,[48] compared to the intermediate to typically high[49] values of the metals,[50] and the zero or low (mostly)[51] to intermediate reflectivity values of the non-metals.[52] See also, for example:
    • Hill and Hollman,[13] who 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'.
    • Bond,[53] who suggests that 'one criterion for distinguishing semi-metals from true metals under normal conditions is that the co-ordination number of the former is never greater than eight'.
    • Edwards et al.,[54] who state that, 'Using the Goldhammer-Herzfeld criterion with measured atomic electronic polarizabilities and condensed phase molar volumes allows one to readily predict which elements are metallic, which are nonmetallic, and which are borderline when in their condensed phases (solid or liquid).'
  6. ^ Selenium has an ionization energy (IE) of ~226 kcal/mol and is sometimes described as a semiconductor. It has a relatively high 2.55 electronegativity (EN). Polonium has an IE of ~196 kcal/mol and a 2.0 EN, but has a metallic band structure.[57] Astatine has an estimated IE of ~210±10 kcal/mol[58] and an EN of 2.2. Its electronic band structure is not known with any great certainty.
  7. ^ The Goldhammer-Herzfeld criterion is a ratio that compares the force holding an individual atom's valence electrons in place with the forces, acting on the same electrons, arising 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. Metallic behaviour is then predicted.[60] Otherwise nonmetallic behaviour is anticipated.
  8. ^ Gallium is unusual (for a metal) in having a packing efficiency of just 39%.[63] Other notable values are 42.9 for bismuth[64] and 58.5 for liquid mercury.[65]
  9. ^ As the ratio is based on classical arguments[68] it does not accommodate the finding that polonium, which has a value of ~0.95, adopts a metallic (rather than covalent) crystalline structure, on relativistic grounds.[69] It nevertheless offers a relatively simple first order rationalization for the occurrence of metallic character amongst the elements.[70]
  10. ^ Jones[81] 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'.
  11. ^ Oderberg[85] argues on ontological grounds that anything not a metal is therefore a nonmetal, and that this includes semi-metals (i.e. metalloids).
  12. ^ Copernicium is reportedly the only metal known to be a gas at room temperature.[100]
  13. ^ Metals have electrical conductivity values of from 6.9 × 103 S•cm−1 for manganese to 6.3 × 105 for silver.[104]
  14. ^ Metalloids have electrical conductivity values of from 1.5 × 10−6 S•cm−1 for boron to 3.9 × 104 for arsenic.[106] If selenium is included as a metalloid the applicable conductivity range would start from ~10−9 to 10−12 S•cm−1.[107]
  15. ^ Nonmetals have electrical conductivity values of from ~10−18 S•cm−1 for the elemental gases to 3 × 104 in graphite.[108]
  16. ^ Chedd[115] defines metalloids as having electronegativity values of 1.8 to 2.2 (Allred-Rochow scale). He included boron, silicon, germanium, arsenic, antimony, tellurium, polonium and astatine in this category. In reviewing Chedd's work, Adler[116] described this choice as arbitrary, as other elements with electronegativities in this range include copper, silver, phosphorus, mercury and bismuth. He went on to suggest defining a metalloid as 'a semiconductor or semimetal' and included bismuth and selenium in his book.
  17. ^ Olmsted and Williams[120] 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'.
  18. ^ Research published in 2012 suggests that metal-metalloid glasses can be characterised by an interconnected atomic packing scheme in which metallic and covalent bonding structures coexist.[170]
  19. ^ 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).[220]
  20. ^ On the analogy between boron and metals, Greenwood[229] 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...'.
  21. ^ Although SiO2 is classified as an acidic oxide, and hence reacts with alkalis to give silicates, its reaction with phosphoric acid yields a silicon oxide orthophosphate Si5O(PO4)6,[254] and with hydrofluoric acid to give hexafluorosilicic acid H2SiF6.[255]
  22. ^ Temperatures above 400 °C are required to form a noticeable surface oxide layer.[259]
  23. ^ Arsenic also exists as a naturally occurring (but rare) allotrope (arsenolamprite), a semiconductor with a band gap of around 0.3 eV or 0.4 eV. It can be prepared in a semiconducting amorphous form, with a band gap of around 1.2–1.4 eV.[279]
  24. ^ The formulae of AsPO4 and As2(SO4)3 suggest straightforward ionic formulations, with As3+, but compounds in which arsenic is present as a cation are extremely rare.[287] AsPO4, 'which is virtually a covalent oxide,' has been referred to as a double oxide, of the form As2O3·P2O5. It comprises AsO3 pyramids and PO4 tetrahedra, joined together by all their corner atoms to form a continuous polymeric network.[288] As2(SO4)3 has a structure in which each SO4 tetrahedron is bridged by two AsO3 trigonal pyramida.[289]
  25. ^ As2O3 is usually regarded as being amphoteric but a few sources say it is (weakly)[292] acidic. They describe its 'basic' properties (that is, its reaction with concentrated hydrochloric acid to form arsenic trichloride) as being alcoholic, by analogy with the formation of covalent alklyl chlorides by covalent alcohols (e.g., R-OH + HCl RCl + H2O)[293]
  26. ^ Antimony can also be prepared in an amorphous semiconducting black form, with an estimated (temperature-dependent) band gap of 0.06–0.18 eV.[299]
  27. ^ Liquid carbon may[334] or may not[335] be a metallic conductor, depending on pressure and temperature; see also.[336]
  28. ^ For the sulfate, the method of preparation is (careful) direct oxidation of graphite in concentrated sulfuric acid by an 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. Analogous salts are formed in other strong acids, such as perchloric acid.[338]
  29. ^ Only a very small fraction of dissolved CO2 is present in water as carbonic acid so, even though H2CO3 is actually a medium-strong acid, solutions of carbonic acid are only weakly acidic.[346]
  30. ^ A mnemonic that captures the elements commonly recognised as metalloids goes: Up, up-down, up-down, up...are the metalloids![358]
  31. ^ Rochow,[361] who later wrote his 1966 monograph The metalloids,[362] commented that, 'In some respects selenium acts like a metalloid and tellurium certainly does'.
  32. ^ A third option is to include astatine both as a nonmetal and as a metalloid.[387]
  33. ^ A visible piece of astatine would be immediately and completely vaporized because of the heat generated by its intense radioactivity.[393]
  34. ^ The literature is contradictory as to whether boron exhibits metallic conductivity in liquid form. Krishnan et al.[395] found that liquid boron behaved like a metal. Glorieux et al.[396] characterised liquid boron as a semiconductor, on the basis of its low electrical conductivity. Millot et al.[397] reported that the emissivity of liquid boron was not consistent with that of a liquid metal.
  35. ^ Korenman[401] similarly noted that 'the ability precipitate with hydrogen sulfide distinguishes astatine from other halogens and brings it closer to bismuth and other heavy metals.'
  36. ^ 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).[413] This is thought due to electronic interactions between the molecules in each layer of iodine, which in turn give rise to its semiconducting properties and shiny appearance.[414]
  37. ^ For example: intermediate electrical conductivity;[426] a relatively narrow band gap;[427] light sensitivity.[426]
  38. ^ White phosphorus is the most common, industrially important,[428] and easily reproducible allotrope. For those reasons it is the standard state of the element.[429] Paradoxically, it is also thermodynamically the least stable, as well as the most volatile and reactive form.[430]

Citations

  1. ^ Chedd 1969, pp. 58, 78; National Research Council 1984, p.  43
  2. ^ a b Atkins et al. 2010, p. 20
  3. ^ a b Roher 2001, pp. 4–6
  4. ^ Rouvray 1995, p. 546. Rouvray submits that classifying the electrical conductivity of the elements using the overlapping domains of metals, metalloids, and nonmetals better reflects reality than a strictly black or white paradigm.
  5. ^ Cobb & Fetterolf 2005, p. 64: 'The division between metals and nonmetals is rather fuzzy, so the elements in the immediate vicinity of the zigzag staircase line are called metalloids, which means they don't fit either definition exactly.'
  6. ^ Fellet 2011: 'Chemistry has all sorts of fuzzy definitions'.
  7. ^ Rochow 1977, p. 14
  8. ^ Goldsmith 1982, p. 526; Hawkes 2001, p. 1686
  9. ^ Hawkes 2001, p. 1687
  10. ^ Sharp 1981, p. 299
  11. ^ Cusack 1987, p. 360
  12. ^ Kelter, Mosher & Scott 2009, p. 268
  13. ^ a b Hill & Holman 2000, p. 41
  14. ^ King 1979, p. 13
  15. ^ Moore 2011, p. 81
  16. ^ Gray 2010
  17. ^ a b Hopkins & Bailar 1956, p. 458
  18. ^ Glinka 1965, p. 77
  19. ^ Wiberg 2001, p. 1279
  20. ^ Belpassi et al. 2006, pp. 4543–4
  21. ^ Schmidbaur & Schier 2008, pp. 1931–51
  22. ^ Tyler Miller 1987, p. 59
  23. ^ Goldsmith 1982, p. 526; Hawkes 2001, p. 1686; Boylan 1962, p. 493; Sherman & Weston 1966, p. 64; Wulfsberg 1991, p. 201; Kotz, Treichel & Weaver 2009, p. 62
  24. ^ a b Mann et al. 2000, p. 2783
  25. ^ Hawkes 2001, p. 1686; Segal 1989, p. 965; McMurray & Fay 2009, p. 767
  26. ^ Bucat 1983, p. 26; Brown c. 2007
  27. ^ a b Swift & Schaefer 1962, p. 100
  28. ^ Hawkes 2001, p. 1686; Hawkes 2010; Holt, Rinehart & Wilson c. 2007
  29. ^ 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).
  30. ^ Tilden 1876, pp. 172, 198–201; Smith 1994, p. 252; Bodner & Pardue 1993, p. 354
  31. ^ Bassett et al. 1966, p. 127
  32. ^ Rausch 1960
  33. ^ Thayer 1977, p. 604; Warren & Geballe 1981; Masters & Ela 2008, p. 190
  34. ^ Warren & Geballe 1981; Chalmers 1959, p. 72; US Bureau of Naval Personnel 1965, p. 26
  35. ^ Siebring 1967, p. 513
  36. ^ Wiberg 2001, p. 282
  37. ^ Rausch 1960; Friend 1953, p. 68
  38. ^ Murray 1928, p. 1295
  39. ^ Hampel & Hawley 1966, p. 950; Stein 1985; Stein 1987, pp. 240, 247–8
  40. ^ Hatcher 1949, p. 223; Secrist & Powers 1966, p. 459
  41. ^ Taylor 1960, p. 614
  42. ^ Considine & Considine 1984, p. 568; Cegielski 1998, p. 147; The American heritage science dictionary 2005 p. 397
  43. ^ Woodward 1948, p. 1
  44. ^ NIST 2010. Values shown in the above table have been converted from the NIST values, which are given in eV.
  45. ^ Berger 1997; Lovett 1977, p. 3
  46. ^ a b Goldsmith 1982, p. 526
  47. ^ Rochow 1966, pp. 1, 4–7
  48. ^ Lagrenaudie 1953; Rochow 1966, pp. 23, 25
  49. ^ Askeland, Fulay & Wright 2011, p. 806
  50. ^ Born & Wolf 1999, p. 746
  51. ^ Burakowski & Wierzchoń 1999, p. 336
  52. ^ Olechna & Knox 1965, pp. A991‒2
  53. ^ Bond 2005, p. 3
  54. ^ Edwards et al. 2010, p. 958
  55. ^ Jones 2010, p. 169
  56. ^ 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, 'since very little is known about their chemical and physical properties, and such classification must be rather arbitrary.'
  57. ^ Kraig, Roundy & Cohen 2004, p. 412; Alloul 2010, p. 83
  58. ^ NIST 2011. They cite Finkelnburg & Humbach (1955) who give a figure of 9.2±0.4 eV = 212.2±9.224 kcal/mol.
  59. ^ Edwards & Sienko 1983, p. 695; Edwards et al. 2010
  60. ^ Herzfeld 1927; Edwards 2000, pp. 100–3
  61. ^ Van Setten et al. 2007, pp. 2460–1; Russell & Lee 2005, p. 7 (Si, Ge); Pearson 1972, p. 264 (As, Sb, Te; also black P)
  62. ^ Russell & Lee 2005, p. 1
  63. ^ Russell & Lee 2005, pp. 6–7, 387
  64. ^ a b Pearson 1972, p. 264
  65. ^ Okakjima & Shomoji 1972, p. 258
  66. ^ Kitaĭgorodskiĭ 1961, p. 108
  67. ^ a b c Neuburger 1936
  68. ^ Edwards 1999, p. 416
  69. ^ Steurer 2007, p. 142; Pyykkö 2012, p. 56
  70. ^ Edwards & Sienko 1983, p. 695
  71. ^ a b Hamm 1969, p. 653
  72. ^ Horvath 1973, p. 336
  73. ^ a b Gray 2009, p.  9
  74. ^ Booth & Bloom 1972, p. 426; Cox 2004, pp. 17, 18, 27–8; Silberberg 2006, p. 305–13
  75. ^ Cox 2004, pp. 17–18, 27–8; Silberberg 2006, p. 305–13
  76. ^ Rayner-Canham 2011
  77. ^ Emsley 1971, p. 1
  78. ^ James et al. 2000, p. 480
  79. ^ Kneen, Rogers & Simpson 1972, pp. 218–220
  80. ^ 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...'.
  81. ^ Jones 2010, p. 170
  82. ^ Tyler 1948, p. 105; Reilly 2002, pp. 5–6
  83. ^ Hampel & Hawley 1976, p. 174
  84. ^ Goodrich 1844, p. 264; The Chemical News 1897, p. 189; Hampel & Hawley 1976, p. 191; Lewis 1993, p. 835; Hérold 2006, pp. 149–50
  85. ^ Oderberg 2007, p. 97
  86. ^ Brown & Holme 2006, p. 57
  87. ^ Wiberg 2001, p. 282; Simple Memory Art c. 2005
  88. ^ Chedd 1969, pp. 12–13
  89. ^ Nickelès 1861, p. 416; US Air Force Medical Service 1966, p. 3-3
  90. ^ Brady, Senese & Jespersen 2009, p. 53
  91. ^ Remy 1956, p. 1
  92. ^ Johnston 1992, p. 57
  93. ^ Malerba 1985, p. 13
  94. ^ Rochow 1966, p. 14
  95. ^ Boikess & Edelson 1985, p. 85
  96. ^ Aldridge 1998, p. 290
  97. ^ Hawkes 2001, p. 1686
  98. ^ Kneen, Rogers & Simpson, 1972, p. 263. Columns 2 and 4 are sourced from this reference unless otherwise indicated.
  99. ^ Stoker 2010, p. 62: Chang 2002, p. 304. Chang speculates that the melting point of francium would be about 23 °C.
  100. ^ New Scientist 1975; Soverna 2004; Eichler, Aksenov & Belozeroz et al. 2007; Austen 2010
  101. ^ a b Rochow 1966, p. 4
  102. ^ Hunt 2000, p. 256
  103. ^ McQuarrie & Rock 1987, p. 85
  104. ^ Desai, James & Ho 1984, p. 1160; Matula 1979, p. 1260
  105. ^ Choppin & Johnsen 1972, p. 351
  106. ^ Schaefer 1968, p. 76; Carapella 1968, p. 30
  107. ^ a b Kozyrev 1959, p. 104; Chizhikov & Shchastlivyi 1968, p. 25; Glazov, Chizhevskaya & Glagoleva 1969, p. 86
  108. ^ Bogoroditskii & Pasynkov 1967, p. 77; Jenkins & Kawamura 1976, p. 88
  109. ^ Hampel & Hawley 1976, p. 191; Wulfsberg 2000, p. 620
  110. ^ Swalin 1962, p. 216
  111. ^ Bailar et al. 1989, p. 742
  112. ^ Metcalfe, Williams & Castka 1974, p. 86
  113. ^ Chang 2002, p. 306
  114. ^ Pauling 1988, p. 183
  115. ^ Chedd 1969, pp. 24–5
  116. ^ Adler 1969, pp. 18–19
  117. ^ Hultgren 1966, p. 648; Young & Sessine 2000, p. 849; Bassett et al. 1966, p. 602
  118. ^ Rochow 1966, p. 4; Atkins et al. 2006, pp. 8, 122–3
  119. ^ Russell & Lee 2005, pp. 421, 423; Gray 2009, p. 23
  120. ^ Olmsted & Williams 1997, p. 975
  121. ^ a b c Russell & Lee 2005, p. 401; Büchel, Moretto & Woditsch 2003, p. 278
  122. ^ Desch 1914, p. 86
  123. ^ Phillips & Williams 1965, p. 620
  124. ^ Van der Put 1998, p. 123
  125. ^ Klug & Brasted 1958, p. 199
  126. ^ Good et al. 1813
  127. ^ Sequeira 2011, p. 776
  128. ^ Gary 2013
  129. ^ Russell & Lee 2005, pp. 423–4; 405–6
  130. ^ Davidson & Lakin 1973, p. 627
  131. ^ Wiberg 2001, p. 589
  132. ^ Greenwood & Earnshaw 2002, p. 749; Schwartz 2002, p. 679
  133. ^ Antman 2001
  134. ^ Řezanka & Sigler 2008; Sekhon 2012
  135. ^ Emsley 2001, p. 67
  136. ^ Zhang et al. 2008, p. 360
  137. ^ a b Science Learning Hub 2009
  138. ^ Büchel 1983, p. 226
  139. ^ Emsley 2001, p. 391
  140. ^ Schauss 1991; Tao & Bolger 1997
  141. ^ Eagleson 1994, p. 450; EVM 2003, pp. 197‒202
  142. ^ a b Nielsen 1998
  143. ^ a b Jaouen & Gibaud 2010
  144. ^ Stevens & Klarner, p. 205
  145. ^ Keall, Martin and Tunbridge 1946
  146. ^ Emsley 2001, p. 426
  147. ^ Oldfield et al. 1974, p. 65; Turner 2011
  148. ^ Ba et al. 2010; Daniel-Hoffmann, Sredni & Nitzan 2012; Molina-Quiroz et al. 2012
  149. ^ Peryea 1998
  150. ^ Hager 2006, p. 299
  151. ^ Apseloff 1999
  152. ^ Trivedi, Yung & Katz 2013, p. 209
  153. ^ Thomas, Bialek & Hensel 2013, p. 1
  154. ^ Le Bras, Wilkie & Bourbigot 2005, p. v
  155. ^ Wilkie & Morgan 2009, p. 187
  156. ^ Locke et al. 1956, p. 88
  157. ^ Carlin 2011, p. 6.2
  158. ^ Evans 1993, pp.  257–8
  159. ^ Corbridge 2013, p. 1149
  160. ^ Kaminow & Li 2002, p. 118
  161. ^ Deming 1925, pp. 330 (As2O3), 418 (B2O3; SiO2; Sb2O3); Witt & Gatos 1968, p. 242 (GeO2)
  162. ^ Eagleson 1994, p. 421 (GeO2); Rothenberg 1976, 56, 118–19 (TeO2)
  163. ^ Geckeler 1987, p. 20
  164. ^ Kreith & Goswami 2005, p. 12–109
  165. ^ Russell & Lee 2005, p. 397
  166. ^ Butterman & Jorgenson 2005, pp. 9–10
  167. ^ Butterman & Carlin 2004, p. 22; Russell & Lee 2005, p. 422
  168. ^ Träger 2007, pp. 438, 958; Eranna 2011, p. 98
  169. ^ Rao 2002, p. 552; Löffler, Kündig & Dalla Torre 2007, p. 17-11
  170. ^ Guan et al. 2012; WPI-AIM 2012
  171. ^ Klement, Willens & Duwez 1960; Wanga, Dongb & Shek 2004, p. 45
  172. ^ Demetriou et al 2011; Oliwenstein 2011
  173. ^ Karabulut et al. 2001, p. 15; Haynes 2012, p. 4-26
  174. ^ Schwartz 2002, pp. 679–680
  175. ^ Carter & Norton 2013, p. 403
  176. ^ Maeder 2013, pp. 3, 9–11
  177. ^ Tominaga 2006, p. 327–8; Chung 2010, p. 285–6; Kolobov & Tominaga 2012, p. 149
  178. ^ Berger 1997, p. 91; Hampel 1968, passim
  179. ^ Rochow 1966, p. 41; Berger 1997, pp. 42–3
  180. ^ a b Bomgardner 2013, p. 20
  181. ^ Russell & Lee 2005, p. 395; Brown et al. 2009, p. 489
  182. ^ 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)
  183. ^ Bianco et. al. 2013
  184. ^ Russell & Lee 2005, pp. 421–2, 424
  185. ^ Berger 1997, p. 91
  186. ^ ScienceDaily 2012
  187. ^ Reardon 2005; Meskers, Hagelüken & Van Damme 2009, p. 1131
  188. ^ The Economist 2012
  189. ^ Whitten 2007, p. 488
  190. ^ Jaskula 2013
  191. ^ German Energy Society 2008, p. 43–44
  192. ^ Patel 2012, p. 248
  193. ^ Oxford English Dictionary 1989, 'metalloid'; Gordh, Gordh & Headrick 2003, p. 753
  194. ^ Foster 1936, pp. 212–13; Brownlee et al. 1943, p. 293
  195. ^ Calderazzo, Ercoli & Natta 1968, p. 257
  196. ^ a b Klemm 1950, pp. 133–42; Reilly 2004, p. 4
  197. ^ Walters 1982, pp. 32–3
  198. ^ Tyler 1948, p. 105
  199. ^ 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.'
  200. ^ Slade 2006, p. 16
  201. ^ Corwin 2005, p. 80
  202. ^ Barsanov & Ginzburg 1974, p. 330; Prokhorov 1983, p. 329
  203. ^ Bradbury et al. 1957, pp. 157, 659
  204. ^ Miller, Lee & Choe 2002, p. 21
  205. ^ King 2004, pp. 196–8; Ferro & Saccone 2008, p. 233
  206. ^ Pashaey & Seleznev 1973, p. 565; Gladyshev & Kovaleva 1998, p. 1445; Eason 2007, p. 294
  207. ^ Johansen & Mackintosh 1970, pp. 121–4; Divakar, Mohan & Singh 1984, p. 2337; Dávila et al. 2002, p. 035411-3
  208. ^ Jezequel & Thomas 1997, pp. 6620–6
  209. ^ 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...'
  210. ^ Pinkerton 1800, p. 81
  211. ^ Friend 1953, p. 68; IUPAC 1959, p. 10; IUPAC 1971, p. 11
  212. ^ IUPAC 2005; IUPAC 2006–
  213. ^ Van Setten et al. 2007, pp. 2460–1; Oganov et al. 2009, pp. 863–4
  214. ^ Housecroft & Sharpe 2008, p. 331; Oganov 2010, p. 212
  215. ^ Housecroft & Sharpe 2008, p. 333
  216. ^ Berger 1997, p. 37
  217. ^ Greenwood & Earnshaw 2002, p. 144
  218. ^ Kopp, Lipták & Eren 2003, p.&nbsp221
  219. ^ Prudenziati 1977, p. 242
  220. ^ Berger 1997, pp. 87, 84
  221. ^ Rayner-Canham & Overton 2006, p. 291
  222. ^ Bowser 1993, p. 393; Grimes 2011, pp. 17–18
  223. ^ Greenwood & Earnshaw 2002, p. 141; Henderson 2000, p. 58; Housecroft & Sharpe 2008, pp. 360–72
  224. ^ Parry et al. 1970, pp. 438, 448–51
  225. ^ a b Fehlner 1990, p. 202
  226. ^ a b Greenwood & Earnshaw 2002, p. 145
  227. ^ Houghton 1979, p. 59
  228. ^ Fehlner 1990, pp. 204, 207
  229. ^ Greenwood 2001, p. 2057
  230. ^ 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
  231. ^ Watt 1958, p. 387; Sharp 1983
  232. ^ a b c d e f Puddephatt & Monaghan 1989, p. 59
  233. ^ Mahan 1965, p. 485
  234. ^ a b c d e f g h Rao 2002, p. 22
  235. ^ Haiduc & Zuckerman 1985, p. 82
  236. ^ a b Greenwood & Earnshaw 2002, p. 331
  237. ^ Wiberg 2001, p. 824
  238. ^ Rochow 1973, p. 1337‒38
  239. ^ Rochow 1973, p. 1337, 1340
  240. ^ Allen & Ordway 1968, p. 152
  241. ^ Eagleson 1994, pp. 48, 127, 438, 1194; Massey 2000, p. 191
  242. ^ Orton 2004, p. 7. The listed figure is a typical value for high-purity silicon.
  243. ^ Russell & Lee 2005, p. 393
  244. ^ Coles & Caplin 1976, p. 106
  245. ^ Glazov, Chizhevskaya & Glagoleva 1969, pp. 59–63; Allen & Broughton 1987, p. 4967
  246. ^ Cotton, Wilkinson & Gaus 1995, p. 393
  247. ^ Partington 1944, p. 723
  248. ^ a b c d e Cox 2004, p. 27
  249. ^ a b c d e Hiller & Herber 1960, inside front cover; p. 225
  250. ^ Kneen, Rogers and Simpson 1972, p. 384
  251. ^ a b c Bailar, Moeller & Kleinberg 1965, p. 513
  252. ^ Cotton, Wilkinson & Gaus 1995, pp. 319, 321
  253. ^ Smith 1990, p. 175
  254. ^ Poojary, Borade & Clearfield 1993
  255. ^ Wiberg 2001, pp. 851, 858
  256. ^ Powell 1988, p. 1
  257. ^ Greenwood & Earnshaw 2002, p. 371
  258. ^ Cusack 1967, p. 193
  259. ^ Russell & Lee 2005, pp. 399–400
  260. ^ a b Greenwood & Earnshaw 2002, p. 373
  261. ^ Moody 1991, p. 273
  262. ^ Russell & Lee 2005, p. 399
  263. ^ Berger 1997, pp. 71–2
  264. ^ Jolly 1966, pp. 125–6
  265. ^ Schwartz 2002, p. 269
  266. ^ Eggins 1972, p. 66; Wiberg 2001, p. 895
  267. ^ Greenwood & Earnshaw 2002, p. 383
  268. ^ Glockling 1969, p. 38; Wells 1984, p. 1175
  269. ^ Cooper 1968, pp. 28–9
  270. ^ Steele 1966, pp. 178, 188–9
  271. ^ Cite error: The named reference Haller EE 2006, p. 3 was invoked but never defined (see the help page).
  272. ^ Wiberg 2001, p. 742
  273. ^ a b c Gray, Whitby & Mann 2011
  274. ^ a b Greenwood & Earnshaw 2002, p. 552
  275. ^ Parkes & Mellor 1943, p. 740
  276. ^ Russell & Lee 2005, p. 420
  277. ^ Carapella 1968, p. 30
  278. ^ a b Barfuß et al. 1981, p. 967
  279. ^ Greaves, Knights & Davis 1974, p. 369; Madelung 2004, pp. 405, 410
  280. ^ Bailar & Trotman-Dickenson 1973, p. 558; Li 1990
  281. ^ Bailar, Moeller & Kleinberg 1965, p. 477
  282. ^ Eagleson 1994, p. 91
  283. ^ a b Massey 2000, p. 267
  284. ^ Timm 1944, p. 454
  285. ^ Partington 1944, p. 641; Kleinberg, Argersinger & Griswold 1960, p. 419
  286. ^ Morgan 1906, p. 163; Moeller 1954, p. 559
  287. ^ Burford & Royan 1989, p. 3746
  288. ^ Corbridge 2012, pp. 122, 215
  289. ^ Douglade 1982
  290. ^ Zingaro 1994, p. 197; Emeleús & Sharpe 1959, p. 418; Addison & Sowerby 1972, p. 209; Mellor 1964, p. 337
  291. ^ Pourbaix 1974, p. 521; Eagleson 1994, p. 92; Greenwood & Earnshaw 2002, p. 572
  292. ^ Wiberg 2001, pp. 750, 975; Silberberg 2006, p. 314
  293. ^ Sidgwick 1950, p. 784; Moody 1991, pp. 248–9, 319
  294. ^ Krannich & Watkins 2006
  295. ^ Greenwood & Earnshaw 2002, p. 553
  296. ^ Dunstan 1968, p. 433
  297. ^ Parise 1996, p. 112
  298. ^ Carapella 1968a, p. 23
  299. ^ Moss 1952, pp. 174, 179
  300. ^ Dupree, Kirby & Freyland 1982, p. 604; Mhiaoui, Sar, & Gasser 2003
  301. ^ Kotz, Treichel & Weaver 2009, p. 62
  302. ^ Friend 1953, p. 87
  303. ^ Fesquet 1872, pp. 109–14
  304. ^ Greenwood & Earnshaw 2002, p. 553; Massey 2000, p. 269
  305. ^ King 1994, p.171
  306. ^ Torova 2011, p. 46
  307. ^ Pourbaix 1974, p. 530
  308. ^ a b Wiberg 2001, p. 764
  309. ^ House 2008, p. 497
  310. ^ Mendeléeff 1897, p. 274
  311. ^ Emsley 2001, p. 428
  312. ^ a b Kudryavtsev 1974, p. 78
  313. ^ Bagnall 1966, pp. 32–3, 59, 137
  314. ^ Swink et al. 1966; Anderson et al. 1980
  315. ^ Ahmed, Fjellvåg & Kjekshus 2000
  316. ^ Chizhikov & Shchastlivyi 1970, p. 28
  317. ^ Kudryavtsev 1974, p. 77
  318. ^ Stuke 1974, p. 178; Donohue 1982, pp. 386–7; Cotton et al. 1999, p. 501
  319. ^ Becker, Johnson & Nussbaum 1971, p. 56
  320. ^ a b Berger 1997, p. 90
  321. ^ Chizhikov & Shchastlivyi 1970, p. 16
  322. ^ Jolly 1966, pp. 66–7
  323. ^ Mellor 1964, p.  30; Wiberg 2001, p. 589
  324. ^ Greenwood & Earnshaw 2002, p. 765–6
  325. ^ Bagnall 1966, p. 134–51; Greenwood & Earnshaw 2002, p. 786
  326. ^ Detty & O'Regan 1994, pp. 1–2
  327. ^ Hill & Holman 2000, p. 124
  328. ^ Chang 2002, p. 314
  329. ^ Kent 1950, pp. 1–2; Clark 1960, p. 588; Warren & Geballe 1981
  330. ^ Housecroft & Sharpe 2008, p. 384; IUPAC 2006–, rhombohedral graphite entry
  331. ^ Mingos 1998, p. 171
  332. ^ Wiberg 2001, p. 781
  333. ^ a b c Atkins et al. 2006, pp. 320–1
  334. ^ Savvatimskiy 2005, p. 1138
  335. ^ Togaya 2000
  336. ^ Savvatimskiy 2009
  337. ^ Inagaki 2000, p. 216; Yasuda et al. 2003, pp. 3–11
  338. ^ a b Wiberg 2001, p. 795
  339. ^ Traynham 1989, pp. 930–1; Prakash & Schleyer 1997
  340. ^ Olmsted & Williams 1997, p. 436
  341. ^ Bailar et al. 1989, p. 743
  342. ^ Moore et al. 1985
  343. ^ House & House 2010, p. 526
  344. ^ Wiberg 2001, p. 798
  345. ^ Eagleson 1994, p. 175
  346. ^ Atkins et al. 2006, p. 121
  347. ^ Russell & Lee 2005, pp. 358–9
  348. ^ Russell & Lee 2005, pp. 358–60 et seq
  349. ^ a b Metcalfe et al. 1974, p. 539
  350. ^ Cobb & Fetterolf 2005, p. 64; Metcalfe, Williams & Castka 1974, p. 539
  351. ^ Ogata, Li & Yip 2002; Boyer et al. 2004, p. 1023; Russell & Lee 2005, p. 359
  352. ^ Cooper 1968, p. 25; Henderson 2000, p. 5; Silberberg 2006, p. 314
  353. ^ Wiberg 2001, p. 1014
  354. ^ Stott 1956, p. 100
  355. ^ Steele 1966, p. 60
  356. ^ 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.'
  357. ^ Holt, Rinehart & Wilson c. 2007
  358. ^ Tuthill 2011
  359. ^ Emsley 2001, p. 382
  360. ^ Young et al. 2010, p. 9; Craig 2003, p. 391. Selenium is 'near metalloidal'.
  361. ^ Rochow 1957
  362. ^ Rochow 1966
  363. ^ Moss 1952, p. 192
  364. ^ a b Glinka 1965, p. 356
  365. ^ Evans 1966, pp. 124–5
  366. ^ Regnault 1853, p. 208
  367. ^ Scott & Kanda 1962, p. 311
  368. ^ Cotton et al. 1999, pp. 496, 503–4
  369. ^ Arlman 1939; Bagnall 1966, pp. 135, 142–3
  370. ^ Chao & Stenger 1964
  371. ^ a b Berger 1997, pp. 86–7
  372. ^ Snyder 1966, p. 242
  373. ^ Fritz & Gjerde 2008, p. 235
  374. ^ Meyer et al. 2005, p. 284; Manahan 2001, p. 911; Szpunar et al. 2004, p. 17
  375. ^ US Environmental Protection Agency 1988, p. 1; Uden 2005, pp. 347‒8
  376. ^ De Zuane 1997, p. 93; Dev 2008, pp. 2‒3
  377. ^ a b Cotton et al. 1999, p. 502
  378. ^ Wiberg 2001, p. 594
  379. ^ Greenwood & Earnshaw 2002, p. 786; Schwietzer & Pesterfield 2010, pp. 242–3
  380. ^ Bagnall 1966, p. 41; Nickless 1968, p. 79
  381. ^ 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]...'
  382. ^ Bagnall 1957, p. 62; Fernelius 1982, p. 741
  383. ^ Bagnall 1966, p. 41; Barrett 2003, p. 119
  384. ^ Hawkes 2010; Holt, Rinehart & Wilson c. 2007; Hawkes 1999, p. 14; Roza 2009, p. 12
  385. ^ Keller 1985
  386. ^ Harding, Johnson & Janes 2002, p. 61
  387. ^ Long & Hentz 1986, p. 58
  388. ^ Vasáros & Berei 1985, p. 109
  389. ^ Haissinsky & Coche 1949, p. 400
  390. ^ Brownlee et al. 1950, p. 173
  391. ^ Hermann, Hoffmann & Ashcroft 2013
  392. ^ Siekierski & Burgess 2002, pp. 65, 122
  393. ^ Emsley 2001, p. 48
  394. ^ Rao & Ganguly 1986
  395. ^ Krishnan et al. 1998
  396. ^ Glorieux, Saboungi & Enderby 2001
  397. ^ Millot et al. 2002
  398. ^ Vasáros & Berei 1985, p. 117
  399. ^ Kaye & Laby 1973, p. 228
  400. ^ Samsonov 1968, p. 590
  401. ^ Korenman 1959, p. 1368
  402. ^ Rossler 1985, pp. 143–4
  403. ^ Champion et al. 2010
  404. ^ Borst 1982, pp. 465, 473
  405. ^ Batsanov 1971, p. 811
  406. ^ Swalin 1962, p. 216; Feng & Lin 2005, p. 157
  407. ^ Schwietzer & Pesterfield 2010, pp. 258–60
  408. ^ Hawkes 1999, p. 14
  409. ^ Olmsted & Williams 1997, p. 328; Daintith 2004, p. 277
  410. ^ Eberle1985, pp. 213–16, 222–7
  411. ^ Restrepo et al. 2004, p. 69; Restrepo et al. 2006, p. 411
  412. ^ Greenwood & Earnshaw 2002, p. 804
  413. ^ Greenwood & Earnshaw 2002, p. 803
  414. ^ Wiberg 2001, p. 416
  415. ^ Craig 2003, p. 391; Schroers 2013, p. 32; Vernon 2013, pp. 1704–1705
  416. ^ Cotton et al. 1999, p. 42
  417. ^ Marezio & Licci 2000, p. 11
  418. ^ a b Vernon 2013, p. 1705
  419. ^ Russell & Lee 2005, p. 5
  420. ^ Parish 1977, pp. 178, 192–3
  421. ^ Eggins 1972, p. 66; Rayner-Canham & Overton 2006, pp. 29–30
  422. ^ Atkins et al. 2006, pp. 320–1; Bailar et al. 1989, p. 742–3
  423. ^ 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, pp. 1–38: '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).
  424. ^ Stuke 1974, p. 178; Cotton et al. 1999, p. 501; Craig 2003, p. 391
  425. ^ 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...'.
  426. ^ a b Lutz 2011, p. 16
  427. ^ Yacobi & Holt 1990, p. 10; Wiberg 2001, p. 160
  428. ^ Eagleson 1994, p. 820
  429. ^ Oxtoby, Gillis & Campion 2008, p. 508
  430. ^ Greenwood & Earnshaw 2002, pp. 479, 482
  431. ^ Brescia et al. 1980, pp. 166–71
  432. ^ Fine & Beall 1990, p. 578
  433. ^ Wiberg 2001, p. 901
  434. ^ Berger 1997, p. 80
  435. ^ Lovett 1977, p. 101
  436. ^ Cohen & Chelikowsky 1988, p. 99
  437. ^ Taguena-Martinez, Barrio & Chambouleyron 1991, p. 141
  438. ^ Ebbing & Gammon 2010, p. 891
  439. ^ Asmussen & Reinhard 2002, p. 7
  440. ^ Deprez & McLachan 1988
  441. ^ Addison 1964 (P, Se, Sn); Marković, Christiansen & Goldman 1998 (Bi); Nagao et al. 2004

References

  • Addison WE 1964, The Allotropy of the Elements, Oldbourne Press, London
  • Addison CC & Sowerby DB 1972, Main Group Elements: Groups V and VI, Butterworths, London, ISBN 0839110057
  • Adler D 1969, 'Half-way Elements: The Technology of Metalloids', book review, Technology Review, vol. 72, no. 1, Oct/Nov, pp. 18–19, ISSN 00401692 Parameter error in {{issn}}: Invalid ISSN.
  • Ahmed MAK, Fjellvåg H & Kjekshus A 2000, 'Synthesis, Structure and Thermal Stability of Tellurium Oxides and Oxide Sulfate Formed from Reactions in Refluxing Sulfuric Acid', Journal of the Chemical Society, Dalton Transactions, no. 24, pp. 4542–9, doi:10.1039/B005688J
  • Aldridge BG 1998, Science Interactions: Course 2, 3rd ed., Glencoe/McGraw-Hill, Westerville, Ohio, ISBN 9780028281575
  • Allen DS & Ordway RJ 1968, Physical Science, 2nd ed., Van Nostrand, Princeton, New Jersey, ISBN 9780442002909
  • Allen PB & Broughton JQ 1987, 'Electrical Conductivity and Electronic Properties of Liquid Silicon', Journal of Physical Chemistry, vol. 91, no. 19, pp. 4964–70, doi:10.1021/j100303a015
  • Alloul H 2010, Introduction to the Physics of Electrons in Solids, Springer-Verlag, Berlin, ISBN 3642135641
  • Anderson JB, Rapposch MH, Anderson CP & Kostiner E 1980, 'Crystal Structure Refinement of Basic Tellurium Nitrate: A Reformulation as (Te2O4H)+(NO3)', Monatshefte für Chemie/ Chemical Monthly, vol. 111, no. 4, pp. 789–96, doi:10.1007/BF00899243
  • Antman KH 2001, 'Introduction: The History of Arsenic Trioxide in Cancer Therapy', The Oncologist, vol. 6, suppl. 2, pp. 1–2, doi:10.1634/theoncologist.6-suppl_2-1
  • Apseloff G 1999, 'Therapeutic Uses of Gallium Nitrate: Past, Present, and Future', American Journal of Therapeutics, vol. 6, no. 6, pp. 327–39, ISSN 15363686 Parameter error in {{issn}}: Invalid ISSN.
  • Arlman EJ 1939, 'The Complex Compounds P(OH)4.ClO4 and Se(OH)3.ClO4', Recueil des Travaux Chimiques des Pays-Bas, vol. 58, no. 10, pp. 871–4, ISSN 01650513 Parameter error in {{issn}}: Invalid ISSN.
  • Askeland DR, Fulay PP & Wright JW 2011, The Science and Engineering of Materials, 6th ed., Cengage Learning, Stamford, CT, ISBN 0495668022
  • Asmussen J & Reinhard DK 2002, Diamond Films Handbook, Marcel Dekker, New York, ISBN 0824795776
  • Atkins P, Overton T, Rourke J, Weller M & Armstrong F 2006, Shriver & Atkins' Inorganic Chemistry, 4th ed., Oxford University Press, Oxford, ISBN 0716748789
  • Atkins P, Overton T, Rourke J, Weller M & Armstrong F 2010, Shriver & Atkins' Inorganic Chemistry, 5th ed., Oxford University Press, Oxford, ISBN 1429218207
  • Austen K 2012, 'A Factory for Elements that Barely Exist', New Scientist, 21 Apr, p. 12
  • Ba LA, Döring M, Jamier V & Jacob C 2010, 'Tellurium: an Element with Great Biological Potency and Potential', Organic & Biomolecular Chemistry, vol. 8, pp. 4203–16, doi:10.1039/C0OB00086H
  • Bagnall KW 1957, Chemistry of the Rare Radioelements: Polonium-actinium, Butterworths Scientific Publications, London
  • Bagnall KW 1966, The Chemistry of Selenium, Tellurium and Polonium, Elsevier, Amsterdam
  • Bagnall KC 1990, 'Compounds of Polonium', in KC Buschbeck & C Keller (eds), Gmelin Handbook of Inorganic and Organometallic Chemistry, 8th ed., Po Polonium, Supplement vol. 1, Springer-Verlag, Berlin, pp. 285–340, ISBN 3540936165
  • Bailar JC, Moeller T & Kleinberg J 1965, University Chemistry, DC Heath, Boston
  • Bailar JC & Trotman-Dickenson AF 1973, Comprehensive Inorganic Chemistry, vol. 4, Pergamon, Oxford
  • Bailar JC, Moeller T, Kleinberg J, Guss CO, Castellion ME & Metz C 1989, Chemistry, 3rd ed., Harcourt Brace Jovanovich, San Diego, ISBN 0155064568
  • Barfuß H, Böhnlein G, Freunek P, Hofmann R, Hohenstein H, Kreische W, Niedrig H and Reimer A 1981, 'The Electric Quadrupole Interaction of 111Cd in Arsenic Metal and in the System Sb1–xInx and Sb1–xCdx', Hyperfine Interactions, vol. 10, nos 1–4, pp. 967–72, doi:10.1007/BF01022038
  • Barrett J 2003, Inorganic Chemistry in Aqueous Solution, The Royal Society of Chemistry, Cambridge, ISBN 085404471X
  • Barsanov GP & Ginzburg AI 1974, 'Mineral', in AM Prokhorov (ed.), Great Soviet Encyclopedia, 3rd ed., vol. 16, Macmillan, New York, pp. 329–32
  • Bassett LG, Bunce SC, Carter AE, Clark HM & Hollinger HB 1966, Principles of Chemistry, Prentice-Hall, Englewood Cliffs, New Jersey
  • Batsanov SS 1971, 'Quantitative Characteristics of Bond Metallicity in Crystals', Journal of Structural Chemistry, vol. 12, no. 5, pp. 809–13, doi:10.1007/BF00743349
  • Becker WM, Johnson VA & Nussbaum 1971, 'The Physical Properties of Tellurium', in WC Cooper (ed.), Tellurium, Van Nostrand Reinhold, New York
  • Belpassi L, Tarantelli F, Sgamellotti A & Quiney HM 2006, 'The Electronic Structure of Alkali Aurides. A Four-Component Dirac−Kohn−Sham study', The Journal of Physical Chemistry A, vol. 110, no. 13, April 6, pp. 4543–54, doi:10.1021/jp054938w
  • Berger LI 1997, Semiconductor Materials, CRC Press, Boca Raton, Florida, ISBN 0849389127
  • Bianco E, Butler S, Jiang S, Restrepo OD, Windl W & Goldberger JE 2013, 'Stability and Exfoliation of Germanane: A Germanium Graphane Analogue,' ACS Nano, March 19 (web), doi:10.1021/nn4009406
  • Bodner GM & Pardue HL 1993, Chemistry, An Experimental Science, John Wiley & Sons, New York, ISBN 0471593869
  • Bogoroditskii NP & Pasynkov VV 1967, Radio and Electronic Materials, Iliffe Books, London
  • Boikess RS & Edelson E 1985, Chemical Principles, 3rd ed., Harper & Row, New York, ISBN 0060408057
  • Bomgardner MM 2013, 'Odds Narrow for Thin-Film Solar', Chemical & Engineering News, vol. 91, no. 20, pp. 20–1, ISSN 00092347 Parameter error in {{issn}}: Invalid ISSN.
  • Bond GC 2005, Metal-Catalysed Reactions of Hydrocarbons, Springer, New York, ISBN 0387241418
  • Booth VH & Bloom ML 1972, Physical Science: A Study of Matter and Energy, Macmillan, New York
  • Born M & Wolf E 1999, Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light, 7th ed., Cambridge University Press, Cambridge, ISBN 0521642221
  • Borst KE 1982, 'Characteristic Properties of Metallic Crystals', Journal of Educational Modules for Materials Science and Engineering, vol. 4, no. 3, pp. 457–92, ISSN 01973940 Parameter error in {{issn}}: Invalid ISSN.
  • Bowser JR 1993, Inorganic Chemistry, Brooks/Cole, Pacific Grove, California, ISBN 0534175325
  • Boyer RD, Li J, Ogata S & Yip S 2004, 'Analysis of Shear Deformations in Al and Cu: Empirical Potentials Versus Density Functional Theory', Modelling and Simulation in Materials Science and Engineering, vol. 12, no. 5, pp. 1017–29, doi:10.1088/0965-0393/12/5/017
  • Boylan PJ 1962, Elements of Chemistry, Allyn & Bacon, Boston
  • Bradbury GM, McGill MV, Smith HR & Baker PS 1957, Chemistry and You, Lyons and Carnahan, Chicago
  • Brady JE, Senese F & Jespersen ND 2009, Chemistry: Matter and its Changes, John Wiley & Sons, Hoboken, New Jersey, ISBN 9780470120941
  • Bresica F, Arents J, Meislich H & Turk A 1980, Fundamentals of Chemistry, 4th ed., Academic Press, New York, ISBN 0121323927
  • Brown L & Holme T 2006, Chemistry for Engineering Students, Thomson Brooks/Cole, Belmont California, ISBN 0495017183
  • Brown TL, LeMay HE, Bursten BE, Murphy CJ, Woodward P 2009, Chemistry: The Central Science, 11th ed., Pearson Education, Upper Saddle River, New Jersey, ISBN 9780132358484
  • Brown WP c. 2007 'The Properties of Semi-Metals or Metalloids,' Doc Brown's Chemistry: Introduction to the Periodic Table, viewed 8 February 2013
  • Brownlee RB, Fuller RW, Hancock WJ, Sohon MD & Whitsit JE 1943, Elements of Chemistry, Allyn and Bacon, Boston
  • Brownlee RB, Fuller RT, Whitsit JE Hancock WJ & Sohon MD 1950, Elements of Chemistry, Allyn and Bacon, Boston
  • Bucat RB (ed.) 1983, Elements of Chemistry: Earth, Air, Fire & Water, vol. 1, Australian Academy of Science, Canberra, ISBN 0858471132
  • Büchel KH (ed.) 1983, Chemistry of Pesticides, John Wiley & Sons, New York, ISBN 0471056820
  • Büchel KH, Moretto H-H, Woditsch P 2003, Industrial Inorganic Chemistry, 2nd ed., Wiley-VCH, ISBN 3527298495
  • Burakowski T & Wierzchoń T 1999, Surface Engineering of Metals: Principles, Equipment, Technologies, CRC Press, Boca Raton, Florida, ISBN 0849382254
  • Burrows A, Holman J, Parsons A, Pilling G & Price G 2009, Chemistry3: Introducing Inorganic, Organic and Physical Chemistry, Oxford University, Oxford, ISBN 0199277893
  • Butterman WC & Carlin JF 2004, Mineral Commodity Profiles: Antimony, US Geological Survey
  • Butterman WC & Jorgenson JD 2005, Mineral Commodity Profiles: Germanium, US Geological Survey
  • Buzea C & Robbie K 2005, 'Assembling the Puzzle of Superconducting Elements: A Review', Superconductor Science and Technology, vol 18, no. 1, pp. R1–R8, doi:10.1088/0953-2048/18/1/R01
  • Calderazzo F, Ercoli R & Natta G 1968, 'Metal Carbonyls: Preparation, Structure, and Properties', in I Wender & P Pino (eds), Organic Syntheses via Metal Carbonyls: Volume 1, Interscience Publishers, New York, pp. 1–272
  • Carapella SC 1968a, 'Arsenic' in CA Hampel (ed.), The Encyclopedia of the Chemical Elements, Reinhold, New York, pp. 29–32
  • Carapella SC 1968, 'Antimony' in CA Hampel (ed.), The Encyclopedia of the Chemical Elements, Reinhold, New York, pp. 22–5
  • Carlin JF 2011, Minerals Year Book: Antimony, United States Geological Survey
  • Carter CB & Norton MG 2013, Ceramic Materials: Science and Engineering, 2nd ed., Springer Science+Business Media, New York, ISBN 9781461435235
  • Cegielski C 1998, Yearbook of Science and the Future, Encyclopædia Britannica, Chicago, ISBN 0852296576
  • Chalmers B 1959, Physical Metallurgy, John Wiley & Sons, New York
  • Champion J, Alliot C, Renault E, Mokili BM, Chérel M, Galland N & Montavon G 2010, 'Astatine Standard Redox Potentials and Speciation in Acidic Medium', The Journal of Physical Chemistry A, vol. 114, no. 1, pp. 576–82, doi:10.1021/jp9077008
  • Chang R 2002, Chemistry, 7th ed., McGraw Hill, Boston, ISBN 0072465336
  • Chao MS & Stenger VA 1964, 'Some Physical Properties of Highly Purified Bromine', Talanta, vol. 11, no. 2, pp. 271–81, doi:10.1016/0039-9140(64)80036-9
  • Chatt J 1951, 'Metal and Metalloid Compounds of the Alkyl Radicals', in EH Rodd (ed.), Chemistry of Carbon Compounds: A Modern Comprehensive Treatise, vol. 1, part A, Elsevier, Amsterdam, pp. 417–58
  • Chedd G 1969, Half-Way Elements: The Technology of Metalloids, Doubleday, New York
  • Chizhikov DM & Shchastlivyi VP 1968, Selenium and Selenides, translated from the Russian by EM Elkin, Collet’s, London
  • Chizhikov DM & Shchastlivyi 1970, Tellurium and the Tellurides, Collet's, London
  • Choppin GR & Johnsen RH 1972, Introductory Chemistry, Addison-Wesley, Reading, Massachusetts
  • Chung DDL 2010, Composite Materials: Science and Applications, 2nd ed., Springer-Verlag, London, ISBN 978184882830
  • Clark GL 1960, The Encyclopedia of Chemistry, Reinhold, New York
  • Cobb C & Fetterolf ML 2005, The Joy of Chemistry, Prometheus Books, New York, ISBN 1591022312
  • Cohen ML & Chelikowsky JR 1988, Electronic Structure and Optical Properties of Semiconductors, Springer Verlag, Berlin, ISBN 3540188185
  • Coles BR & Caplin AD 1976, The Electronic Structures of Solids, Edward Arnold, London, ISBN 0844808741
  • Considine DM & Considine GD (eds) 1984, 'Metalloid', in Van Nostrand Reinhold Encyclopedia of Chemistry, 4th ed., Van Nostrand Reinhold, New York, ISBN 0442225725
  • Cooper DG 1968, The Periodic Table, 4th ed., Butterworths, London
  • Corbridge DEC 2012, Phosphorus: Chemistry, Biochemistry and Technology, 6th ed., CRC Press, Boca Raton, Florida, ISBN 9781439840887
  • Corwin CH 2005, Introductory Chemistry: Concepts & Connections, 4th ed., Prentice Hall, Upper Saddle River, New Jersey, ISBN 0131448501
  • Cotton FA, Wilkinson G & Gaus P 1995, Basic Inorganic Chemistry, 3rd ed., John Wiley & Sons, New York, ISBN 0471505323
  • Cotton FA, Wilkinson G, Murillo CA & Bochmann 1999, Advanced Inorganic Chemistry, 6th ed., John Wiley & Sons, New York, ISBN 0471199575
  • Cox PA 2004, Inorganic Chemistry, 2nd ed., Instant Notes series, Bios Scientific, London, ISBN 1859962890
  • Craig PJ 2003, Organometallic Compounds in the Environment, John Wiley & Sons, New York, ISBN 0471899933
  • Cusack N 1967, The Electrical and Magnetic Properties of Solids: An Introductory Textbook, 5th ed., John Wiley & Sons, New York
  • Cusack N E 1987, The Physics of Structurally Disordered Matter: An Introduction, A Hilger in association with the University of Sussex Press, Bristol, ISBN 0852745915
  • Daintith J (ed.) 2004, Oxford Dictionary of Chemistry, 5th ed., Oxford University, Oxford, ISBN 0199204632
  • Daniel-Hoffmann M, Sredni B & Nitzan Y 2012, 'Bactericidal Activity of the Organo-Tellurium Compound AS101 Against Enterobacter Cloacae,' Journal of Antimicrobial Chemotherapy, vol. 67, no. 9, pp. 2165–72, doi:10.1093/jac/dks185
  • Daub GW & Seese WS 1996, Basic Chemistry, 7th ed., Prentice Hall, New York, ISBN 013373630X
  • Davidson DF & Lakin HW 1973, 'Tellurium', in DA Brobst & WP Pratt (eds), United States Mineral Resources, Geological survey professional paper 820, United States Government Printing Office, Washington, pp. 627–30
  • Dávila ME, Molotov SL, Laubschat C & Asensio MC 2002, 'Structural Determination of Yb Single-Crystal Films Grown on W(110) Using Photoelectron Diffraction', Physical Review B, vol. 66, no. 3, p. 035411–18, doi:10.1103/PhysRevB.66.035411
  • Demetriou MD, Launey ME, Garrett G, Schramm JP, Hofmann DC, Johnson WL & Ritchie RO 2011, 'A Damage-Tolerant Glass', Nature Materials, vol. 10, February, pp. 123–8, doi:10.1038/nmat2930
  • Deming HG 1925, General Chemistry: An Elementary Survey, 2nd ed., John Wiley & Sons, New York
  • Denniston KJ, Topping JJ & Caret RL 2004, General, Organic, and Biochemistry, 5th ed., McGraw-Hill, New York, ISBN 0072828471
  • Deprez N & McLachan DS 1988, 'The Analysis of the Electrical Conductivity of Graphite Conductivity of Graphite Powders During Compaction', Journal of Physics D: Applied Physics, vol. 21, no. 1, doi:10.1088/0022-3727/21/1/015
  • Desai PD, James HM & Ho CY 1984, 'Electrical Resistivity of Aluminum and Manganese', Journal of Physical and Chemical Reference Data, vol. 13, no. 4, pp. 1131–72, doi:10.1063/1.555725
  • Desch CH 1914, Intermetallic Compounds, Longmans, Green and Co., New York
  • Detty MR & O’Regan MB 1994, Tellurium-Containing Heterocycles, (The Chemistry of Heterocyclic Compounds, vol. 53), John Wiley & Sons, New York
  • Dev N 2008, 'Modelling Selenium Fate and Transport in Great Salt Lake Wetlands', PhD dissertation, University of Utah, ProQuest, Ann Arbor, Michigan, ISBN 054986542X
  • De Zuane J 1997, Handbook of Drinking Water Quality, 2nd ed., John Wiley & Sons, New York, ISBN 047128789X
  • Divakar C, Mohan M & Singh AK 1984, 'The Kinetics of Pressure-Induced Fcc-Bcc Transformation in Ytterbium', Journal of Applied Physics, vol. 56, no. 8, pp. 2337–40, doi:10.1063/1.334270
  • Donohue J 1982, The Structures of the Elements, Robert E. Krieger, Malabar, Florida, ISBN 0898742307
  • Douglade J & Mercier R 1982, 'Structure Cristalline et Covalence des Liaisons dans le Sulfate d'Arsenic(III), As2(SO4)3', Acta Crystallographica Section B, vol. 38, no. 3, pp. 720–3, doi:10.1107/S056774088200394X
  • Du Y, Ouyang C, Shi S & Lei M 2010, 'Ab Initio Studies on Atomic and Electronic Structures of Black Phosphorus', Journal of Applied Physics, vol. 107, no. 9, pp. 093718–1–4, doi:10.1063/1.3386509
  • Dunlap BD, Brodsky MB, Shenoy GK & Kalvius GM 1970, 'Hyperfine Interactions and Anisotropic Lattice Vibrations of 237Np in α-Np Metal', Physical Review B, vol. 1, no. 1, pp. 44–9, doi:10.1103/PhysRevB.1.44
  • Dunstan S 1968, Principles of Chemistry, D. Van Nostrand Company, London
  • Dupree R, Kirby DJ & Freyland W 1982, 'N.M.R. Study of Changes in Bonding and the Metal-Non-metal Transition in Liquid Caesium-Antimony Alloys', Philosophical Magazine Part B, vol. 46 no. 6, pp. 595–606, doi:10.1080/01418638208223546
  • Eagleson M 1994, Concise Encyclopedia Chemistry, Walter de Gruyter, Berlin, ISBN 3110114518
  • Eason R 2007, Pulsed Laser Deposition of Thin Films: Applications-Led Growth of Functional Materials, Wiley-Interscience, New York
  • Ebbing DD & Gammon SD 2010, General Chemistry, 9th ed. enhanced, Brooks/Cole, Belmont, California, ISBN 9780618934690
  • Eberle SH 1985, 'Chemical Behavior and Compounds of Astatine', pp. 183–209, in Kugler & Keller
  • Edwards PP & Sienko MJ 1983, 'On the Occurrence of Metallic Character in the Periodic Table of the Elements', Journal of Chemical Education, vol. 60, no. 9, pp. 691–6, doi:10.1021ed060p691
  • Edwards PP, Lodge MTJ, Hensel F & Redmer R 2010, '...A Metal Conducts and a Non-metal Doesn't', Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, vol. 368, pp. 941–65, doi:10.1098/rsta.2009.0282
  • Eggins BR 1972, Chemical Structure and Reactivity, MacMillan, London, ISBN 0333081455
  • Eichler R, Aksenov NV, Belozerov AV, Bozhikov GA, Chepigin VI, Dmitriev SN, Dressler R, Gäggeler HW, Gorshkov VA, Haenssler F, Itkis MG, Laube A, Lebedev VY, Malyshev ON, Oganessian YT, Petrushkin OV, Piguet D, Rasmussen P, Shishkin SV, Shutov, AV, Svirikhin AI, Tereshatov EE, Vostokin GK, Wegrzecki M & Yeremin AV 2007, 'Chemical Characterization of Element 112,' Nature, vol. 447, pp. 72–5, doi:10.1038/nature05761
  • Emeleús HJ & Sharpe AG 1959, Advances in Inorganic Chemistry and Radiochemistry, vol. 1, Academic Press, New York
  • Emsley J 1971, The Inorganic Chemistry of the Non-metals, Methuen Educational, London, ISBN 0423861204
  • Emsley J 2001, Nature's Building Blocks: An A–Z guide to the Elements, Oxford University Press, Oxford, ISBN 0198503415
  • Eranna G 2011, Metal Oxide Nanostructures as Gas Sensing Devices, Taylor & Francis, Boca Raton, Florida, ISBN 1439863407
  • Evans RC 1966, An Introduction to Crystal Chemistry, Cambridge University, Cambridge
  • Evans KA 1993, 'Properties and Uses of Oxides and Hydroxides,' in AJ Downs (ed.), Chemistry of Aluminium, Gallium, Indium, and Thallium, Blackie Academic & Professional, Bishopbriggs, Glasgow, pp. 248–91, ISBN 075140103X
  • EVM (Expert Group on Vitamins and Minerals) 2003, Safe Upper Levels for Vitamins and Minerals, UK Food Standards Agency, London, ISBN 1904026117
  • Fehlner TP 1990, 'The Metallic Face of Boron,' in AG Sykes (ed.), Advances in Inorganic Chemistry, vol. 35, Academic Press, Orlando, pp. 199–233
  • Fellet M 'Redefining Hydrogen Bonds, the Givers of Life,' New Scientist, 28 July 2011, ISSN 10321233 Parameter error in {{issn}}: Invalid ISSN.
  • Feng & Jin 2005, Introduction to Condensed Matter Physics: Volume 1, World Scientific, Singapore, ISBN 1842653474
  • Fernelius WC 1982, 'Polonium', Journal of Chemical Education, vol. 59, no. 9, pp. 741–2, doi:10.1021/ed059p741
  • Ferro R & Saccone A 2008, Intermetallic Chemistry, Elsevier, Oxford, p. 233, ISBN 0080440991
  • Fesquet AA 1872, A Practical Guide for the Manufacture of Metallic Alloys, trans. A. Guettier, Henry Carey Baird, Philadelphia
  • Fine LW & Beall H 1990, Chemistry for Engineers and Scientists, Saunders College Publishing, Philadelphia, ISBN 0030215374
  • Foster W 1936, The Romance of Chemistry, D Appleton-Century, New York
  • Foster LS & Wrigley AN 1958, 'Periodic Table', in GL Clark, GG Hawley & WA Hamor (eds), The Encyclopedia of Chemistry (Supplement), Reinhold, New York, pp. 215–20
  • Friend JN 1953, Man and the Chemical Elements, 1st ed., Charles Scribner's Sons, New York
  • Fritz JS & Gjerde DT 2008, Ion Chromatography, John Wiley & Sons, New York, ISBN 3527613250
  • Gary S 2013, 'Poisoned Alloy' the Metal of the Future', News in science, viewed 28 August 2013
  • Geckeler S 1987, Optical Fiber Transmission Systems, Artech Hous, Norwood, Massachusetts, ISBN 0890062269
  • Geman Energy Society 2008, Planning and Installing Photovoltaic Systems: A Guide for Installers, Architects and Engineers, 2nd ed., Earthscan, London, ISBN 9781844074426
  • Gordh G, Gordh G & Headrick D 2003, A Dictionary of Entomology, CABI Publishing, Wallingford, ISBN 0851996558
  • Gladyshev VP & Kovaleva SV 1998, 'Liquidus Shape of the Mercury–Gallium System', Russian Journal of Inorganic Chemistry, vol. 43, no. 9, pp. 1445–6
  • Glazov VM, Chizhevskaya SN & Glagoleva NN 1969, Liquid Semiconductors, Plenum, New York
  • Glinka N 1965, General Chemistry, trans. D Sobolev, Gordon & Breach, New York
  • Glockling F 1969, The Chemistry of Germanium, Academic, London
  • Glorieux B, Saboungi ML & Enderby JE 2001, 'Electronic Conduction in Liquid Boron', Europhysics Letters (EPL), vol. 56, no. 1, pp. 81–5, doi:10.1209/epl/i2001-00490-0
  • Goldsmith RH 1982, 'Metalloids', Journal of Chemical Education, vol. 59, no. 6, pp. 526–7, doi:10.1021/ed059p526
  • Good JM, Gregory O & Bosworth N 1813, 'Arsenicum', in Pantologia: A New Cyclopedia...of Essays, Treatises, and Systems...with a General Dictionary of Arts, Sciences, and Words..., Kearsely, London
  • Goodrich BG 1844, A Glance at the Physical Sciences, Bradbury, Soden & Co., Boston
  • Gray T 2009, The Elements: A Visual Exploration of Every Known Atom in the Universe, Black Dog & Leventhal, New York, ISBN 9781579128142
  • Gray T 2010, 'Metalloids (7)', viewed 8 February 2013
  • Gray T, Whitby M & Mann N 2011, Mohs Hardness of the Elements, viewed 12 Feb 2012
  • Greaves GN, Knights JC & Davis EA 1974, 'Electronic Properties of Amorphous Arsenic', in J Stuke & W Brenig (eds), Amorphous and Liquid Semiconductors: Proceedings, vol. 1, Taylor & Francis, London, pp. 369–74, ISBN 9780470834855
  • Greenwood NN 2001, 'Main Group Element Chemistry at the Millennium', Journal of the Chemical Society, Dalton Transactions, issue 14, pp. 2055–66, doi:10.1039/b103917m
  • Greenwood NN & Earnshaw A 2002, Chemistry of the Elements, 2nd ed., Butterworth-Heinemann, ISBN 0750633654
  • Grimes RN 2011, Carboranes, 2nd ed., Academic Press, London, ISBN 012374170X
  • Guan PF, Fujita T, Hirata A, Liu YH & Chen MW 2012, 'Structural Origins of the Excellent Glass-forming Ability of Pd40Ni40P20', Physical Review Letters, vol. 108, no. 17, pp. 175501–1–5, doi:10.1103/PhysRevLett.108.175501
  • Hager T 2006, The Demon under the Microscope, Three Rivers Press, New York, ISBN 9781400082148
  • Haiduc I & Zuckerman JJ 1985, Basic Organometallic Chemistry, Walter de Gruyter, Berlin, ISBN 0899250068
  • Haissinsky M & Coche A 1949, 'New Experiments on the Cathodic Deposition of Radio-elements', Journal of the Chemical Society, pp. S397–400
  • Haller EE 2006, 'Germanium: From its Discovery to SiGe Devices', Materials Science in Semiconductor Processing, vol. 9, nos 4–5, doi:10.1016/j.mssp.2006.08.063, viewed 8 February 2013
  • Hamm DI 1969, Fundamental Concepts of Chemistry, Meredith Corporation, New York, ISBN 0390406511
  • Hampel CA & Hawley GG 1966, The Encyclopedia of Chemistry, 3rd ed., Van Nostrand Reinhold, New York
  • Hampel CA (ed.) 1968, The Encyclopedia of the Chemical Elements, Reinhold, New York
  • Hampel CA & Hawley GG 1976, Glossary of Chemical Terms, Van Nostrand Reinhold, New York, ISBN 0442232381
  • Harding C, Johnson DA & Janes R 2002, Elements of the p Block, Royal Society of Chemistry, Cambridge, ISBN 0854046909
  • Hasan H 2009, The Boron Elements: Boron, Aluminum, Gallium, Indium, Thallium, The Rosen Publishing Group, New York, ISBN 1435853334
  • Hatcher WH 1949, An Introduction to Chemical Science, John Wiley & Sons, New York
  • Hawkes SJ 1999, 'Polonium and Astatine are not Semimetals', Chem 13 News, February, p. 14, ISSN 07031157 Parameter error in {{issn}}: Invalid ISSN.
  • Hawkes SJ 2001, 'Semimetallicity', Journal of Chemical Education, vol. 78, no. 12, pp. 1686–7, doi:10.1021/ed078p1686
  • Hawkes SJ 2010, 'Polonium and Astatine are not Semimetals', Journal of Chemical Education, vol. 87, no. 8, p. 783, doi:10.1021ed100308w
  • Haynes WM (ed.) 2012, CRC Handbook of Chemistry and Physics, 93rd ed., CRC Press, Boca Raton, Florida, ISBN 1439880492
  • Hein M, Best LR, Pattison S & Arena S 2004, 8th ed., Introduction to General, Organic, and Biochemistry, John Wiley & Sons, Hoboken, New Jersey, ISBN 0471451967
  • Henderson M 2000, Main Group Chemistry, The Royal Society of Chemistry, Cambridge, ISBN 0854046178
  • Hermann A, Hoffmann R & Ashcroft NW 2013, 'Condensed Astatine: Monatomic and Metallic', Physical Review Letters, vol. 111, pp. 11604–1−11604-5, doi:10.1103/PhysRevLett.111.116404
  • Hérold A 2006, 'An Arrangement of the Chemical Elements in Several Classes Inside the Periodic Table According to their Common Properties', Comptes Rendus Chimie, vol. 9, no. 1, pp. 148–53, doi:10.1016/j.crci.2005.10.002
  • Hill G & Holman J 2000, Chemistry in Context, 5th ed., Nelson Thornes, Cheltenham, ISBN 0174483074
  • Hiller LA & Herber RH 1960, Principles of Chemistry, McGraw-Hill, New York
  • Hindman JC 1968, 'Neptunium', in CA Hampel (ed.), The Encyclopedia of the Chemical Elements, Reinhold, New York, pp. 432–7
  • Hoddeson L 2007, 'In the Wake of Thomas Kuhn's Theory of Scientific Revolutions: The Perspective of an Historian of Science,' in S Vosniadou, A Baltas & X Vamvakoussi (eds), Reframing the Conceptual Change Approach in Learning and Instruction, Elsevier, Amsterdam, pp. 25–34, ISBN 9780080453552
  • Holt, Rinehart & Wilson c. 2007 'Why Polonium and Astatine are not Metalloids in HRW texts', viewed 8 February 2013
  • Hopkins BS & Bailar JC 1956, General Chemistry for Colleges, 5th ed., D. C. Heath, Boston
  • Horvath 1973, 'Critical Temperature of Elements and the Periodic System', Journal of Chemical Education, vol. 50, no. 5, pp. 335–6, doi:10.1021/ed050p335
  • Houghton RP 1979, Metal Complexes in Organic Chemistry, Cambridge University Press, Cambridge, ISBN 0521219922
  • House JE 2008, Inorganic Chemistry, Academic Press (Elsevier), Burlington, Massachusetts, ISBN 0123567866
  • House JE & House KA 2010, Descriptive Inorganic Chemistry, 2nd ed., Academic Press, Burlington, Massachusetts, ISBN 012088755X
  • Housecroft CE & Sharpe AG 2008, Inorganic Chemistry, 3rd ed., Pearson Education, Harlow, ISBN 9780131755536
  • Hultgren HH 1966, 'Metalloids', in GL Clark & GG Hawley (eds), The Encyclopedia of Inorganic Chemistry, 2nd ed., Reinhold Publishing, New York
  • Hunt A 2000, The Complete A-Z Chemistry Handbook, 2nd ed., Hodder & Stoughton, London, ISBN 0340772182
  • Iler RK 1979, The Chemistry of Silica: Solubility, Polymerization, Colloid and Surface Properties, and Biochemistry, John Wiley, New York, ISBN 9780471024040
  • Inagaki M 2000, New Carbons: Control of Structure and Functions, Elsevier, Oxford, ISBN 0080437133
  • Isbell C 1998, 'Copper', in F Habashi (ed.), Alloys: Preparation, Properties, Applications, Wiley-VCH, Weinheim, pp. 65–108, ISBN 3527295917
  • IUPAC 1959, Nomenclature of Inorganic Chemistry, 1st ed., Butterworths, London
  • IUPAC 1971, Nomenclature of Inorganic Chemistry, 2nd ed., Butterworths, London, ISBN 0408701684
  • IUPAC 2005, Nomenclature of Inorganic Chemistry (the "Red Book"), NG Connelly & T Damhus eds, RSC Publishing, Cambridge, ISBN 0854044388
  • IUPAC 2006–, Compendium of Chemical Terminology (the "Gold Book"), 2nd ed., by M Nic, J Jirat & B Kosata, with updates compiled by A Jenkins, ISBN 0967855098, doi:10.1351/goldbook
  • James M, Stokes R, Ng W & Moloney J 2000, Chemical Connections 2: VCE Chemistry Units 3 & 4, John Wiley & Sons, Milton, Queensland, ISBN 0701634383
  • Jaouen G & Gibaud S 2010, 'Arsenic-based Drugs: From Fowler's solution to Modern Anticancer Chemotherapy', Medicinal Organometallic Chemistry, vol. 32, pp. 1–20, doi:10.1007/978-3-642-13185-1_1
  • Jaskula BW 2013, Mineral Commodity Profiles: Gallium, US Geological Survey
  • Jenkins GM & Kawamura K 1976, Polymeric Carbons—Carbon Fibre, Glass and Char, Cambridge University Press, Cambridge, ISBN 0521206936
  • Jezequel G & Thomas J 1997, 'Experimental Band Structure of Semimetal Bismuth', Physical Review B, vol. 56, no. 11, pp. 6620–6, doi:10.1103/PhysRevB.56.6620
  • Johansen G & Mackintosh AR 1970, 'Electronic Structure and Phase Transitions in Ytterbium', Solid State Communications, vol. 8, no. 2, pp. 121–4
  • Johnston FJ 1992, 'Metalloid' in McGraw-Hill Encyclopedia of Science & Technology, 7th ed., vol. 11, McGraw-Hill, New York, ISBN 0079092063 (set)
  • Jolly WL 1966, The Chemistry of the Non-metals, Prentice-Hall, Englewood Cliffs, New Jersey
  • Jones BW 2010, Pluto: Sentinel of the Outer Solar System, Cambridge University, Cambridge, ISBN 9780521194365
  • Kaminow IP & Li T 2002 (eds), Optical Fiber Telecommunications, Volume IVA, Academic Press, San Diego, ISBN 0123951720
  • Karabulut M, Melnik E, Stefan R, Marasinghe GK, Ray CS, Kurkjian CR & Day DE 2001, 'Mechanical and Structural Properties of Phosphate Glasses', Journal of Non-Crystalline Solids, vol. 288, nos. 1–3, pp. 8–17, doi:10.1016/S0022-3093(01)00615-9
  • Kaye GWC & Laby TH 1973, Tables of Physical and Chemical Constants, 14th ed., Longman, London, ISBN 0582463262
  • Keall JHH, Martin NH & Tunbridge RE 1946, 'A Report of Three Cases of Accidental Poisoning by Sodium Tellurite', British Journal of Industrial Medicine, vol. 3, no. 3, pp. 175–6
  • Keller C 1985, 'Preface', in Kugler & Keller
  • Kelter P, Mosher M & Scott A 2009, Chemistry: the Practical Science, Houghton Mifflin, Boston, ISBN 0547053932
  • Kent W 1950, Kent's Mechanical Engineers' Handbook, 12th ed., vol. 1, John Wiley & Sons, New York
  • King EL 1979, Chemistry, Painter Hopkins, Sausalito, California, ISBN 0052507262
  • King RB 1994, 'Antimony: Inorganic Chemistry', in RB King (ed), Encyclopedia of Inorganic Chemistry, John Wiley, Chichester, pp. 170–5, ISBN 0471936200
  • King RB 2004, 'The Metallurgist's Periodic Table and the Zintl-Klemm Concept', in DH Rouvray & RB King (eds), The Periodic Table: Into the 21st Century, Research Studies Press, Baldock, Hertfordshire, pp. 191–206, ISBN 0863802923
  • Kitaĭgorodskiĭ AI 1961, Organic Chemical Crystallography, Consultants Bureau, New York
  • Kleinberg J, Argersinger WJ & Griswold E 1960, Inorganic Chemistry, DC Health, Boston
  • Klement W, Willens RH & Duwez P 1960, 'Non-Crystalline Structure in Solidified Gold–Silicon Alloys', Nature, vol. 187, pp. 869–70, doi|10.1038/187869b0
  • Klemm W 1950, 'Einige Probleme aus der Physik und der Chemie der Halbmetalle und der Metametalle', Angewandte Chemie, vol. 62, no. 6, pp. 133–42
  • Klug HP & Brasted RC 1958, Comprehensive Inorganic Chemistry: The Elements and Compounds of Group IV A, Van Nostrand, New York
  • Kneen WR, Rogers MJW & Simpson P 1972, Chemistry: Facts, Patterns, and Principles, Addison-Wesley, London, ISBN 0201037793
  • Kolobov AV & Tominaga J 2012, Chalcogenides: Metastability and Phase Change Phenomena, Springer-Verlag, Heidelberg, ISBN 9783642287053
  • Kopp JG, Lipták BG & Eren H 000, 'Magnetic Flowmeters', in BG Lipták (ed.), Instrument Engineers' Handbook, 4th ed., vol. 1, Process Measurement and Analysis, CRC Press, Boca Raton, Florida, pp. 208–224, ISBN 0849310830
  • Korenman IM 1959, 'Regularities in Properties of Thallium', Journal of General Chemistry of the USSR, English translation, Consultants Bureau, New York, vol. 29, no. 2, pp. 1366–90, ISSN 00221279 Parameter error in {{issn}}: Invalid ISSN.
  • Kotz JC, Treichel P & Weaver GC 2009, Chemistry and Chemical Reactivity, 7th ed., Brooks/Cole, Belmont, California, ISBN 1439041318
  • Kozyrev PT 1959, 'Deoxidized Selenium and the Dependence of its Electrical Conductivity on Pressure. II', Physics of the Solid State, translation of the journal Solid State Physics (Fizika tverdogo tela) of the Academy of Sciences of the USSR, vol. 1, pp. 102–10
  • Kraig RE, Roundy D & Cohen ML 2004, 'A Study of the Mechanical and Structural Properties of Polonium', Solid State Communications, vol. 129, issue 6, Feb, pp. 411–13, doi:10.1016/j.ssc.2003.08.001
  • Krannich LK & Watkins CL 2006, 'Arsenic: Organoarsenic chemistry,' Encyclopedia of inorganic chemistry, viewed 12 Feb 2012
  • Kreith F & Goswami DY (eds) 2005, The CRC Handbook of Mechanical Engineering, 2nd ed., Boca Raton, Florida, ISBN 0849308666
  • Krishnan S, Ansell S, Felten J, Volin K & Price D 1998, 'Structure of Liquid Boron', Physical Review Letters, vol. 81, no. 3, pp. 586–9, doi:10.1103/PhysRevLett.81.586
  • Kudryavtsev AA 1974, The Chemistry & Technology of Selenium and Tellurium, translated from the 2nd Russian edition and revised by EM Elkin, Collet's, London, ISBN 0569080096
  • Kugler HK & Keller C (eds) 1985, Gmelin Handbook of Inorganic and Organometallic chemistry, 8th ed., 'At, Astatine', system no. 8a, Springer-Verlag, Berlin, ISBN 3540935169
  • Lagrenaudie J 1953, 'Semiconductive Properties of Boron' (in French), Journal de Chimie Physique, vol. 50, nos. 11–12, Nov–Dec, pp. 629–33, ISSN 00217689 Parameter error in {{issn}}: Invalid ISSN.
  • Le Bras M, Wilkie CA & Bourbigot S (eds) 2005, Fire Retardancy of Polymers: New Applications of Mineral Fillers, Royal Society of Chemistry, Cambridge, ISBN 0854045821
  • Lehto Y & Hou X 2011, Chemistry and Analysis of Radionuclides: Laboratory Techniques and Methodology, Wiley-VCH, Weinheim, ISBN 9783527326587
  • Lewis RJ 1993, Hawley's Condensed Chemical Dictionary, 12th ed., Van Nostrand Reinhold, New York, ISBN 0442011318
  • Li XP 1990, 'Properties of Liquid Arsenic: A Theoretical Study', Physical Review B, vol. 41, no. 12, pp. 8392–406, doi:10.1103/PhysRevB.41.8392
  • Locke EG, Baechler RH, Beglinger E, Bruce HD, Drow JT, Johnson KG, Laughnan DG, Paul BH, Rietz RC, Saeman JF & Tarkow H 1956, 'Wood', in RE Kirk & DF Othmer (eds), Encyclopedia of Chemical Technology, vol. 15, The Interscience Encyclopedia, New York, pp. 72–102
  • Löffler JF, Kündig AA & Dalla Torre FH 2007, 'Rapid Solidification and Bulk Metallic Glasses—Processing and Properties,' in JR Groza, JF Shackelford, EJ Lavernia EJ & MT Powers (eds), Materials Processing Handbook, CRC Press, Boca Raton, Florida, pp. 17–1–44, ISBN 0849332168
  • Long GG & Hentz FC 1986, Problem Exercises for General Chemistry, 3rd ed., John Wiley & Sons, New York, ISBN 0471828408
  • Lovett DR 1977, Semimetals & Narrow-Bandgap Semi-conductors, Pion, London, ISBN 0850860601
  • Lutz J, Schlangenotto H, Scheuermann U, De Doncker R 2011, Semiconductor Power Devices: Physics, Characteristics, Reliability, Springer-Verlag, Berlin, ISBN 3642111246
  • MacKay KM, MacKay RA & Henderson W 2002, Introduction to Modern Inorganic Chemistry, 6th ed., Nelson Thornes, Cheltenham, ISBN 0748764208
  • Madelung O 2004, Semiconductors: Data Handbook, 3rd ed., Springer-Verlag, Berlin, ISBN 9783540404880
  • Maeder T 2013, 'Review of Bi2O3 Based Glasses for Electronics and Related Applications, International Materials Reviews, vol. 58, no. 1, pp. 3‒40, doi:10.1179/1743280412Y.0000000010
  • Mahan BH 1965, University Chemistry, Addison-Wesley, Reading, Massachusetts
  • Malerba F 1985, The Semiconductor Business: The Economics of Rapid Growth and Decline, Frances Pinter, London, ISBN 0861875427
  • Manahan SE 2001, Fundamentals of Environmental Chemistry, 2nd ed., CRC Press, Boca Raton, Florida, ISBN 156670491X
  • Mann JB, Meek TL & Allen LC 2000, 'Configuration Energies of the Main Group Elements', Journal of the American Chemical Society, vol. 122, no. 12, pp. 2780–3, doi:10.1021ja992866e
  • Marezio M & Licci F 2000, 'Strategies for Tailoring New Superconducting Systems', in X Obradors, F Sandiumenge & J Fontcuberta (eds), Applied Superconductivity 1999: Large scale applications, volume 1 of Applied Superconductivity 1999: Proceedings of EUCAS 1999, the Fourth European Conference on Applied Superconductivity, held in Sitges, Spain, 14–17 September 1999, Institute of Physics, Bristol, pp. 11–16, ISBN 0750307455
  • Marković N, Christiansen C & Goldman AM 1998, 'Thickness-Magnetic Field Phase Diagram at the Superconductor-Insulator Transition in 2D', Physical Review Letters, vol. 81, no. 23, pp. 5217–20, doi:10.1103/PhysRevLett.81.5217
  • Massey AG 2000, Main Group Chemistry, 2nd ed., John Wiley & Sons, Chichester, ISBN 0471490393
  • Masters GM & Ela W 2008, Introduction to Environmental Engineering and Science, 3rd ed., Prentice Hall, Upper Saddle River, New Jersey, ISBN 9780131481930
  • Masterton WL & Slowinski EJ 1977, Chemical Principles, 4th ed., W. B. Saunders, Philadelphia, ISBN 0721661734
  • Matula RA 1979, 'Electrical Resistivity of Copper, Gold, Palladium, and Silver,' Journal of Physical and Chemical Reference Data, vol. 8, no. 4, pp. 1147–298, doi:10.1063/1.555614
  • McMurray J & Fay RC 2009, General Chemistry: Atoms First, Prentice Hall, Upper Saddle River, New Jersey, ISBN 0321571630
  • McQuarrie DA & Rock PA 1987, General Chemistry, 3rd ed., WH Freeman, New York, ISBN 0716721694
  • Mellor JW 1964, A Comprehensive Treatise on Inorganic and Theoretical Chemistry, vol. 9, John Wiley, New York
  • Mellor JW 1964, A Comprehensive Treatise on Inorganic and Theoretical Chemistry, vol. 11, John Wiley, New York
  • Mendeléeff DI 1897, The Principles of Chemistry, vol. 2, 5th ed., trans. G Kamensky, AJ Greenaway (ed.), Longmans, Green & Co., London
  • Meskers CEM, Hagelüken C & Van Damme G 2009, 'Green Recycling of EEE: Special and Precious Metal EEE', in SM Howard, P Anyalebechi & L Zhang (eds), Proceedings of Sessions and Symposia Sponsored by the Extraction and Processing Division (EPD) of The Minerals, Metals and Materials Society (TMS), held during the TMS 2009 Annual Meeting & Exhibition San Francisco, California, February 15–19, 2009, The Minerals, Metals and Materials Society, Warrendale, Pennsylvania, ISBN 9780873397322, pp. 1131–6
  • Metcalfe HC, Williams JE & Castka JF 1974, Modern Chemistry, Holt, Rinehart and Winston, New York, ISBN 0030894506
  • Meyer JS, Adams WJ, Brix KV, Luoma SM, Mount DR, Stubblefield WA & Wood CM (eds) 2005, Toxicity of Dietborne Metals to Aquatic Organisms, Proceedings from the Pellston Workshop on Toxicity of Dietborne Metals to Aquatic Organisms, 27 July–1 August 2002, Fairmont Hot Springs, British Columbia, Canada, Society of Environmental Toxicology and Chemistry, Pensacola, Florida, ISBN 1880611708
  • Mhiaoui S, Sar F, Gasser J 2003, 'Influence of the History of a Melt on the Electrical Resistivity of Cadmium–Antimony Liquid Alloys', Intermetallics, vol. 11, nos 11–12, pp. 1377–82, doi:10.1016/j.intermet.2003.09.008
  • Miller GJ, Lee C & Choe W 2002, 'Structure and Bonding Around the Zintl border', in G Meyer, D Naumann & L Wesermann (eds), Inorganic chemistry highlights, Wiley-VCH, Weinheim, pp. 21–53, ISBN 3527302654
  • Millot F, Rifflet JC, Sarou-Kanian V & Wille G 2002, 'High-Temperature Properties of Liquid Boron from Contactless Techniques', International Journal of Thermophysics, vol. 23, no. 5, pp. 1185–95, doi:10.1023/A:1019836102776
  • Mingos DMP 1998, Essential Trends in Inorganic Chemistry, Oxford University, Oxford, ISBN 0198501080
  • Moeller T 1954, Inorganic Chemistry: An Advanced Textbook, John Wiley & Sons, New York
  • Molina-Quiroz RC, Muñoz-Villagrán CM, de la Torre E, Tantaleán JC, Vásquez CC & Pérez-Donoso JM 2012, 'Enhancing the Antibiotic Antibacterial Effect by Sub Lethal Tellurite Concentrations: Tellurite and Cefotaxime Act Synergistically in Escherichia Coli', PloS (Public Library of Science) ONE, vol. 7, no. 4, doi:10.1371/journal.pone.0035452
  • Moody B 1991, Comparative Inorganic Chemistry, 3rd ed., Edward Arnold, London, ISBN 0713136790
  • Moore LJ, Fassett JD, Travis JC, Lucatorto TB & Clark CW 1985, 'Resonance-Ionization Mass Spectrometry of Carbon', Journal of the Optical Society of America B, vol. 2, no. 9, pp. 1561–5, doi:10.1364/JOSAB.2.001561
  • Moore JT 2011, Chemistry for Dummies, 2nd ed., John Wiley & Sons, New York, ISBN 1118092929
  • Morgan WC 1906, Qualitative Analysis as a Laboratory Basis for the Study of General Inorganic Chemistry, The Macmillan Company, New York
  • Morita A 1986, 'Semiconducting Black Phosphorus', Journal of Applied Physics A, vol. 39, no. 4, pp. 227–42, doi:10.1007/BF00617267
  • Moss TS 1952, Photoconductivity in the Elements, London, Butterworths
  • Murray JF 1928, 'Cable-Sheath Corrosion', Electrical World, vol. 92, Dec 29, pp. 1295–7, ISSN 00134457 Parameter error in {{issn}}: Invalid ISSN.
  • Nagao T, Sadowski1 JT, Saito M, Yaginuma S, Fujikawa Y, Kogure T, Ohno T, Hasegawa Y, Hasegawa S & Sakurai T 2004, 'Nanofilm Allotrope and Phase Transformation of Ultrathin Bi Film on Si(111)-7×7', Physical Review Letters, vol. 93, no. 10, pp. 105501–1–4, doi:10.1103/PhysRevLett.93.105501
  • Nemodruk AA & Karalova ZK 1969, Analytical Chemistry of Boron, R Kondor trans., Ann Arbor Humphrey Science, Ann Arbor, Michigan
  • Neuburger MC 1936, 'Gitterkonstanten für das Jahr 1936' (in German), Zeitschrift für Kristallographie, vol. 93, pp. 1–36, ISSN 00442968 Parameter error in {{issn}}: Invalid ISSN.
  • Nickelès J 1861, 'On a New Characteristic of the So-Called Semi-Metals', American Journal of Science, vol. 82, ISSN 00029599 Parameter error in {{issn}}: Invalid ISSN.
  • Nickless G 1968, Inorganic Sulphur Chemistry, Elsevier, Amsterdam
  • Nielsen FH 1998, 'Ultratrace Elements in Nutrition: Current Knowledge and Speculation', The Journal of Trace Elements in Experimental Medicine, vol. 11, pp. 251–74, doi:10.1002/(SICI)1520-670X(1998)11:2/3<251::AID-JTRA15>3.0.CO;2-Q
  • NIST (National Institute of Standards and Technology) 2010, Ground Levels and Ionization Energies for Neutral Atoms, by WC Martin, A Musgrove, S Kotochigova & JE Sansonetti, viewed 8 February 2013
  • NIST (National Institute of Standards and Technology) 2011, Astatine, viewed 8 February 2013
  • National Research Council 1984, The Competitive Status of the U.S. Electronics Industry: A Study of the Influences of Technology in Determining International Industrial Competitive Advantage, National Academy Press, Washington, DC, ISBN 0309033977
  • New Scientist 1975, 'Chemistry on the Islands of Stability', 11 Sep, p. 574, ISSN 10321233 Parameter error in {{issn}}: Invalid ISSN.
  • Oderberg DS 2007, Real Essentialism, Routledge, New York, ISBN 1134348851
  • Oxford English Dictionary 1989, 2nd ed., Oxford University, Oxford, ISBN 0198612133
  • Oganov AR, Chen J, Gatti C, Ma Y, Ma Y, Glass CW, Liu Z, Yu T, Kurakevych OO & Solozhenko VL 2009, 'Ionic High-Pressure Form of Elemental Boron', Nature, vol. 457, 12 Feb, pp. 863–8, doi:10.1038/nature07736
  • Oganov AR 2010, 'Boron Under Pressure: Phase Diagram and Novel High Pressure Phase,' in N Ortovoskaya N & L Mykola L (eds), Boron Rich Solids: Sensors, Ultra High Temperature Ceramics, Thermoelectrics, Armor, Springer, Dordrecht, pp. 207–25, ISBN 9048198232
  • Ogata S, Li J & Yip S 2002, 'Ideal Pure Shear Strength of Aluminium and Copper', Science, vol. 298, no. 5594, 25 October, pp. 807–10, doi:10.1126/science.1076652
  • Okakjima Y & Shomoji M 1972, Viscosity of Dilute Amalgams', Transactions of the Japan Institute of Metals, vol. 13, no. 4, pp. 255–8, ISSN 00214434 Parameter error in {{issn}}: Invalid ISSN.
  • Oldfield JE, Allaway WH, HA Laitinen, HW Lakin & OH Muth 1974, 'Tellurium', in Geochemistry and the Environment, Volume 1: The Relation of Selected Trace Elements to Health and Disease, US National Committee for Geochemistry, Subcommittee on the Geochemical Environment in Relation to Health and Disease, National Academy of Sciences, Washington, ISBN 0309022231
  • Olechna DJ & Knox RS 1965, 'Energy-Band Structure of Selenium Chains', Physical Review, vol. 140, no. 3A, pp. A986–93, doi:10.1103/PhysRev.140.A986
  • Oliwenstein L 2011, 'Caltech-Led Team Creates Damage-Tolerant Metallic Glass', California Institute of Technology, 12 January, viewed 8 February 2013
  • Olmsted J & Williams GM 1997, Chemistry, the Molecular Science, 2nd ed., Wm C Brown, Dubuque, Iowa, ISBN 0815184506
  • Orton JW 2004, The Story of Semiconductors, Oxford University, Oxford, ISBN 0198530838
  • Oxtoby DW, Gillis HP & Campion A 2008, Principles of Modern Chemistry, 6th ed., Thomson Brooks/Cole, Belmont, California, ISBN 0534493661
  • Parise JB, Tan K, Norby P, Ko Y & Cahill C 1996, 'Examples of Hydrothermal Titration and Real Time X-ray Diffraction in the Synthesis of Open Frameworks', MRS Proceedings, vol. 453, pp. 103–14, doi:10.1557/PROC-453-103
  • Parish RV 1977, The Metallic Elements, Longman, London, ISBN 0582442788
  • Parkes GD & Mellor JW 1943, Mellor's Nodern Inorganic Chemistry, Longmans, Green and Co., London
  • Parry RW, Steiner LE, Tellefsen RL & Dietz PM 1970, Chemistry: Experimental Foundations, Prentice-Hall/Martin Educational, Sydney, ISBN 0725301007
  • Partington 1944, A Text-book of Inorganic Chemistry, 5th ed., Macmillan, London
  • Partington JR 1964, A History of Chemistry, vol. 4, Macmillan, London
  • Pashaey BP & Seleznev VV 1973, 'Magnetic Susceptibility of Gallium-Indium Alloys in Liquid State', Russian Physics Journal, vol. 16, no. 4, pp. 565–6, doi:10.1007/BF00890855
  • Patel MR 2012, Introduction to Electrical Power and Power Electronics CRC Press, Boca Raton, ISBN 9781466556607
  • Pauling L 1988, General Chemistry, Dover Publications, New York, ISBN 0486656225
  • Pearson WB 1972, The Crystal Chemistry and Physics of Metals and Alloys, Wiley-Interscience, New York, ISBN 0471675407
  • Peryea FJ 1998, 'Historical Use of Lead Arsenate Insecticides, Resulting Soil Contamination and Implications for Soil Remediation, Proceedings', 16th World Congress of Soil Science, Montpellier, France, 20–26 August
  • Phillips CSG & Williams RJP 1965, Inorganic Chemistry, I: Principles and Non-metals, Clarendon Press, Oxford
  • Pinkerton J 1800, Petralogy. A Treatise on Rocks, vol. 2, White, Cochrane, and Co., London
  • Poltavtsev YG 1974, 'X-ray Diffraction Studies on Changes in the Nature of the Interatomic Bond and Close Range Order Structure During Melting of AIVBVI and AIIIBVCVIII Semiconductors,' Proceedings of all-union conference on chemical bond in semiconductors and submetals (in Russian), Nauka, Minsk
  • Poojary DM, Borade RB & Clearfield A 1993, 'Structural Characterization of Silicon Orthophosphate', Inorganica Chimica Acta, vol. 208, no. 1, pp. 23–9, doi:10.1016/S0020-1693(00)82879-0
  • Pourbaix M 1974, Atlas of Electrochemical Equilibria in Aqueous Solutions, 2nd English edition, National Association of Corrosion Engineers, Houston, ISBN 0915567989
  • Powell P 1988, Principles of Organometallic Chemistry, Chapman and Hall, London, ISBN 041242830X
  • Prakash GKS & Schleyer PvR (eds) 1997, Stable Carbocation Chemistry, John Wiley & Sons, New York, ISBN 0471594628
  • Prudenziati M 1977, IV. 'Characterization of Localized States in β-Rhombohedral Boron', in VI Matkovich (ed.), Boron and Refractory Borides, Springer-Verlag, Berlin, pp. 241–61, ISBN 038708181X
  • Puddephatt RJ & Monaghan PK 1989, The Periodic Table of the Elements, 2nd ed., Oxford University, Oxford, ISBN 0198555164
  • Pyykkö P 2012, 'Relativistic Effects in Chemistry: More Common Than You Thought', Annual Review of Physical Chemistry, vol. 63, pp. 45‒64 (56), doi:10.1146/annurev-physchem-032511-143755
  • Rao CNR & Ganguly P 1986, 'A New Criterion for the Metallicity of Elements', Solid State Communications, vol. 57, no. 1, pp. 5–6, doi:10.1016/0038-1098(86)90659-9
  • Rao KY 2002, Structural Chemistry of Glasses, Elsevier, Oxford, ISBN 0080439586
  • Rausch MD 1960, 'Cyclopentadienyl Compounds of Metals and Metalloids', Journal of Chemical Education, vol. 37, no. 11, pp. 568–78, doi:10.1021/ed037p568
  • Rayner-Canham G & Overton T 2006, Descriptive Inorganic Chemistry, 4th ed., WH Freeman, New York, ISBN 0716789639
  • Rayner-Canham G 2011, 'Isodiagonality in the Periodic Table', Foundations of chemistry, vol. 13, no. 2, pp. 121–9, doi:10.1007/s10698-011-9108-y
  • Reardon M 2005, 'IBM Doubles Speed of Germanium chips', CNET News, August 4, viewed 27 December 2013
  • Regnault MV 1853, Elements of Chemistry, vol. 1, 2nd ed., Clark & Hesser, Philadelphia
  • Reilly C 2002, Metal Contamination of Food, Blackwell Science, Oxford, ISBN 0632059273
  • Reilly 2004, The Nutritional Trace Metals, Blackwell, Oxford, ISBN 1405110406
  • Remy H 1956, Treatise on Inorganic Chemistry, vol. II, Elsevier, Amsterdam
  • Restrepo G, Mesa H, Llanos EJ & Villaveces JL 2004, 'Topological Study of the Periodic System', Journal of Chemical Information and Modelling, vol. 44, no. 1, pp. 68–75, doi:10.1021/ci034217z
  • Restrepo G, Llanos EJ & Mesa H 2006, 'Topological Space of the Chemical Elements and its Properties', Journal of Mathematical Chemistry, vol. 39, no. 2, pp. 401–16, doi:10.1007/s10910-005-9041-1
  • Řezanka T & Sigler K 2008, 'Biologically Active Compounds of Semi-Metals', Studies in Natural Products Chemistry, vol. 35, pp. 585–606, doi:10.1016/S1572-5995(08)80018-X
  • Rochow EG 1957, The Chemistry of Organometallic Compounds, John Wiley & Sons, New York
  • Rochow EG 1966, The Metalloids, DC Heath and Company, Boston
  • Rochow EG 1973, 'Silicon', in JC Bailar, HJ Emeléus, R Nyholm & AF Trotman-Dickenson (eds), Comprehensive Inorganic Chemistry, vol. 1, Pergamon, Oxford, pp. 1323–1467, ISBN 008015655X
  • Rochow EG 1977, Modern Descriptive Chemistry, Saunders, Philadelphia, ISBN 0721676286
  • Roher GS 2001, Structure and Bonding in Crystalline Materials, Cambridge University Press, Cambridge, ISBN 0521663792
  • Rossler K 1985, 'Handling of Astatine', pp. 140–56, in Kugler & Keller
  • Rothenberg GB 1976, Glass Technology, Recent Developments, Noyes Data Corporation, Park Ridge, New Jersey, ISBN 0815506090
  • Rouvray DH 1995, 'That Fuzzy Feeling in Chemistry', Chemistry in Britain, vol. 31, no. 71, pp. 544–6
  • Roza G 2009, Bromine, Rosen Publishing, New York, ISBN 1435850688
  • Russell AM & Lee KL 2005, Structure-Property Relations in Nonferrous Metals, Wiley-Interscience, New York, ISBN 047164952X
  • Sacks O 2001, Uncle Tungsten: Memories of a Chemical Boyhood, Alfred A. Knopf, New York, ISBN 0375704043
  • Salentine CG 1987, 'Synthesis, Characterization, and Crystal Structure of a New Potassium Borate, KB3O5•3H2O', Inorganic Chemistry, vol. 26, no. 1, pp. 128–32, doi:10.1021/ic00248a025
  • Samsonov GV 1968, Handbook of the Physiochemical Properties of the Elements, I F I/Plenum, New York
  • Sanderson RT 1960, Chemical Periodicity, Reinhold Publishing, New York
  • Savvatimskiy AI 2005, 'Measurements of the Melting Point of Graphite and the Properties of Liquid Carbon (a review for 1963–2003)', Carbon, vol. 43, no. 6, pp. 1115–42, doi:10.1016/j.carbon.2004.12.027
  • Savvatimskiy AI 2009, 'Experimental Electrical Resistivity of Liquid Carbon in the Temperature Range from 4800 to ~20,000 K', Carbon, vol. 47, no. 10, pp. 2322–8, doi:10.1016/j.carbon.2009.04.009
  • Schaefer JC 1968, 'Boron' in CA Hampel (ed.), The Encyclopedia of the Chemical Elements, Reinhold, New York, pp. 73–81
  • Schauss AG 1991, 'Nephrotoxicity and Neurotoxicity in Humans from Organogermanium Compounds and Germanium Dioxide', Biological Trace Element Research, vol. 29, no. 3, pp. 267–80, doi:10.1007/BF03032683
  • Schmidbaur H & Schier A 2008, 'A Briefing on Aurophilicity,' Chemical Society Reviews, vol. 37, pp. 1931–51, doi:10.1039/B708845K
  • Schroers J 2013, 'Bulk Metallic Glasses', Physics Today, vol. 66, no. 2, pp. 32–7, doi:10.1063/PT.3.1885
  • Schwartz MM 2002, Encyclopedia of Materials, Parts, and Finishes, 2nd ed., CRC Press, Boca Raton, Florida, ISBN 1566766613
  • Schwietzer GK and Pesterfield LL 2010, The Aqueous Chemistry of the Elements, Oxford University, Oxford, ISBN 019539335X
  • ScienceDaily 2012, 'Recharge Your Cell Phone With a Touch? New nanotechnology converts body heat into power', February 22, viewed 13 Jan 2013
  • Scott EC & Kanda FA 1962, The Nature of Atoms and Molecules: A General Chemistry, Harper & Row, New York
  • Secrist JH & Powers WH 1966, General Chemistry, D. Van Nostrand, Princeton, New Jersey
  • Segal BG 1989, Chemistry: Experiment and Theory, 2nd ed., John Wiley & Sons, New York, ISBN 0471849294
  • Sekhon BS 2012, 'Metalloid Compounds as Drugs', Research in Pharmaceutical Sciences, vol. 8, no. 3, pp. 145–58, ISSN 17359414 Parameter error in {{issn}}: Invalid ISSN.
  • Sequeira CAC 2011, 'Copper and Copper Alloys', in R Winston Revie (ed.), Uhlig's Corrosion Handbook, 3rd ed., John Wiley & Sons, Hoboken, New Jersey, pp. 757–86, ISBN 111811003X
  • Seybolt AU & Burke JE 1953, Procedures in Experimental Metallurgy, John Wiley & Sons, New York
  • Sharp DWA 1981, 'Metalloids', in Miall's Dictionary of Chemistry, 5th ed, Longman, Harlow, ISBN 0582351529
  • Sharp DWA 1983, The Penguin Dictionary of Chemistry, 2nd ed., Harmondsworth, Middlesex, ISBN 014051113X
  • Sherman E & Weston GJ 1966, Chemistry of the Non-metallic Elements, Pergamon Press, New York
  • Sidgwick NV 1950, The Chemical Elements and Their Compounds, vol. 1, Clarendon, Oxford
  • Siebring BR 1967, Chemistry, MacMillan, New York
  • Siekierski S & Burgess J 2002, Concise Chemistry of the Elements, Horwood, Chichester, ISBN 1898563713
  • Silberberg MS 2006, Chemistry: The Molecular Nature of Matter and Change, 4th ed., McGraw-Hill, New York, ISBN 0071116583
  • Simple Memory Art c. 2005, Periodic Table, EVA vinyl shower curtain, San Francisco
  • Slade S 2006, Elements and the Periodic Table, The Rosen Publishing Group, New York, ISBN 1404221654
  • Science Learning Hub 2009, 'The Essential Elements', The University of Waikato, viewed 16 January 2013
  • Smith DW 1990, Inorganic Substances: A Prelude to the Study of Descriptive Inorganic Chemistry, Cambridge University, Cambridge, ISBN 0521337380
  • Smith R 1994, Conquering Chemistry, 2nd ed., McGraw-Hill, Sydney, ISBN 0074701460
  • Snyder MK 1966, Chemistry: Structure and Reactions, Holt, Rinehart and Winston, New York
  • Soverna S 2004, 'Indication for a Gaseous Element 112', in U Grundinger (ed.), GSI Scientific Report 2003, GSI Report 2004-1, p. 187, ISSN 01740814 Parameter error in {{issn}}: Invalid ISSN.
  • Steele D 1966, The Chemistry of the Metallic Elements, Pergamon Press, Oxford
  • Stein L 1985, 'New Evidence that Radon is a Metalloid Element: Ion-Exchange Reactions of Cationic Radon', Journal of the Chemical Society, Chemical Communications, vol. 22, pp. 1631–2, doi:10.1039/C39850001631
  • Stein L 1987, 'Chemical Properties of Radon' in PK Hopke (ed.) 1987, Radon and its Decay products: Occurrence, Properties, and Health Effects, American Chemical Society, Washington DC, pp. 240–51, ISBN 0841210152
  • Steudel R 1977, Chemistry of the Non-metals: With an Introduction to atomic Structure and Chemical Bonding, Walter de Gruyter, Berlin, ISBN 3110048825
  • Steurer W 2007, 'Crystal Structures of the Elements' in JW Marin (ed.), Concise Encyclopedia of the Structure of Materials, Elsevier, Oxford, pp. 127–45, ISBN 0080451276
  • Stevens SD & Klarner A 1990, Deadly Doses: A Writer's Guide to Poisons, Writer's Digest Books, Cincinnati, Ohio, ISBN 0898793718
  • Stoker HS 2010, General, Organic, and Biological Chemistry, 5th ed., Brooks/Cole, Cengage Learning, Belmont California, ISBN 0495831468
  • Stott RW 1956, A Companion to Physical and Inorganic Chemistry, Longmans, Green and Co., London
  • Stuke J 1974, 'Optical and Electrical Properties of Selenium', in RA Zingaro & WC Cooper (eds), Selenium, Van Nostrand Reinhold, New York, pp. 174–297, ISBN 0442295758
  • Swalin RA 1962, Thermodynamics of Solids, John Wiley & Sons, New York
  • Swift EH & Schaefer WP 1962, Qualitative Elemental Analysis, WH Freeman, San Francisco
  • Swink LN & Carpenter GB 1966, 'The Crystal Structure of Basic Tellurium Nitrate, Te2O4•HNO3', Acta Crystallographica, vol. 21, no. 4, pp. 578–83, doi:10.1107/S0365110X66003487
  • Szpunar J, Bouyssiere B & Lobinski R 2004, 'Advances in Analytical Methods for Speciation of Trace Elements in the Environment', in AV Hirner & H Emons (eds), Organic Metal and Metalloid Species in the Environment: Analysis, Distribution Processes and Toxicological Evaluation, Springer-Verlag, Berlin, pp. 17–40, ISBN 3540208291
  • Taguena-Martinez J, Barrio RA & Chambouleyron I 1991, 'Study of Tin in Amorphous Germanium', in JA Blackman & J Tagüeña (eds), Disorder in Condensed Matter Physics: A Volume in Honour of Roger Elliott, Clarendon Press, Oxford, ISBN 019853938X, pp. 139–44
  • Taniguchi M, Suga S, Seki M, Sakamoto H, Kanzaki H, Akahama Y, Endo S, Terada S & Narita S 1984, 'Core-Exciton Induced Resonant Photoemission in the Covalent Semiconductor Black Phosphorus', Solid State Communications, vo1. 49, no. 9, pp. 867–70
  • Tao SH & Bolger PM 1997, 'Hazard Assessment of Germanium Supplements', Regulatory Toxicology and Pharmacology, vol. 25, no. 3, pp. 211–19, doi:10.1006/rtph.1997.1098
  • Taylor MD 1960, First Principles of Chemistry, D. Van Nostrand, Princeton, New Jersey
  • Thayer JS 1977, 'Teaching Bio-Organometal Chemistry. I. The Metalloids', Journal of Chemical Education, vol. 54, no. 10, pp. 604–6, doi:10.1021/ed054p604
  • The Economist 2012, 'Phase-Change Memory: Altered States', Technology Quarterly, September 1
  • The American Heritage Science Dictionary 2005, Houghton Mifflin Harcourt, Boston, ISBN 0618455043
  • The Chemical News 1897, 'Notices of Books: A Manual of Chemistry, Theoretical and Practical, by WA Tilden', vol. 75, no. 1951, p. 189
  • Thomas F, Bialek B & Hensel R 2013, 'Medical Use of Bismuth: The Two Sides of the Coin', Journal of Clinical Toxicology, special issue 3, article 4, doi:10.4172/2161-0495
  • Thomson, T. 1830, The History of Chemistry, volumes 1–2, Henry Colburn, and Richard Bentley, London
  • Tilden WA 1876, Introduction to the Study of Chemical Philosophy, D. Appleton and Co., New York
  • Timm JA 1944, General Chemistry, McGraw-Hill, New York
  • Togaya M 2000, 'Electrical Resistivity of Liquid Carbon at High Pressure', in MH Manghnani, W Nellis & MF.Nicol (eds), Science and Technology of High Pressure, proceedings of AIRAPT-17, Honolulu, Hawaii, 25–30 July 1999, vol. 2, Universities Press, Hyderabad, pp. 871–4, ISBN 8173713391
  • Tominaga J 2006, 'Application of Ge–Sb–Te Glasses for Ultrahigh Density Optical Storage', in AV Kolobov (ed.), Photo-Induced Metastability in Amorphous Semiconductors, Wiley-VCH, pp. 327–7, ISBN 3527608664
  • Träger F 2007, Springer Handbook of Lasers and Optics, Springer, New York, ISBN 9780387955797
  • Traynham JG 1989, 'Carbonium Ion: Waxing and Waning of a Name', Journal of Chemical Education, vol. 63, no. 11, pp. 930–3, doi:10.1021/ed063p930
  • Trivedi Y, Yung E & Katz DS 2013, 'Imaging in Fever of Unknown Origin', in BA Cunha (ed.), Fever of Unknown Origin, Informa Healthcare USA, New York, pp. 209–228, ISBN 0849336155
  • Turner M 2011, 'German E. Coli Outbreak Caused by Previously Unknown Strain', Nature News, 2 Jun, doi:10.1038/news.2011.345
  • Turova N 2011, Inorganic Chemistry in Tables, Springer, Heidelberg, ISBN 9783642204869
  • Tuthill G 2011, 'Faculty profile: Elements of Great Teaching', The Iolani School Bulletin, Winter, viewed 29 October 2011
  • Tyler PM 1948, From the Ground Up: Facts and Figures of the Mineral Industries of the United States, McGraw-Hill, New York
  • Tyler Miller G 1987, Chemistry: A Basic Introduction, 4th ed., Wadsworth Publishing Company, Belmont, California, ISBN 0534069126
  • Uden PC 2005, 'Speciation of Selenium,' in R Cornelis, J Caruso, H Crews & K Heumann (eds), Handbook of Elemental Speciation II: Species in the Environment, Food, Medicine and Occupational Health, John Wiley & Sons, Chichester, pp. 346–65, ISBN 0470855983
  • US Air Force Medical Service 1966, The Pharmacy Technician, Department of the Air Force, Washington
  • US Bureau of Naval Personnel 1965, Shipfitter 3 & 2, US Government Printing Office, Washington
  • US Environmental Protection Agency 1988, Ambient Aquatic Life Water Quality Criteria for Antimony (III), draft, Office of Research and Development, Environmental Research Laboratories, Washington
  • Van der Put PJ 1998, The Inorganic Chemistry of Materials: How to Make Things Out of Elements, Plenum, New York, ISBN 0306457318
  • Van Setten MJ, Uijttewaal MA, de Wijs GA & Groot RA 2007, 'Thermodynamic Stability of Boron: The Role of Defects and Zero Point Motion', Journal of the American Chemical Society, vol. 129, no. 9, pp. 2458–65, doi:10.1021/ja0631246
  • Vasáros L & Berei K 1985, 'General Properties of Astatine', pp. 107–28, in Kugler & Keller
  • Vernon RE 2013, 'Which Elements Are Metalloids?', Journal of Chemical Education, vol. 90, no. 12, pp. 1703–1707, doi:10.1021/ed3008457
  • Walters D 1982, Chemistry, Franklin Watts Science World series, Franklin Watts, London, ISBN 0531045811
  • Wanga WH, Dongb C & Shek CH 2004, 'Bulk Metallic Glasses', Materials Science and Engineering Reports, vol. 44, nos 2–3, pp. 45–89, doi:10.1016/j.mser.2004.03.001
  • Warren J & Geballe T 1981, 'Research Opportunities in New Energy-Related Materials', Materials Science and Engineering, vol. 50, no. 2, pp. 149–98, doi:10.1016/0025-5416(81)90177-4
  • Watt GW 1958, Basic Concepts in Chemistry, McGraw-Hill, New York
  • Wells AF 1984, Structural Inorganic Chemistry, 5th ed., Clarendon, Oxford, ISBN 0198553706
  • Whitten KW, Davis RE, Peck LM & Stanley GG 2007, Chemistry, 8th ed., Thomson Brooks/Cole, Belmont, California, ISBN 0495014494
  • Wiberg N 2001, Inorganic Chemistry, Academic Press, San Diego, ISBN 0123526515
  • Wilkie CA & Morgan AB 2009, Fire Retardancy of Polymeric Materials, CRC Press, Boca Raton, Florida, ISBN 1420083996
  • Witt AF & Gatos HC 1968, 'Germanium', in CA Hampel (ed.), The Encyclopedia of the Chemical Elements, Reinhold, New York, pp. 237–44
  • Woodward WE 1948, Engineering Metallurgy, Constable, London
  • WPI-AIM (World Premier Institute – Advanced Institute for Materials Research) 2012, 'Bulk Metallic Glasses: An Unexpected Hybrid', AIMResearch, Tohoku University, Sendai, Japan, 30 April
  • Wulfsberg G 1991, Principles of Descriptive Inorganic Chemistry, Brooks/Cole, Monterey, California, ISBN 0935702660
  • Wulfsberg G 2000, Inorganic Chemistry, University Science Books, Sausalito California, ISBN 1891389017
  • Yacobi BG & Holt DB 1990, Cathodoluminescence Microscopy of Inorganic Solids, Plenum, New York, ISBN 0306433141
  • Yasuda E, Inagaki M, Kaneko K, Endo M, Oya A & Tanabe Y 2003, Carbon Alloys: Novel Concepts to Develop Carbon Science and Technology, Elsevier Science, Oxford, pp. 3–11 et seq, ISBN 0080441637
  • Young RV & Sessine S (eds) 2000, World of Chemistry, Gale Group, Farmington Hills, Michigan, ISBN 0787636509
  • Young TF, Finley K, Adams WF, Besser J, Hopkins WD, Jolley D, McNaughton E, Presser TS, Shaw DP & Unrine J 2010, 'What You Need to Know About Selenium', in PM Chapman, WJ Adams, M Brooks, CJ Delos, SN Luoma, WA Maher, H Ohlendorf, TS Presser & P Shaw (eds), Ecological Assessment of Selenium in the Aquatic Environment, CRC, Boca Raton, Florida, pp. 7–45, ISBN 1439826773
  • Zhang TC, Lai KCK & Surampalli AY 2008, 'Pesticides', in A Bhandari, RY Surampalli, CD Adams, P Champagne, SK Ong, RD Tyagi & TC Zhang (eds), Contaminants of Emerging Environmental Concern, American Society of Civil Engineers, Reston, Virginia, ISBN 9780784410141, pp. 343–415
  • Zingaro RA 1994, 'Arsenic: Inorganic Chemistry', in RB King (ed.) 1994, Encyclopedia of Inorganic Chemistry, John Wiley & Sons, Chichester, pp. 192–218, ISBN 0471936200

Monographs

  • Brady JE, Humiston GE & Heikkinen H 1980, 'Chemistry of the Representative Elements: Part II, The Metalloids and Nonmetals', in General Chemistry: Principles and Structure, 2nd ed., SI version, John Wiley & Sons, New York, pp. 537–591, ISBN 0471063150
  • Chedd G 1969, Half-way Elements: The Technology of Metalloids, Doubleday, New York
  • Dunstan S 1968, 'The Metalloids', in Principles of Chemistry, D. Van Nostrand Company, London, pp. 407–39
  • Goldsmith RH 1982, 'Metalloids', Journal of Chemical Education, vol. 59, no. 6, pp. 526–527, doi:10.1021/ed059p526
  • Hawkes SJ 2001, 'Semimetallicity', Journal of Chemical Education, vol. 78, no. 12, pp. 1686–7, doi:10.1021/ed078p1686
  • Metcalfe HC, Williams JE & Castka JF 1974, 'Aluminum and the Metalloids', in Modern Chemistry, Holt, Rinehart and Winston, New York, pp. 538–57, ISBN 0030894506
  • Moeller T, Bailar JC, Kleinberg J, Guss CO, Castellion ME & Metz C 1989, 'Carbon and the Semiconducting Elements', in Chemistry, with Inorganic Qualitative Analysis, 3rd ed., Harcourt Brace Jovanovich, San Diego, pp. 751–75, ISBN 0155064924
  • Rieske M 1998, 'Metalloids', in Encyclopedia of Earth and Physical Sciences, Marshall Cavendish, New York, vol. 6, pp. 758–9, ISBN 0761405518 (set)
  • Rochow EG 1966, The Metalloids, DC Heath and Company, Boston
  • Vernon RE 2013, 'Which Elements are Metalloids?', Journal of Chemical Education, vol. 90, no. 12, pp. 1703–7, doi:10.1021/ed3008457

Template:Link GA