Nonmetal: Difference between revisions

Source: Wikipedia, the free encyclopedia.
Content deleted Content added
Extended confirmed users
15,284 edits
→‎Types {{anchor|Classes}}: Add new metal/nonmetal comparison paragraph
Tag: Reverted
Extended confirmed users
15,284 edits
→‎Types {{anchor|Classes}}: Remove old metal/nonmetal comparison paragraph
Tag: Reverted
Line 247: Line 247:


The greatest discrepancy between authors occurs in the metalloid "frontier territory".<ref>[[#Russell|Russell & Lee 2005, p. 419]]</ref> Some consider metalloids distinct from both metals and nonmetals, while others classify them as nonmetals.<ref>[[#Goodrich|Goodrich 1844, p. 264]]; [[#TheChemical1897|''The Chemical News'' 1897, p. 189]]; [[#Hampel|Hampel & Hawley 1976, pp. 174, 191]]; [[#Lewis|Lewis 1993, p. 835]]; [[#Hérold|Hérold 2006, pp. 149–50]]</ref> Some categorize certain metalloids as metals (e.g., arsenic and antimony due to their similarities to [[heavy metals]]).<ref>[[#Tyler|Tyler 1948, p. 105]]; [[#Reilly|Reilly 2002, pp. 5–6]]</ref>{{ambiguous|These two refs in the preceding tag should be either before the opening paren if they support the general statement or else before the closing paren if they support Ar & Sb. If they support both they should be after the close paren & period. If the two refs belong in different places they should be in separate ref tags.}}{{#tag:ref|Jones takes a philosophical or pragmatic view to these questions. He writes: "Though classification is an essential feature of all branches of science, there are always hard cases at the boundaries. The boundary of a class is rarely sharp{{nbsp}}... Scientists should not lose sleep over the hard cases. As long as a classification system is beneficial to economy of description, to structuring knowledge and to our understanding, and hard cases constitute a small minority, then keep it. If the system becomes less than useful, then scrap it and replace it with a system based on different shared characteristics".<ref name="Jones">[[#Jones|Jones 2010, pp. 169–71]]</ref>|group=n}} This article includes metalloids for comparative purposes{{#tag:ref|For a related comparison of the [[properties of metals, metalloids, and nonmetals]], see [[#RDM|Rudakiya & Patel (2021), p. 36]]|group=n}} and due to their relatively low densities, high electronegativity, and chemical behavior.<ref name="Bailar"/>
The greatest discrepancy between authors occurs in the metalloid "frontier territory".<ref>[[#Russell|Russell & Lee 2005, p. 419]]</ref> Some consider metalloids distinct from both metals and nonmetals, while others classify them as nonmetals.<ref>[[#Goodrich|Goodrich 1844, p. 264]]; [[#TheChemical1897|''The Chemical News'' 1897, p. 189]]; [[#Hampel|Hampel & Hawley 1976, pp. 174, 191]]; [[#Lewis|Lewis 1993, p. 835]]; [[#Hérold|Hérold 2006, pp. 149–50]]</ref> Some categorize certain metalloids as metals (e.g., arsenic and antimony due to their similarities to [[heavy metals]]).<ref>[[#Tyler|Tyler 1948, p. 105]]; [[#Reilly|Reilly 2002, pp. 5–6]]</ref>{{ambiguous|These two refs in the preceding tag should be either before the opening paren if they support the general statement or else before the closing paren if they support Ar & Sb. If they support both they should be after the close paren & period. If the two refs belong in different places they should be in separate ref tags.}}{{#tag:ref|Jones takes a philosophical or pragmatic view to these questions. He writes: "Though classification is an essential feature of all branches of science, there are always hard cases at the boundaries. The boundary of a class is rarely sharp{{nbsp}}... Scientists should not lose sleep over the hard cases. As long as a classification system is beneficial to economy of description, to structuring knowledge and to our understanding, and hard cases constitute a small minority, then keep it. If the system becomes less than useful, then scrap it and replace it with a system based on different shared characteristics".<ref name="Jones">[[#Jones|Jones 2010, pp. 169–71]]</ref>|group=n}} This article includes metalloids for comparative purposes{{#tag:ref|For a related comparison of the [[properties of metals, metalloids, and nonmetals]], see [[#RDM|Rudakiya & Patel (2021), p. 36]]|group=n}} and due to their relatively low densities, high electronegativity, and chemical behavior.<ref name="Bailar"/>

A broadly comparable range of types occurs among the metals, from [[Reactivity_series#Comparison_with_electronegativity_values|highly reactive to less reactive (even noble)]]. On the left side of the periodic table, and below its main body, are highly to fairly reactive metals, such as [[sodium]], [[calcium]] and [[uranium]]. Towards the middle of the periodic table are [[transition metal]]s, such as [[scandium]], [[iron]] and [[nickel]], of high to low reactivity. To the right of the transition metals, (from group 13 onwards) are metals such as [[tin]] and [[lead]], none of which are particularly reactive. A subset of the transition metals (including [[platinum]] and [[gold]]) are referred to as [[noble metal]]s on account of their reluctance to engage in chemical activity.<ref>[[#Parish|Parish 1977, pp. 37, 112, 115, 145, 163, 182]]</ref>{{#tag:ref|'''1.''' For aluminium, Whitten and Davis<ref>[[#Whitten1996|Whitten & Davis 1996, p. 853]]</ref> write, "[It] is quite reactive , but a thin, transparent film of Al<sub>2</sub>O<sub>3</sub> forms when Al comes into contact with air. This protects it from further oxidation For this reason it is even passive toward nitric acid (HNO<sub>3</sub>), a strong oxidizing agent. When the oxide coating is sanded off, Al reacts vigorously with HNO3."<br>
'''2.''' For the transition metal [[manganese]], Parish<ref>[[#Parish|Parish 1977, pp. 53]]</ref> writes: "Very reactive. Reacts with water." Russell & Lee<ref>[[#Russell|Russell & Lee 2005, p. 247]]</ref> add, "Mn reacts with all [[mineral acid]]s, and it even slowly dissociates pure water, liberating H<sub>2</sub> gas as it forms [[manganese hydroxide|Mn hydroxide]]."|group=n}}


=== Noble gases ===
=== Noble gases ===

Revision as of 06:29, 15 November 2023

Nonmetals in their periodic table context
  usually/always counted as a nonmetal
  sometimes counted as a nonmetal
At astatine's status is unclear; while usually counted as a nonmetal, relativistic effects suggest it may be a metal[1]
Cn Fl Og copernicium, flerovium, and/or oganesson may turn out to be nonmetallic however their status has not been confirmed.

A nonmetal is a chemical element that mostly lacks metallic properties. Seventeen elements are generally considered nonmetals, though some authors recognize more or fewer depending on the properties considered most representative of metallic or nonmetallic character. Some borderline elements further complicate the situation.

Nonmetals tend to have low

acidic
.

The two lightest nonmetals, hydrogen and

oceans and biosphere.[n 1]

The distinct properties of nonmetallic elements allow for specific uses that metals often cannot achieve. Elements like hydrogen, oxygen, carbon, and nitrogen are essential building blocks for life itself. Moreover, nonmetallic elements are integral to industries such as electronics, energy storage, agriculture, and chemical production.

Most nonmetallic elements were not identified until the 18th and 19th centuries. While a distinction between metals and other minerals had existed since antiquity, a basic classification of chemical elements as metallic or nonmetallic emerged only in the late 18th century. Since then nigh on two dozen properties have been suggested as single criteria for distinguishing nonmetals from metals.

Definition and applicable elements

Properties mentioned hereafter refer to the elements in their most stable forms in ambient conditions unless otherwise mentioned
tarnishing) vaporises rather than melts when heated. The vapor is lemon-yellow and smells like garlic.[2] The chemistry of arsenic is predominately nonmetallic in nature.[3]

Nonmetal

electrical conductivity; and a capacity to (mostly) produce basic oxides when combined with oxygen.[4][n 2]

There is no precise definition of a nonmetal;[6] any list of such is open to debate and revision.[7] Which elements are included depends on the properties regarded as most representative of nonmetallic or metallic character.[n 3]

These fourteen elements are effectively always recognized as nonmetals:[7][8]

Hydrogen, Nitrogen, Oxygen, Sulfur
Fluorine, Chlorine, Bromine, Iodine
Helium, Neon, Argon, Krypton, Xenon, Radon

Three more are commonly called nonmetals, but some sources list them as

metalloids:[9]

Carbon, Phosphorus, Selenium

The six elements most commonly recognized as metalloids, which are typically seen as intermediates between metals and nonmetals, are here counted as a type of nonmetal due to their relatively low densities and predominantly nonmetallic chemistry, and for comparative purposes:[10]

Boron, Silicon, Germanium, Arsenic, Antimony, Tellurium

Of the 118 known elements,[11] roughly 20% are classified as nonmetals.[12] Opinions differ as to the status of astatine. Its rarity and extreme radioactivity has resulted in it being frequently ignored in the literature.[13] With no comprehensive understanding of its properties, its classification remains uncertain. As a halogen it has usually been presumed to be a nonmetal.[14] Chemically, studies on trace quantities of astatine, which are not necessarily reliable,[15] have demonstrated characteristics of both metals and nonmetals.[16] Alternatively, given the near-metallic character of its lighter congener iodine,[n 4] a succession of authors suggest astatine may be a metal.[18] A 2013 study based on relativistic chemistry concluded that it would be a monatomic metal with a close-packed crystalline structure,[19] but this has not been experimentally verified. Astatine is not considered further in this article due to uncertainty as to its behavior and status. The superheavy elements copernicium (element 112), flerovium (114), and oganesson (118) may or may not turn out to be nonmetals; their status has not been confirmed.[20]

General properties

Physical

Variety in color and form
of some nonmetallic elements
rhombohedral phase
A shiny grey-black cuboid nugget with a rough surface.
Metallic appearance of carbon as graphite
A pale blue liquid in a clear beaker
Blue color of liquid oxygen
A glass tube, is inside a larger glass tube, has some clear yellow liquid in it
Pale yellow liquid fluorine in a cryogenic bath
Yellow powdery chunks
Sulfur as a yellow powder
A small capped jar a quarter filled with a very dark liquid
Liquid bromine at room temperature
Shiny violet-black colored crystalline shards.
Metallic appearance of iodine under white light
A partly filled ampoule containing a colorless liquid

About half of nonmetallic elements are gases; most of the rest are shiny solids. Bromine, the only liquid, is so volatile that it is usually topped by a layer of its fumes; sulfur is the only colored solid nonmetal.[n 5] The gaseous and liquid nonmetals have very low densities, melting and boiling points, and are poor conductors of heat and electricity.[23] The solid elements have low densities and low mechanical and structural strength (being brittle or crumbly),[24] but a wide range of electrical conductivity.[n 6]

These diverse forms are caused by varied internal structures and bonding arrangements. Nonmetals existing as discrete atoms like xenon, or as small molecules, such as oxygen, sulfur, and bromine, have low melting and boiling points; many are gases at room temperature, as they are held together by weak London dispersion forces acting between their atoms or molecules.[28] In contrast, nonmetals that form giant structures, such as chains of up to 1,000 atoms (e.g., selenium),[29] sheets (e.g., carbon as graphite),[30] or three-dimensional lattices (e.g., silicon)[31] have higher melting and boiling points, and are all solids, as it takes more energy to overcome their stronger covalent bonds.[32] Nonmetals closer to the left or bottom of the periodic table, often have some weak metallic interactions between their molecules, chains, or layers, consistent with their proximity to the metals; this occurs in boron,[33] carbon,[34] phosphorus,[35] arsenic,[36] selenium,[37] antimony,[38] tellurium[39] and iodine.[40]

The structures of nonmetallic elements differ from those of metals primarily due to variations in valence electrons and atomic size. Metals typically have fewer valence electrons than available orbitals, leading them to share electrons with many nearby atoms, resulting in centrosymmetrical crystalline structures.[41] In contrast, nonmetals share only the electrons required to achieve a noble gas electron configuration.[42] For example, nitrogen forms diatomic molecules featuring a triple bonds between each atom, both of which thereby attain the configuration of the noble gas neon; while antimony's larger atomic size prevents triple bonding, resulting in buckled layers in which each antimony atom is singly bonded with three other nearby atoms.[43]

Nonmetals vary greatly in appearance. The shininess of boron, graphitic carbon, silicon, black phosphorus, germanium, arsenic, selenium, antimony, tellurium, and iodine is a result of their structures featuring varying degrees of delocalized (free-moving) electrons that scatter incoming visible light.[44] The colored nonmetals (sulfur, fluorine, chlorine, bromine) absorb some colors (wavelengths) and transmit the complementary or opposite colors. For example, chlorine’s "familiar yellow-green colour ... is due to a broad region of absorption in the violet and blue regions of the spectrum".[45][n 7] For the colorless nonmetals (hydrogen, nitrogen, oxygen, and the noble gases), their electrons are held sufficiently strongly so that no absorption happens in the visible part of the spectrum, and all visible light is transmitted.[47]

The electrical and thermal conductivities of nonmetals, along with the brittle nature of solid nonmetals are likewise related to their internal arrangements. Whereas good conductivity and plasticity (malleability, ductility) are ordinarily associated with the presence of

free-moving and evenly distributed electrons in metals,[48] the electrons in nonmetals typically lack such mobility.[49] Among nonmetallic elements, good electrical and thermal conductivity is seen only in carbon (as graphite, along its planes), arsenic, and antimony.[n 8] Good thermal conductivity otherwise occurs only in boron, silicon, phosphorus, and germanium;[25] such conductivity is transmitted though vibrations of the crystalline lattices of these elements.[50] Moderate electrical conductivity is observed in boron, silicon, phosphorus, germanium, selenium, tellurium, and iodine.[n 9] Plasticity occurs under limited circumstances in carbon, as seen in exfoliated (expanded) graphite[52][53] and carbon nanotube wire,[54] in white phosphorus (soft as wax, pliable and can be cut with a knife, at room temperature),[55] in plastic sulfur,[56] and in selenium which can be drawn into wires from its molten state.[57]

The physical differences between metals and nonmetals arise from internal and external atomic forces. Internally, the

positive charge stemming from the protons in an atom's nucleus acts to hold the atom's outer electrons in place. Externally, the same electrons are subject to attractive forces from protons in neighboring atoms. When the external forces are greater than, or equal to, the internal force, the outer electrons are expected to become relatively free to move between atoms, and metallic properties are predicted. Otherwise nonmetallic properties are expected.[58]

Allotropes

a haphazard aggregate of brownish crystals
Brownish crystals of buckminsterfullerene60), a semiconducting allotrope of carbon
Some chemistry-based typical
differences between metals and nonmetals[59]
Aspect Metals Nonmetals
Electronegativity Lower than nonmetals,
with some exceptions[60]
Relatively high
Chemical
bonding
Seldom form
covalent bonds
Frequently form
covalent bonds
Metallic bonds (alloys)
between metals
Covalent bonds
between nonmetals
Ionic bonds
between nonmetals and metals
Oxidation
states
Positive Negative or positive
Oxides Basic in lower oxides;
increasingly acidic
in higher oxides
Acidic;
never basic[61]
In aqueous
solution
[62]
Exist as
cations
Exist as
anions
or oxyanions

Many nonmetallic elements exhibit a range of allotropic forms, each with distinct physical properties that may vary between metallic and nonmetallic.

amorphous[66] and paracrystalline (mixed amorphous and crystalline)[67]
variations. Allotropes also occur for the other unclassified nonmetals, the metalloids, and iodine among the halogen nonmetals.[68]

Chemical

Nonmetals have relatively high values of electronegativity, and their oxides are therefore usually acidic. Exceptions occur when the oxidation state is low, the nonmetal is not very electronegative, or both: thus, water (H2O) is amphoteric[69] and nitrous oxide (N2O) is neutral.[70][n 10]

They tend to gain or share electrons during chemical reactions, in contrast to metals which tend to donate electrons. This behavior is closely related to the stability of electron configurations in the noble gases, which have complete outer shells. Nonmetals generally gain enough electrons to attain the electron configuration of the following noble gas, while metals tend to lose electrons, achieving the electron configuration of the preceding noble gas. These tendencies in nonmetallic elements are succinctly summarized by the duet and octet rules of thumb.

Furthermore, nonmetals typically exhibit higher

standard reduction potentials than metals. Generally, the higher these values are (including electronegativity) the more nonmetallic the element tends to be.[73] For example, the chemically very active nonmetals fluorine, chlorine, bromine, and iodine have an average electronegativity of 3.19—a figure[n 11] higher than that of any individual metal. On the other hand, the 2.05 average[n 12] of the chemically weak metalloid nonmetals falls within the 0.70 to 2.54 range of metals.[74]

The chemical distinctions between metals and nonmetals primarily stem from the attractive force between the positive nuclear charge of an individual atom and its negatively charged outer electrons. From left to right across each period of the periodic table, the nuclear charge increases in tandem with the number of protons in the atomic nucleus.[75] Consequently, there is a corresponding reduction in atomic radius[76] as the heightened nuclear charge draws the outer electrons closer to the nucleus core.[77] In metals, the impact of the nuclear charge is generally weaker compared to nonmetallic elements. As a result, in chemical bonding, metals tend to lose electrons, leading to the formation of positively charged or polarized atoms or ions, while nonmetals tend to gain these electrons due to their stronger nuclear charge, resulting in negatively charged ions or polarized atoms.[78]

The number of compounds formed by nonmetals is vast.

teflon ((C
2
F
4
)n), used to create non-stick coatings for pans and other cookware.[84]

Complications

Adding complexity to the chemistry of the nonmetals are anomalies occurring in the first row of each

periodic table block
; non-uniform periodic trends; higher oxidation states; and property overlaps with metals.

First row anomaly

H and He are in the first row of the s-block. B through Ne take up the first row of the p-block. Sc through Zn occupy the first row of the d-block. La to Yb make up the first row of the f block. The elements within scope of the article are hydrogen, helium, boron, carbon, nitrogen, oxygen, fluorine, neon, silicon, phosphorus, sulfur, chlorine, argon, germanium, arsenic, selenium, bromine, krypton, antimony, tellurium, iodine, xenon, and radon.
Periodic table highlighting the first row of each block.[n 13] Helium (He), as a noble gas, is normally shown over neon (Ne) with the rest of the noble gases. The elements within scope of this article are inside the thick black borders. The status of oganesson (Og, element 118) is not yet known.

Starting with hydrogen, the

double helix; shapes the three-dimensional forms of proteins; and even raises water's boiling point high enough to make a decent cup of tea."[87]

Hydrogen and helium, as well as boron through neon, have unusually small atomic radii. This phenomenon arises because the

2p subshells lack inner analogues (meaning there is no zero shell and no 1p subshell), and they therefore experience no electron repulsion effects, unlike the 3p, 4p, and 5p subshells of heavier elements.[88] A a result, ionization energies and electronegativities among these elements are higher than what periodic trends would otherwise suggest. The compact atomic radii of carbon, nitrogen, and oxygen facilitate the formation of double or triple bonds.[89]

While it would normally be expected, on electron configuration consistency grounds, that hydrogen and helium would be placed atop the s-block elements, the significant first row anomaly shown by these two elements justifies alternative placements. Hydrogen is occasionally positioned above fluorine, in group 17, rather than above lithium in group 1. Helium is commonly placed above neon, in group 18, rather than above beryllium in group 2.[90]

Secondary periodicity

A graph with a vertical electronegativity axis and a horizontal atomic number axis. The five elements plotted are O, S, Se, Te and Po. The electronegativity of Se looks too high, and causes a bump in what otherwise be a smooth curve.
Electronegativity values of the group 16 chalcogen elements showing a W-shaped alternation or secondary periodicity going down the group

A

fourteen f-block metals located between barium and lutetium, ultimately leading to atomic radii that are smaller than expected for elements from hafnium (Hf) onward.[93]

Higher oxidation states

The larger atomic radii of the heavier group 15–18 nonmetals enable higher

bulk coordination numbers, and result in lower electronegativity values that better tolerate higher positive charges. The elements involved are thereby able to exhibit oxidation states other than would be indicated by the octet rule, that is 3, 2, 1, or 0 in, for example, nitrogen triiodide (NI3), hydrogen sulfide (H2S), hydrogen fluoride (HF), and elemental xenon (Xe). Thus, the maximum possible oxidation state increases from +5 in group 15 to +8 in group 18, for example, in phosphorus pentachloride (PCl5), sulfur hexafluoride (SF6), iodine heptafluoride (IF7), and xenon difluoride (XeF2).[94]

Property overlaps with metals

While certain elements have traditionally been classified as nonmetals and others as metals, some overlapping of properties occurs. Writing early in the twentieth century, by which time the era of modern chemistry had been well-established,[95] Humphrey[96] observed that:

... these two groups, however, are not marked off perfectly sharply from each other; some nonmetals resemble metals in certain of their properties, and some metals approximate in some ways to the non-metals.

Examples of metal-like properties occurring in nonmetallic elements include:

  • the electrical conductivity of graphite exceeds that of some metals;[n 15]
  • selenium can be drawn into a wire;[57]
  • just over half of nonmetallic elements can form homopolyatomic cations;[n 16] and
  • silicon has an electronegativity (1.9) comparable with metals such as cobalt (1.88), copper (1.9), nickel (1.91) and silver (1.93).[74]

Types

Noble gases: He, Ne, Ar, Kr, Xe, Rn; Halogen nonmetals: F, Cl, Br, I; Unclassified nonmetals: H, C, N, P, O, S, Se; Metalloids: B, Si, Ge, As, Sb, Te. Nearby metals are Al, Ga, In, Tl; Sn, Pb; Bi; Po; and At. The nonmetals N, S and I are shown as moderately strong oxidizing agents; O, F, Cl and Br are relatively strong oxidizing agents.
Periodic table excerpt highlighting the four types of nonmetals. Hydrogen is typically found in group 1, but occasionally placed in group 17.
† moderately strong oxidizing agents
‡ strong oxidizing agents[n 17]

Encyclopaedia Britannica recognizes noble gases, halogens, and other nonmetals, and splits the elements commonly recognized as metalloids between "other metals" and "other nonmetals".[111] On the other hand, the Royal Society of Chemistry periodic table uses different colors for each of its eight main groups, with nonmetals represented in seven of these.[112]

Traversing the periodic table from right to left, three or four types of nonmetals can be discerned:

  • the relatively inert noble gases;[113]
  • a set of chemically strong halogen elements—fluorine, chlorine, bromine and iodine—sometimes referred to as nonmetal halogens[114] or halogen nonmetals[115] (as used here) or stable halogens;[116]
  • a set of unclassified nonmetals, encompassing elements like hydrogen, carbon, nitrogen, and oxygen, for which there is no widely recognized collective name;[n 19] and
  • the chemically weak nonmetallic metalloids[125] which are sometimes considered nonmetals and sometimes considered a third category distinct from metals and nonmetals.[n 20]

In the periodic table, metalloids – so metallic they are often not considered nonmetals – come beside

noble metals
buried within transition metals.

The boundaries between these sets of nonmetals are not sharp.

cationic behavior, which is unusual for a nonmetal.[128]

The greatest discrepancy between authors occurs in the metalloid "frontier territory".[129] Some consider metalloids distinct from both metals and nonmetals, while others classify them as nonmetals.[130] Some categorize certain metalloids as metals (e.g., arsenic and antimony due to their similarities to heavy metals).[131][ambiguous][n 22] This article includes metalloids for comparative purposes[n 23] and due to their relatively low densities, high electronegativity, and chemical behavior.[125]

Noble gases

a glass tube, held upside down by some tongs, has a clear-looking ice-like plug in it which is slowly melting judging from the clear drops falling out of the open end of the tube
A small (about 2 cm long) piece of rapidly melting argon ice

Six nonmetals are classified as noble gases: helium, neon, argon, krypton, xenon, and the radioactive radon. In conventional periodic tables they occupy the rightmost column. They are called noble gases due to their exceptionally low

chemical reactivity.[113]

These elements exhibit remarkably similar properties, characterized by their colorlessness, odorlessness, and nonflammability. Due to their closed outer electron shells, noble gases possess feeble

interatomic forces of attraction, leading to exceptionally low melting and boiling points.[132] As a consequence, they all exist as gases under standard conditions, even those with atomic masses surpassing many typically solid elements.[133]

Chemically, the noble gases exhibit relatively high ionization energies, negligible or negative electron affinities, and high to very high electronegativities. The number of compounds formed by noble gases is in the hundreds and continues to expand,[134] with most of these compounds involving the combination of oxygen or fluorine with either krypton, xenon, or radon.[135]

About 1015 tonnes of noble gases are present in the Earth's atmosphere.

intermetallic compounds.[n 24] This could explain why "studies of the Earth's atmosphere have shown that more than 90% of the expected amount of Xe is depleted."[140]

Halogen nonmetals

table salt
(NaCl)
Corrosive chlorine, a halogen nonmetal, combines with highly reactive sodium to form stable, unreactive table salt.

Although the halogen nonmetals are notably reactive and corrosive elements, they can also be found in everyday compounds like

table salt (NaCl); swimming pool disinfectant (NaBr); or food supplements (KI). The term "halogen" itself means "salt former".[141]

Physically, fluorine and chlorine exist as pale yellow and yellowish-green gases, respectively, while bromine is a reddish-brown liquid, typically covered by a layer of its fumes; iodine, when observed under white light, appears as a metallic-looking[100] solid. Electrically, the first three elements function as insulators while iodine behaves as a semiconductor (along its planes).[142]

Chemically, the halogen nonmetals exhibit high ionization energies, electron affinities, and electronegativity values, and are mostly relatively strong

covalent compounds with metals.[n 25] The highly reactive and strongly electronegative nature of the halogen nonmetals epitomizes nonmetallic character.[149]

The halogen nonmetals are commonly found in salt-related minerals. Fluorine, for instance, is present in fluorite (CaF2), a mineral found widely. Chlorine, bromine, and iodine are typically found in brines. Exceptionally, a study reported in 2012 noted the presence of 0.04% native fluorine (F
2
) by weight in antozonite, attributing these inclusions to radiation from tiny amounts of uranium.[150]

Metalloids

a cluster of bright cherry-red crystals
A crystal of realgar, also known as "ruby sulphur" or "ruby of arsenic", an arsenic sulfide mineral As4S4. The two elements involved each have a predominately nonmetallic chemistry.[151]

The six elements more commonly recognized as metalloids are boron, silicon, germanium, arsenic, antimony, and tellurium, all of which have a metallic appearance. (There are other elements less commonly recongised as metalloids, including carbon, aluminium, selenium and polonium. They have metallic and nonmetallic properties, but one or the other kind predominates.) On a standard periodic table, they occupy a diagonal region within the p-block extending from boron at the upper left to tellurium at the lower right, along the dividing line between metals and nonmetals shown on some tables.[9]

They are brittle and poor-to-good conductors of heat and electricity. Specifically, boron, silicon, germanium, and tellurium are semiconductors. Arsenic and antimony have the

forms.[9]

Chemically, metalloids generally behave like (weak) nonmetals. Among the nonmetallic elements they tend to have the lowest ionization energies, electron affinities, and electronegativity values, and are relatively weak oxidizing agents. Additionally, they tend to form alloys when combined with metals.[9]

The metalloids are commonly found combined with oxygen, sulfur, or, in the case of tellurium, gold or silver.

silica (sand). Germanium, arsenic, and antimony are mainly found as components of sulfide ores. Tellurium is often found in telluride minerals of gold or silver. In some instances, native forms of arsenic, antimony, and tellurium have been reported.[153]

Unclassified nonmetals

After classifying the nonmetallic elements into noble gases, halogens, and metalloids, the remaining seven nonmetals are hydrogen, carbon, nitrogen, oxygen, phosphorus, sulfur, and selenium.

In their most stable forms, three of these are colorless gases (H, N, O); three have a metal-like appearance (C, P, Se); and one appears yellow (S). Electrically, graphitic carbon behaves as a semimetal along its planes[155] and a semiconductor perpendicular to its planes;[156] phosphorus and selenium are semiconductors;[157] while hydrogen, nitrogen, oxygen, and sulfur are insulators.[n 26]

These elements are often considered too diverse to merit a collective classification,[159] and have been referred to as other nonmetals,[160] or simply as nonmetals, located between the metalloids and the halogens.[161] As a result, their chemistry is typically taught disparately, according to their respective periodic table groups:[162] hydrogen in group 1; the group 14 nonmetals (including carbon, and possibly silicon and germanium); the group 15 nonmetals (including nitrogen, phosphorus, and possibly arsenic and antimony); and the group 16 nonmetals (including oxygen, sulfur, selenium, and possibly tellurium). Authors may choose other subdivisions based on their preferences.[n 27]

Hydrogen, in particular, behaves in some respects like a metal and in others like a nonmetal.

transition metals.[167] Conversely, it is an insulating diatomic gas, akin to the nonmetals nitrogen, oxygen, fluorine and chlorine. In chemical reactions, it tends to ultimately attain the electron configuration of helium (the following noble gas) behaving in this way as a nonmetal.[168] It attains this configuration by forming a covalent or ionic bond[169] or, if it has initially given up its electron, by attaching itself to a lone pair of electrons.[170]

Some or all of these nonmetals share several properties. Being less reactive than the halogens,

interstitial or refractory) compounds[177] due to their relatively small atomic radii and sufficiently low ionization energies.[159] They also exhibit a tendency to bond to themselves, particularly in solid compounds.[178] Additionally, diagonal periodic table relationships among these nonmetals mirror similar relationships among the metalloids.[179]

Unclassified nonmetals are typically found in elemental forms or in association with other elements:[152]

Abundance, sources, and uses

Abundance of nonmetallic elements

Domain Abundance[n 28] of
main components
Next most
abundant
Universe[188] H 70.5%, He 27.5% O 1%
Atmosphere[189] N 78%, O 21% Ar 0.5%
Hydrosphere[189] O 66.2%, H 33.2% Cl 0.3%
Biomass[190] O 63%, C 20%, H 10% N 3.0%
Crust[189] O 61%, Si 20% H 2.9%

Hydrogen and helium dominate the universe, making up an estimated 98% of all ordinary matter by mass.[n 29] Oxygen, the next most abundant element, constitutes around 1% of the universe's composition.[192]

Five nonmetals—hydrogen, carbon, nitrogen, oxygen, and silicon—dominate the accessible structure of the earth, forming the vast majority of the

atmosphere, hydrosphere, and biomass as shown in the accompanying table.[193]

Sources of nonmetallic elements

Group (1, 13−18) Period
13 14 15 16 1/17 18 (1−6)
  H He 1
  B C N O F Ne 2
  Si P S Cl Ar 3
  Ge As Se Br Kr 4
  Sb Te I Xe 5
  Rn 6

Nonmetals and metalloids are extracted in their raw forms from:[172]

   mineral ores—boron (
silica); phosphorus (phosphates); antimony (stibnite, tetrahedrite); iodine (in sodium iodate and sodium iodide
);
   mining byproducts—germanium (zinc ores); arsenic (copper and lead ores); selenium and tellurium (copper ores); and radon (uranium-bearing ores);
   liquid air—nitrogen, oxygen, neon, argon, krypton, xenon;
   natural gas—hydrogen (methane), helium, sulfur (hydrogen sulfide); and
   seawater brine—chlorine, bromine, iodine.

Uses of nonmetallic elements

Nearly all nonmetals have uses in:
pharmaceuticals
Most nonmetals have uses in:
dyestuffs
Some nonmetals have uses in or as:
welding gases, and vulcanization
Metalloids have uses in:
semiconductors

Nonmetallic elements have distinct properties[195] that enable a wide range of natural and technological uses. In living organisms, hydrogen, oxygen, carbon, and nitrogen serve as the foundational building blocks of life. Some key technological uses of nonmetallic elements are in lighting and lasers, medicine and pharmaceuticals, and ceramics and plastics. The accompanying table groups nonmetallic elements according to their uses.

Some specific uses of later-discovered or rarer nonmetallic elements include:

  • Black phosphorus, first reported in 1916,
    thermoelectric materials.[200]

History, background, and taxonomy

Discovery

; the glow emanates from the combustion of phosphorus inside the flask.

Most nonmetallic elements were identified during the 18th and 19th centuries. However, a few nonmetals were recognized in ancient times and later historical periods. Carbon, sulfur, and antimony were among the early nonmetals known to humanity. The discovery of arsenic can be traced back to the Middle Ages, credited to the work of Albertus Magnus. A significant moment in the history of nonmetal discovery occurred in 1669 when Hennig Brand successfully isolated phosphorus from urine. Helium, identified in 1868, holds a unique distinction as the only element not initially discovered on Earth itself.[n 30] Radon is the most recently identified nonmetal, with its detection occurring at the end of the 19th century.[172]

The isolation of nonmetallic elements depended on a range of chemical and physical techniques. These methods encompassed

displacement reactions, combustion
, and controlled heating processes. While some nonmetals were naturally occurring as free elements, others required intricate extraction procedures:

Origin and use of the term

Traité élémentaire de chimie (1789),[216] listing the elemental gases oxygen, hydrogen and nitrogen (and erroneously including light and caloric), and the nonmetallic substances sulfur, phosphorus, and carbon, and including the chloride, fluoride and borate
ions

Although a distinction had existed between metals and other mineral substances since ancient times, it was only towards the end of the 18th century that a basic classification of chemical elements as either metallic or nonmetallic substances began to emerge. It would take another nine decades before the term "nonmetal" was widely adopted.

Around the year 340 BCE, in Book III of his treatise

ruddle, sulfur, cinnabar, and other substances that he referred to as "stones which cannot be melted".[217]

Up until the Middle Ages the classification of minerals remained largely unchanged, albeit with varying terminology. In the fourteenth century, an English alchemist named Richardus Anglicus expanded upon the classification of minerals in his work Correctorium Alchemiae. In this text, he proposed the existence of two primary types of minerals. The first category, which he referred to as "major minerals", included well-known metals such as gold, silver, copper, tin, lead, and iron. On the other hand, the second category, labeled as "minor minerals", encompassed substances like salts, atramenta (iron sulfate), alums, vitriol, arsenic, orpiment, sulfur, and similar substances that were not metallic bodies.[218]

The term "nonmetallic" has historical origins dating back to at least the 16th century. In a 1566 medical treatise, the French physician Loys de L'Aunay discussed the distinct properties exhibited by substances derived from plant sources. In his writings, he made a significant comparison between the characteristics of materials originating from what he referred to as metallic soils and non-metallic soils.[219]

Later, the French chemist Nicolas Lémery discussed metallic minerals and nonmetallic minerals in his work Universal Treatise on Simple Drugs, Arranged Alphabetically published in 1699. In his writings, he contemplated whether the substance "cadmia" belonged to either the first category, akin to cobaltum (cobaltite), or the second category, exemplified by what was then known as calamine—a mixed ore containing zinc carbonate and silicate.[220]

The pivotal moment in the systematic classification of chemical elements, distinguishing between metallic and nonmetallic substances, came in 1789 with the groundbreaking work of

Traité élémentaire de chimie. In this work he categorized elements into distinct groups, including gases, metallic substances, nonmetallic substances, and earths (heat-resistant oxides).[222] Lavoisier's work gained widespread recognition and was republished in twenty-three editions across six languages within its first seventeen years, significantly advancing the understanding of chemistry in Europe and America.[223]

The eventual and widespread adoption of the term "nonmetal" followed a complex and lengthy developmental process that spanned nearly nine decades. In 1811, the Swedish chemist Berzelius introduced the term "metalloids"[224] to describe nonmetallic elements, noting their ability to form negatively charged ions with oxygen in aqueous solutions.[225][226] While Berzelius' terminology gained significant acceptance,[227] it later faced criticism from some who found it counterintuitive,[226] misapplied,[228] or even invalid.[229][230] In 1864, reports indicated that the term "metalloids" was still endorsed by leading authorities,[231] but there were reservations about its appropriateness. The idea of designating elements like arsenic as metalloids had been considered.[231] By as early as 1866, some authors began preferring the term "nonmetal" over "metalloid" to describe nonmetallic elements.[232] In 1875, Kemshead[233] observed that elements were categorized into two groups: non-metals (or metalloids) and metals. He noted that the term "non-metal", despite its compound nature, was more precise and had become universally accepted as the nomenclature of choice.

Suggested distinguishing criteria

Some single properties used to distinguish metals and nonmetals
Physical
  • Critical temperature[242]
  • 3D
    Thermal conductivity[250]
Acid-base character
of oxides[252]
  • Sulfate formation[61]
  • Oxide solubility in acids[253]
  • Electron-related

    In 1809, the British chemist and inventor Humphry Davy made a groundbreaking discovery that reshaped the understanding of metals and nonmetals.[255] His isolation of sodium and potassium represented a significant departure from the conventional method of classifying metals solely based on their ponderousness or high densities.[256] Sodium and potassium, on the contrary, floated on water.[n 34] Nevertheless, their classification as metals was firmly established by their distinct chemical properties.[259]

    As early as 1811, attempts were initiated to enhance the differentiation between metals and nonmetals by examining a range of properties, including physical, chemical, and electron-related characteristics. The table provided here outlines 22 such properties, sorted by type and year of mention.

    One of the most commonly recognized properties used in this context is the effect of heating on electrical conductivity. As temperature rises, the conductivity of metals decreases while that of nonmetals increases.[246] However, plutonium, carbon, arsenic, and antimony defy the norm. When plutonium (a metal) is heated within a temperature range of −175 to +125 °C its conductivity increases.[260] Similarly, despite its common classification as a nonmetal, when carbon (as graphite) is heated it experiences a decrease in electrical conductivity.[261] Arsenic and antimony, which are occasionally classified as nonmetals, show behavior similar to carbon, highlighting the complexity of the distinction between metals and nonmetals.[262]

    Kneen and colleagues[263] proposed that the classification of nonmetals can be achieved by establishing a single criterion for metallicity. They acknowledged that various plausible classifications exist and emphasized that while these classifications may differ to some extent, they would generally agree on the categorization of nonmetals.

    However, Emsley[264] pointed out the complexity of this task, asserting that no single property alone can unequivocally assign elements to either the metal or nonmetal category. Furthermore, Jones[265] emphasized that classification systems typically rely on more than two attributes to define distinct types.

    Characteristics of
    nonmetals according to Johnson[266]
    Tier Elements
    Gaseous H, N, O, F, Cl, noble gases
    Other insulator S, Br
    Semiconductor B, Si, Ge, P, Se, Te, I
    Soft and crumbly
    conductor; acidic oxide
    C, As, Sb[n 35]
    All other elements are metals

    An approach to distinguishing between metallic and nonmetallic properties was suggested by Johnson,

    physical properties
    , while acknowledging the potential need for other properties in certain ambiguous cases. His observations highlighted several key distinctions:

    1. Physical state—Elements that exist as gases or are nonconductors are typically classified as nonmetals.
    2. Solid nonmetals—Solid nonmetals exhibit characteristics such as hardness and brittleness (B, Si, Ge) or softness and crumbliness, setting them apart from metals that are generally malleable and ductile.
    3. Chemical behavior—Nonmetal oxides tend to be acidic, providing another useful criterion for identifying nonmetals.

    Metals and nonmetals sorted by
    density and electronegativity (EN)[n 36]
    EN
    Density < 1.9 ≥ 1.9
    < 7 gm/cm3 Groups 1 and 2
    Sc, Y, La
    Ce, Pr, Eu, Yb
    Ti, Zr, V; Al, Ga
    Noble gases
    F, Cl, Br, I
    H, C, N, P, O, S, Se
    B, Si, Ge, As, Sb, Te^
    > 7 gm/cm3 Nd, Pm, Sm, Gd, Tb, Dy
    Ho, Er, Tm, Lu; Ac–Es;
    Hf, Nb, Ta; Cr, Mn, Fe,
    Co, Zn, Cd, In, Tl, Pb
    Ni, Mo, W, Tc, Re,
    Platinum group metals,
    Coinage metals
    , Hg; Sn,
    Bi, Po, At
    ^ italicized elements are commonly recognized by some authors as metalloids

    Several authors[273] have noted that, in general, and among other properties, nonmetals have low densities and high electronegativity, which is consistent with the data presented in the table. Nonmetallic elements are predominantly located in the top right quadrant of this table, where density is low and electronegativity values are relatively high. In contrast, the other three quadrants are primarily occupied by metals. Goldwhite and Spielman[274] added that, "... lighter elements tend to be more electronegative than heavier ones." The average electronegativity for the elements in the table with densities less than 7 gm/cm3 (metals and nonmetals) is 2.09 compared to 1.68 for the metals having densities of more than 7 gm/cm3.

    While some authors choose to further subdivide elements into metals, metalloids, and nonmetals, Oderberg[275] disagrees with this approach. He argues that according to the principles of categorization, anything not classified as a metal should be considered a nonmetal.

    Development of types

    A side profile set in stone of a distinguished French gentleman
    Gaspard Alphonse Dupasquier (1793–1848), French doctor, pharmacist and chemist as depicted in the Monument aux Grands Hommes de la Martinière [fr] in Lyon, France. In 1844 he put forward a basic taxonomy of nonmetals.

    In 1844, Alphonse Dupasquier [fr], a French doctor, pharmacist, and chemist,[276] established a basic taxonomy of nonmetals to aid in the study of these elements. He wrote:[277]

    They will be divided into four groups or sections, as in the following:
    Organogens O, N, H, C
    Sulphuroids S, Se, P
    Chloroides F, Cl, Br, I
    Boroids B, Si.

    Dupasquier's fourfold classification has echoes in the modern types of nonmetals. The organogens and sulphuroids are akin to the unclassified nonmetals. The chloroide nonmetals were later recognized independently as halogens.[278] The boroid nonmetals eventually evolved into the metalloids, with this classification beginning as early as 1864.[231] The noble gases were also identified as a distinct group among the nonmetals, dating back to as early as 1900.[279]

    Comparison of selected properties

    The two tables in this section list some physical and chemical properties of metals[n 37] and those of the three to four types of nonmetals, based on the most stable forms of the elements in ambient conditions.

    The aim is to show that most properties display a left-to-right progression in metallic-to-nonmetallic character or average values.[280][281] Some overlapping of boundaries can occur as outlier elements of each type exhibit less-distinct, hybrid-like, or atypical properties.[282][n 38] These overlaps or transitional points, along with horizontal, diagonal, and vertical relationships between the elements, form part of the "great deal of information" summarized by the periodic table.[284]

    The dashed lines around the columns for metalloids signify that the treatment of these elements as a distinct type can vary depending on the author, or classification scheme in use.

    Physical

    Physical properties are presented in loose order of ease of their determination.

    Element type
    Property Metals Metalloids Unc. nonmetals Halogen nonmetals Noble gases
    Form and density[285] solid solid solid or gas solid, liquid or gas gas
    often high density such as Fe, Pb, W low to moderately high density low density low density low density
    some
    light metals
    including Be, Mg, Al
    all lighter than Fe H, N lighter than air[286] He, Ne lighter than air[287]
    Appearance lustrous[23] lustrous[288]
    • ◇ lustrous: C, P, Se[289]
    • ◇ colorless: H, N, O[290]
    • ◇ colored: S[291]
    • ◇ colored: F, Cl, Br[292]
    • ◇ lustrous: I[9]
    colorless[293]
    Elasticity mostly malleable and ductile[23] (Hg is liquid) brittle[288] C, black P, S, Se brittle[n 39] iodine is brittle[296] not applicable
    Electrical conductivity good[n 40]
    • ◇ moderate: B, Si, Ge, Te
    • ◇ good: As, Sb[n 41]
    • ◇ poor: H, N, O, S
    • ◇ moderate: P, Se
    • ◇ good: C[n 42]
    • ◇ poor: F, Cl, Br
    • ◇ moderate: I[n 43]
    poor[n 44]
    Electronic structure[299] metallic (Bi is a semimetal) semimetal (As, Sb) or semiconductor
    • ◇ semimetal: C
    • ◇ semiconductor: P, Se
    • ◇ insulator: H, N, O, S
    semiconductor (I) or insulator insulator

    Chemical

    Chemical properties
    start with general characteristics and proceed to more specific details.

    Element type
    Property Metals Metalloids Unc. nonmetals Halogen nonmetals Noble gases
    General chemical behavior
    weakly nonmetallic[n 45] moderately nonmetallic[281] strongly nonmetallic[302]
    • ◇ inert to nonmetallic[303]
    • ◇ Rn shows some cationic behavior[304]
    Oxides basic; some
    amphoteric or acidic[305]
    amphoteric or weakly acidic[306][n 46] acidic[n 47] or neutral[n 48] acidic[n 49] metastable XeO3 is acidic;[311] stable XeO4 strongly so[312]
    few glass formers[n 50] all glass formers[314] some glass formers[n 51] no glass formers reported no glass formers reported
    ionic, polymeric, layer, chain, and molecular structures[316] polymeric in structure[317]
    • ◇ mostly molecular[317]
    • ◇ C, P, S, Se have at least one polymeric form
    • ◇ mostly molecular
    • ◇ iodine has at least one polymeric form, I2O5[318]
    • ◇ mostly molecular
    • XeO2 is polymeric[319]
    Compounds with metals alloys
    intermetallic compounds[320]
    tend to form alloys or intermetallic compounds[321]
    • ◇ salt-like to covalent: H†, C, N, P, S, Se[322]
    • ◇ mainly ionic: O[323]
    mainly ionic[145] simple compounds in ambient conditions not known[n 52]
    Ionization energy (kJ mol−1)[325] ‡ low to high moderate moderate to high high high to very high
    376 to 1,007 762 to 947 941 to 1,402 1,008 to 1,681 1,037 to 2,372
    average 643 average 833 average 1,152 average 1,270 average 1,589
    Electronegativity (Pauling)[n 53][74] ‡ low to high moderate moderate to high high high (Rn) to very high
    0.7 to 2.54 1.9 to 2.18 2.19 to 3.44 2.66 to 3.98 ca. 2.43 to 4.7
    average 1.5 average 2.05 average 2.65 average 3.19 average 3.3

    † Hydrogen can also form alloy-like hydrides[167]
    ‡ The labels low, moderate, high, and very high are arbitrarily based on the value spans listed in the table

    See also

    Notes

    1. ^ By weight, O/Si/H comprise 83.9% of the crust; N/O, 99% of the atmosphere; O/H, 99.4% of the hydrosphere; and O/C/H/N, 96% of the biomass.
    2. ^ While the oxides of most metals are basic, an appreciable number are either amphoteric or acidic.[5]
    3. ^ Metallic or nonmetallic character has usually been taken to be indicated by one property rather than two or more
    4. ^ A 2D semiconductor with a metallic appearance, showing evidence of delocalized electrons[17]
    5. ^ Solid iodine has a silvery metallic appearance under white light at room temperature.[21] It sublimes at ordinary and higher temperatures, passing from solid to gas; its vapours are violet-colored.[22]
    6. ^ The solid nonmetals have electrical conductivity values ranging from 10−18 S•cm−1 for sulfur[25] to 3 × 104 in graphite[26] or 3.9 × 104 for arsenic;[27] cf. 0.69 × 104 for manganese to 63 × 104 for silver, both metals.[25] The conductivity of graphite (a nonmetal) and arsenic (a metalloid nonmetal) exceeds that of manganese. Such overlaps show that it can be difficult to draw a clear line between metals and nonmetals.
    7. ^ The absorbed light may be converted to heat or re-emitted in all directions so that the emission spectrum is thousands of times weaker than the incident light radiation[46]
    8. ^ Thermal conductivity values for metals range from 6.3 W m−1 K−1 for neptunium to 429 for silver; cf. antimony 24.3, arsenic 50, and carbon 2000.[25] Electrical conductivity values of metals range from 0.69 S•cm−1 × 104 for manganese to 63 × 104 for silver; cf. carbon 3 × 104,[26] arsenic 3.9 × 104 and antimony 2.3 × 104.[25]
    9. ^ These elements being semiconductors.[51]
    10. ^ While CO and NO are commonly referred to as being neutral, CO is a slightly acidic oxide, reacting with bases to produce formates (CO + OH → HCOO);[71] and in water, NO reacts with oxygen to form nitrous acid HNO2 (4NO + O2 + 2H2O → 4HNO2).[72]
    11. ^ F−I: 3.98 + 3.16 + 2.96 + 2.66 = 12.76/4 = 3.19
    12. ^ B−Te: 2.04 + 1.9 + 2.01 + 2.18 + 2.05 + 2.1 = 12.28/6 = 2.04
    13. ^ These elements are hydrogen and helium in the s-block; boron to neon in the p-block; scandium to zinc in the d-block; and lanthanum to ytterbium in the f-block
    14. s-block) that is sometimes known as secondary periodicity: elements in even periods have smaller atomic radii and prefer to lose fewer electrons, while elements in odd periods (except the first) differ in the opposite direction. Many properties in the p-block then show a zigzag rather than a smooth trend along the group. For example, phosphorus and antimony in odd periods of group 15 readily reach the +5 oxidation state, whereas nitrogen, arsenic, and bismuth in even periods prefer to stay at +3.[92]
    15. ^ For example, the conductivity of graphite is 3 × 104 S•cm−1[97] whereas that of manganese is 6.9 × 103 S•cm−1[98]
    16. ^ A homopolyatomic cation consists of two or more atoms of the same element bonded together and carrying a positive charge, for example, N5+, O2+ and Cl4+, Such ions are further known for C, P, Sb, S, Se, Te, Br, I and Xe[99].
    17. ^ The seven nonmetals marked with single or double daggers each have a lackluster appearance and discrete molecular structures, but for I which has a metallic appearance under white light.[100] The remaining reactive nonmetallic elements have giant covalent structures, but for H which is a diatomic gas.[101]

      The single dagger nonmetals N, S and iodine are somewhat hobbled as to the strength of their nonmetallic character:

      • While N has a high electronegativity, it is a reluctant anion former,[102] and a pedestrian oxidizing agent unless combined with a more active nonmetal like O or F.[103]
      • S reacts in the cold with alkalic and post-transition metals, and Cu, Ag and Hg,[104] but otherwise has low values of ionization energy, electron affinity, and electronegativity compared to the averages of the others; it is regarded as being not a particularly good oxidizing agent.[105]
      • Iodine is sufficiently corrosive to cause lesions resembling thermal burns, if handled without suitable protection,
        reduction potentials are F +2.87 V; Cl +1.36; Br +1.09; I +0.54. Here Fe3+ + e → Fe3+ +0.77.[109] Thus F2, Cl2 and Br2 will oxidize Fe2+ to Fe3+ but Fe3+ will oxidize I to I2. Iodine has previously been referred to as a moderately strong oxidizing agent.[110]
    18. ^ The quote marks are not found in the source; they are used here to make it clear that the source employs the word non-metals as a formal term for the subset of chemical elements in question, rather than applying to nonmetals generally
    19. ^ Varying configurations of these nonmetals have been referred to as, for example, basic nonmetals,[117] bioelements,[118] central nonmetals,[119] CHNOPS,[120] essential elements,[121] "non-metals",[122][n 18] orphan nonmetals,[123] or redox nonmetals[124]

      The descriptive phrase unclassified nonmetals is used here for convenience.

    20. ^ Tshitoyan et al. (2019) conducted a machine-based analysis of the proximity of names of the elements based on 3.3 million abstracts published between 1922 and 2018 in more than 1,000 journals. The resulting map shows that "chemically similar elements are seen to cluster together and the overall distribution exhibits a topology reminiscent of the periodic table itself".[126]
    21. ^ Such boundary fuzziness and overlap often occur in classification schemes.[127]
    22. ^ Jones takes a philosophical or pragmatic view to these questions. He writes: "Though classification is an essential feature of all branches of science, there are always hard cases at the boundaries. The boundary of a class is rarely sharp ... Scientists should not lose sleep over the hard cases. As long as a classification system is beneficial to economy of description, to structuring knowledge and to our understanding, and hard cases constitute a small minority, then keep it. If the system becomes less than useful, then scrap it and replace it with a system based on different shared characteristics".[127]
    23. properties of metals, metalloids, and nonmetals, see Rudakiya & Patel (2021), p. 36
    24. ^ Xe is expected to be metallic at the pressures encountered in the Earth's core[139]
    25. ^ Metal oxides are usually ionic.[146] On the other hand, oxides of metals with high oxidation states are usually either polymeric or covalent.[147] A polymeric oxide has a linked structure composed of multiple repeating units.[148]
    26. ^ Sulfur, an insulator, and selenium, a semiconductor are each photoconductors—their electrical conductivities increase by up to six orders of magnitude when exposed to light[158]
    27. ^ For example, Wulfsberg divides the nonmetals, including B, Si, Ge, As, Sb, Te, Xe, into very electronegative nonmetals (Pauling electronegativity over 2.8) and electronegative nonmetals (1.9 to 2.8). This results in N and O being very electronegative nonmetals, along with the halogens; and H, C, P, S and Se being electronegative nonmetals. Se is further recognized as a semiconducting metalloid.[163]
    28. ^ Approximate composition by weight
    29. ^ Ordinary matter – including the stars, planets, and all living creatures – constitutes less than 5% of the universe. The rest – dark energy and dark matter – is as yet poorly understood.[191]
    30. Dumas had stated, behaved as a metal".[208]
    31. Berzelius, who discovered selenium, thought it had the properties of a metal, combined with the properties of sulfur[213]
    32. ^ Not to be confused with the modern usage of fossil to refer to the preserved remains, impression, or trace of any once-living thing
    33. ^ The Goldhammer-Herzfeld ratio is roughly equal to the cube of the atomic radius divided by the molar volume.[237] More specifically, it is the ratio of the force holding an individual atom's outer electrons in place with the forces on the same electrons from interactions between the atoms in the solid or liquid element. When the interatomic forces are greater than, or equal to, the atomic force, outer electron itinerancy is indicated and metallic behaviour is predicted. Otherwise nonmetallic behaviour is anticipated.
    34. ^ When Davy isolated sodium and potassium their low densities challenged the conventional wisdom that metals were ponderous substances. It was thus proposed to refer to them as metalloids, meaning "resembling metals in form or appearance".[257] This suggestion was ignored; the two new elements were admitted to the metal club in cognizance of their physical properties (opacity, luster, malleability, conductivity) and "their qualities of chemical combination". Hare[258] observed that the line of demarcation between metals and nonmetals had been "annihilated" by the discovery of alkaline metals having a density less than that of water:
      "Peculiar brilliance and opacity were in the next place appealed to as a means of discrimination; and likewise that superiority in the power of conducting heat and electricity ... Yet so difficult has it been to draw the line between metallic…and non-metallic ... that bodies which are by some authors placed in one class, are by others included in the other. Thus selenium, silicon, and zirconion [sic] have by some chemists been comprised among the metals, by others among non-metallic bodies." ...
    35. amphoteric its very weak acid properties dominate over those of a very weak base[267]
    36. ^ (a) Up to element 99 (Es), with the values taken from Aylward and Findlay.[268]
      (b) Weighable amounts of the extremely radioactive elements At (element 85), Fr (87), and elements with an atomic number higher than Es (99), have not been prepared.[269]
      (c) The density values used for At and Fr are theoretical estimates.[270]
      (d) A survey of definitions of the term "heavy metal" reported density criteria ranging from above 3.5 g/cm3 to above 7 g/cm3.[271]
      (e) Vernon specified a minimum electronegativity of 1.9 for the metalloids, on the revised Pauling scale[9]
      (f) Electronegativity values for the noble gases are from Rahm, Zeng and Hoffmann[272]
    37. ^ Metals are included for reference
    38. ^ A similar phenomenon applies more generally to certain Groups of the periodic table where, for example, the noble gases in Group 18 act as bridge between the nonmetals of the p-block and the metals of the s-block (Groups 1 and 2)[283]
    39. ^ All four have less stable non-brittle forms:[294] carbon as exfoliated (expanded) graphite,[52][295] and as carbon nanotube wire;[54] phosphorus as white phosphorus (soft as wax, pliable and can be cut with a knife, at room temperature);[55] sulfur as plastic sulfur;[56] and selenium as selenium wires[57]
    40. ^ Metals have electrical conductivity values of from 6.9×103 S•cm−1 for manganese to 6.3×105 for silver[297]
    41. ^ Metalloids have electrical conductivity values of from 1.5×10−6 S•cm−1 for boron to 3.9×104 for arsenic[298]
    42. ^ Unclassified nonmetals have electrical conductivity values of from ca. 1×10−18 S•cm−1 for the elemental gases to 3×104 in graphite[97]
    43. ^ The halogen nonmetals have electrical conductivity values of from ca. 1×10−18 S•cm−1 for F and Cl to 1.7×10−8 S•cm−1 for iodine[97][142]
    44. ^ The elemental gases have electrical conductivity values of ca. 1×10−18 S•cm−1[97]
    45. ^ Metalloids always give "compounds less acidic in character than the corresponding compounds of the [typical] nonmetals"[288]
    46. ^ Arsenic trioxide reacts with sulfur trioxide, forming arsenic "sulfate" As2(SO4)3[307]
    47. ^ NO
      2
      , N
      2
      O
      5
      , SO
      3
      , SeO
      3
      are strongly acidic[308]
    48. anhydrides of formic and hyponitrous acid, respectively viz. CO + H2O → H2CO2 (HCOOH, formic acid); N2O + H2O → H2N2O2 (hyponitrous acid)"[309]
    49. ^ ClO
      2
      , Cl
      2
      O
      7
      , I
      2
      O
      5
      are strongly acidic[310]
    50. ^ Metals that form glasses are: V; Mo, W; Al, In, Tl; Sn, Pb; Bi[313]
    51. ^ Unclassified nonmetals that form glasses are P, S, Se;[313] CO2 forms a glass at 40 GPa[315]
    52. ^ Disodium helide (Na2He) is a compound of helium and sodium that is stable at high pressures above 113 GPa. Argon forms an alloy with nickel, at 140 GPa and close to 1,500 K however at this pressure argon is no longer a noble gas[324]
    53. ^ Values for the noble gases are from Rahm, Zeng and Hoffmann[272]

    References

    Citations

    1. ^ Restrepo et al. 2006, p. 411; Thornton & Burdette 2010, p. 86; Hermann, Hoffmann & Ashcroft 2013, pp. 11604‒1‒11604‒5
    2. ^ Parkes & Mellor 1943, p. 740
    3. ^ Pascoe 1982, p. 3
    4. ^ Glinka 1973, p. 56; Oxtoby, Gillis & Butler 2015, p. I.23
    5. ^ Liu, Yang & Zheng 2022, p. 31
    6. ^ Godovikov & Nenasheva 2020, p. 4; Sanderson 1957, p. 229; Morely & Muir 1892, p. 241
    7. ^ a b Larrañaga, Lewis & Lewis 2016, p. 988
    8. ^ Steudel 2020, p. 43: Steudel's monograph is an updated translation of the fifth German edition of 2013, incorporating the literature up to Spring 2019.
    9. ^ a b c d e f Vernon 2013
    10. ^ Vernon 2020, p. 220; Rochow 1966, p. 4
    11. ^ IUPAC Periodic Table of the Elements
    12. ^ Johnson 2007, p. 13
    13. ^ Bodner & Pardue 1993, p. 354; Cherim 1971, p. 98
    14. ^ Chen 2021, p. 33; Burrows et al. 2021, p. 1242; Vallabhajosula 2023, p. 214
    15. ^ Vernon 2013, p. 1204
    16. ^ Nefedov et al. 1968, p. 87
    17. ^ Steudel 2020, p. 601
    18. ^ Vasáros & Berei 1985, p. 109; Seaborg 1948, p. 368; Bladel 1949, pp. 51–52; Kleinberg 1950, p. 32; Fearnside, Jones & Shaw 1954, p. 102; Encyclopedia Britannica 1956, vol. 6, p. 823; Furse & Rendle 1975, p. 82; Siekierski & Burgess 2002, pp. 65, 122; Restrepo et al. 2006, p. 411; Thornton & Burdette 2010, p. 86
    19. ^ Hermann, Hoffmann & Ashcroft 2013, pp. 11604‒1‒11604‒5
    20. ^ Mewes et al. 2019; Smits et al. 2020; Florez et al. 2022
    21. ^ Koenig 1962, p. 108
    22. ^ Tidy 1887, pp. 107–108
    23. ^ a b c d Kneen, Rogers & Simpson 1972, pp. 261–264
    24. ^ Phillips 1973, p. 7
    25. ^ a b c d e Aylward & Findlay 2008, pp. 6–12
    26. ^ a b Jenkins & Kawamura 1976, p. 88
    27. ^ Carapella 1968, p. 30
    28. ^ Zumdahl & DeCoste 2010, pp. 455, 456, 469, A40; Earl & Wilford 2021, p. 3-24
    29. ^ Still 2016, p. 120
    30. ^ Wiberg 2001, pp. 780
    31. ^ Wiberg 2001, pp. 824, 785
    32. ^ Earl & Wilford 2021, p. 3-24
    33. ^ Siekierski & Burgess 2002, p. 86
    34. ^ Charlier, Gonze & Michenaud 1994
    35. ^ Taniguchi et al. 1984, p. 867: "... black phosphorus ... [is] characterized by the wide valence bands with rather delocalized nature."; Morita 1986, p. 230; Carmalt & Norman 1998, p. 7: "Phosphorus ... should therefore be expected to have some metalloid properties."; Du et al. 2010. Interlayer interactions in black phosphorus, which are attributed to van der Waals-Keesom forces, are thought to contribute to the smaller band gap of the bulk material (calculated 0.19 eV; observed 0.3 eV) as opposed to the larger band gap of a single layer (calculated ~0.75 eV).
    36. ^ Wiberg 2001, pp. 742
    37. ^ Evans 1966, pp. 124–25
    38. ^ Wiberg 2001, pp. 758
    39. ^ Stuke 1974, p. 178; Donohue 1982, pp. 386–87; Cotton et al. 1999, p. 501
    40. ^ Steudel 2000, p. 601: "... Considerable orbital overlap can be expected. Apparently, intermolecular multicenter bonds exist in crystalline iodine that extend throughout the layer and lead to the delocalization of electrons akin to that in metals. This explains certain physical properties of iodine: the dark color, the luster and a weak electric conductivity, which is 3400 times stronger within the layers then perpendicular to them. Crystalline iodine is thus a two-dimensional semiconductor."; Segal 1989, p. 481: "Iodine exhibits some metallic properties ..."
    41. ^ Cahn & Haasen 1996, p. 4; Boreskov 2003, p. 44
    42. ^ DeKock & Gray 1989, pp. 423, 426—427
    43. ^ Boreskov 2003, p. 45
    44. ^ Wiberg 2001, p. 416; Wiberg is here referring to iodine.
    45. ^ Elliot 1929, p. 629
    46. ^ Fox 2010, p. 31
    47. ^ Wibaut 1951, p. 33: "Many substances ...are colourless and therefore show no selective absorption in the visible part of the spectrum."
    48. ^ Kneen, Rogers & Simpson 1972, pp. 85–86, 237
    49. ^ Salinas 2019, p. 379
    50. ^ Yang 2004, p. 9
    51. ^ Wiberg 2001, pp. 416, 574, 681, 824, 895, 930; Siekierski & Burgess 2002, p. 129
    52. ^ a b Chung 1987
    53. ^ Godfrin & Lauter 1995
    54. ^ a b Janas, Cabrero-Vilatela & Bulmer 2013
    55. ^ a b Faraday 1853, p. 42; Holderness & Berry 1979, p. 255
    56. ^ a b Partington 1944, p. 405
    57. ^ a b c Regnault 1853, p. 208
    58. ^ Edwards 2000, pp. 100, 102–103; Herzfeld 1927, pp. 701–705
    59. ^ Kneen, Rogers & Simpson 1972, pp. 263‒264
    60. ^ Langley & Hattori 2014, p. 214
    61. ^ a b Abbott 1966, p. 18
    62. ^ Brown et al. 2014, p. 237
    63. ^ Barton 2021, p. 200
    64. ^ Borg & Dienes 1992, p. 26
    65. ^ Wiberg 2001, p. 796
    66. ^ Shang et al. 2021
    67. ^ Tang et al. 2021
    68. ^ Steudel 2020, passim; Carrasco et al. 2023; Shanabrook, Lannin & Hisatsune 1981, pp. 130–133
    69. ^ Eagleson 1994, 1169
    70. ^ Moody 1991, p. 365
    71. ^ House 2013, p. 427
    72. ^ Lewis & Deen 1994, p. 568
    73. ^ Yoder, Suydam & Snavely 1975, p. 58
    74. ^ a b c Aylward & Findlay 2008, p. 126
    75. ^ Young et al. 2018, p. 753
    76. ^ Brown et al. 2014, p. 227
    77. ^ Siekierski & Burgess 2002, pp. 21, 133, 177
    78. ^ Moore 2016; Burford, Passmore & Sanders 1989, p. 54
    79. ^ King & Caldwell 1954, p. 17; Brady & Senese 2009, p. 69
    80. ^ Chemical Abstracts Service 2021
    81. ^ Emsley 2011, pp. 81
    82. ^ Cockell 2019, p. 210
    83. ^ Scott 2014, p. 3
    84. ^ Emsley 2011, p. 184
    85. ^ Lee 1996, p. 240
    86. ^ Greenwood & Earnshaw 2002, p. 43
    87. ^ Cressey 2010
    88. ^ Siekierski & Burgess 2002, pp. 24–25
    89. ^ Siekierski & Burgess 2002, p. 23
    90. ^ Petruševski & Cvetković 2018; Grochala 2018
    91. ^ Kneen, Rogers & Simpson 1972, pp. 226, 360; Siekierski & Burgess 2002, pp. 52, 101, 111, 124, 194
    92. ^ Scerri 2020, pp. 407–420
    93. ^ Greenwood & Earnshaw 2002, pp. 27, 1232, 1234
    94. ^ Cox 2004, p. 146
    95. ^ Dorsey 2023, pp. 12–13
    96. ^ Humphrey 1908
    97. ^ a b c d Bogoroditskii & Pasynkov 1967, p. 77; Jenkins & Kawamura 1976, p. 88
    98. ^ Desai, James & Ho 1984, p. 1160
    99. ^ Engesser & Krossing 2013, p. 947
    100. ^ a b Vernon 2013, p. 1706
    101. ^ Wiberg 2001, passim
    102. ^ Vernon 2020, p. 220
    103. ^ Atkins & Overton 2010, pp. 377, 389
    104. ^ Moody 1991, p. 391
    105. ^ Rodgers 2012, p. 504; Wulfsberg 2000, p. 726
    106. ^ Stellman 1998, chapter 104–211
    107. ^ Nakao 1992, p. 426–427
    108. ^ Hill & Holman 2000, p. 196
    109. ^ Wiberg 2001, pp. 1761–1762
    110. ^ Young 2006, p. 1285
    111. ^ Encyclopaedia Britannica 2021
    112. ^ Royal Society of Chemistry 2021
    113. ^ a b Matson & Orbaek 2013, p. 203
    114. ^ Chambers & Holliday 1982, pp. 273–274; Bohlmann 1992, p. 213; Jentzsch 2015, p. 247
    115. ^ Kernion 2019, p. 191; Cao et al. 2021, pp. 20–21; Hussain et al. 2023
    116. ^ Vassilakis, Kalemos & Mavridis 2014, p. 1; Hanley & Koga 2018, p. 24; Kaiho 2017, ch. 2, p. 1
    117. ^ Williams 2007, pp. 1550–1561: H, C, N, P, O, S
    118. ^ Wächtershäuser 2014, p. 5: H, C, N, P, O, S, Se
    119. ^ Hengeveld & Fedonkin, pp. 181–226: C, N, P, O, S
    120. ^ Wakeman 1899, p. 562
    121. ^ Fraps 1913, p. 11: H, C, Si, N, P, O, S, Cl
    122. ^ Parameswaran at al. 2020, p. 210: H, C, N, P, O, S, Se
    123. ^ Knight 2002, p. 148: H, C, N, P, O, S, Se
    124. ^ Fraústo da Silva & Williams 2001, p. 500: H, C, N, O, S, Se
    125. ^ a b Bailar et al. 1989, p. 742
    126. ^ Tshitoyan et al. 2019, pp. 95–98
    127. ^ a b Jones 2010, pp. 169–71
    128. ^ Stein 1983, p. 165
    129. ^ Russell & Lee 2005, p. 419
    130. ^ Goodrich 1844, p. 264; The Chemical News 1897, p. 189; Hampel & Hawley 1976, pp. 174, 191; Lewis 1993, p. 835; Hérold 2006, pp. 149–50
    131. ^ Tyler 1948, p. 105; Reilly 2002, pp. 5–6
    132. ^ Jolly 1966, p. 20
    133. ^ Clugston & Flemming 2000, pp. 100–101, 104–105, 302
    134. ^ Maosheng 2020, p. 962
    135. ^ Mazej 2020
    136. ^ Cox 2000, pp. 258–259; Möller 2003, p. 173; Trenberth & Smith 2005, p. 864
    137. ^ Emsley 2011, p. 220
    138. ^ Emsley 2011, p. 440
    139. ^ Lee & Steinle-Neumann 2006, p. 1
    140. ^ Zhu et al. 2014, pp. 644–648
    141. ^ Wiberg 2001, pp. 4022
    142. ^ a b Greenwood & Earnshaw 2002, p. 804
    143. ^ Rudolph 1973, p. 133: "Oxygen and the halogens in particular ... are therefore strong oxidizing agents."
    144. ^ Daniel & Rapp 1976, p. 55
    145. ^ a b Cotton et al. 1999, p. 554
    146. ^ Woodward et al. 1999, pp. 133–194
    147. ^ Phillips & Williams 1965, pp. 478–479
    148. ^ Moeller et al. 2012, p. 314
    149. ^ Lanford 1959, p. 176
    150. ^ Schmedt, Mangstl & Kraus 2012, p. 7847‒7849
    151. ^ Bailar, Moeller & Kleinberg 1965, p. 477; Mee 1964, p. 153
    152. ^ a b Cox 1997, pp. 130–132; Emsley 2011, passim
    153. ^ Hurlbut 1961, p. 132
    154. ^ Emsley 2011, p. 478
    155. ^ Greenwood & Earnshaw 2002, p. 277
    156. ^ Atkins et al. 2006, p. 320
    157. ^ Greenwood & Earnshaw 2002, p. 482; Berger 1997, p. 86
    158. ^ Moss 1952, pp. 180, 202
    159. ^ a b c d Cao et al. 2021, p. 20
    160. ^ Challoner 2014, p. 5; Government of Canada 2015; Gargaud et al. 2006, p. 447
    161. ^ Crichton 2012, p. 6; Scerri 2013; Los Alamos National Laboratory 2021
    162. ^ Vernon 2020, p. 218
    163. ^ Wulfsberg 2000, pp. 273–274, 620
    164. ^ Seese & Daub 1985, p. 65
    165. ^ MacKay, MacKay & Henderson 2002, pp. 209, 211
    166. ^ Cousins, Davidson & García-Vivó 2013, pp. 11809–11811
    167. ^ a b Cao et al. 2021, p. 4
    168. ^ Liptrot 1983, p. 161; Malone & Dolter 2008, p. 255
    169. ^ Wiberg 2001, pp. 255–257
    170. ^ Scott & Kanda 1962, p. 153
    171. ^ Taylor 1960, p. 316
    172. ^ a b c d Emsley 2011, passim
    173. ^ Crawford 1968, p. 540; Benner, Ricardo & Carrigan 2018, pp. 167–168: "The stability of the carbon-carbon bond ... has made it the first choice element to scaffold biomolecules. Hydrogen is needed for many reasons; at the very least, it terminates C-C chains. Heteroatoms (atoms that are neither carbon nor hydrogen) determine the reactivity of carbon-scaffolded biomolecules. In ... life, these are oxygen, nitrogen and, to a lesser extent, sulfur, phosphorus, selenium, and an occasional halogen."
    174. ^ Zhao, Tu & Chan 2021
    175. ^ Kosanke et al. 2012, p. 841
    176. ^ Wasewar 2021, pp. 322–323
    177. ^ Messler 2011, p. 10
    178. ^ King et al. 1994, p. 1344; Powell & Tims 1974, pp. 189–191; Cao et al. 2021, pp. 20–21
    179. ^ Vernon 2020, pp. 221–223; Rayner-Canham 2020, p. 216
    180. ^ National Center for Biotechnology Information 2021
    181. ^ Emsley 2011, p. 113
    182. ^ Greenwood & Earnshaw 2002, p. 270–271
    183. ^ Khan 2001, p. 59
    184. ^ Emsley 2011, pp. 376, 380, 640
    185. ^ Cox 1997, pp. 130; Emsley 2011, p. 393
    186. ^ Cox 1997, pp. 130; Emsley 2011, pp. 515–516, 518
    187. ^ Boyd 2011, p. 570
    188. ^ Chandra X-ray Center 2018
    189. ^ a b c Nelson 1987, p. 732
    190. ^ Fortescue 2012, pp. 56, 65
    191. ^ Ostriker & Steinhardt 2001, pp. 46‒53; Zhu 2020, p. 27
    192. ^ Cox 1997, pp. 17–19
    193. ^ Steudel 2020, p. v
    194. ^ a b c d USGS Mineral Commodity Summaries 2023; Beard et al. 2021; Bhuwalka et al. 2021, pp. 10097–10107; Allcock 2020, pp. 61–63; Burke 2020, p. 262; Imberti & Sadler 2020, p. 8; King 2019, p. 408; Gaffney & Marley 2017, p. 27; Csele 2016; Kiiski et al. 2016; Bolin 2017, p. 2-1; Harbison, Bourgeois & Johnson 2015, p. 364; Reinhardt at al. 2015; Royal Society of Chemistry; Emsley 2011, passim; Ward 2010, p. 250
    195. ^ Whitten et al. 2014, p. 133
    196. ^ Weeks ME & Leicester 1968, p. 550
    197. ^ Zhong & Nsengiyumva, p. 19
    198. ^ Angelo & Ravisankar p. 56–57
    199. ^ Greenwood & Earnshaw 2002, p. 482
    200. ^ Sultana et al. 2022
    201. ^ Haller 2006, p. 3
    202. ^ Shanks et al. 2017, pp. I2–I3
    203. ^ Emsley 2011, p. 611
    204. ^ Baja, Cascella & Borger 2022; Webb-Mack 2019
    205. ^ Rodgers 2012, p. 571
    206. ^ Greger 2023
    207. ^ Pawlicki, Scanderbeg & Starkschall 2016, p. 228
    208. ^ Labinger 2019, p. 305
    209. ^ Emsley 2011, pp. 42–43, 219–220, 263–264, 341, 441–442, 596, 609
    210. ^ Toon 2011
    211. ^ Emsley 2011, pp. 84, 128, 180–181, 247
    212. ^ Cook 1923, p. 124
    213. ^ Weeks ME & Leicester 1968, p. 309
    214. ^ Emsley 2011, pp. 113, 363, 378, 477, 514–515
    215. ^ Weeks & Leicester 1968, pp. 95, 97, 103
    216. ^ Lavoisier 1790, p. 175
    217. ^ Jordan 2016
    218. ^ Stillman 1924, p. 213
    219. ^ de L'Aunay 1566, p. 7
    220. ^ Lémery 1699, p. 118; Dejonghe 1998, p. 329
    221. ^ Strathern 2000, p. 239
    222. ^ Criswell p. 1140
    223. ^ Salzberg 1991, p. 204
    224. ^ Berzelius 1811, p. 258
    225. ^ Partington 1964, p. 168
    226. ^ a b Bache 1832, p. 250
    227. ^ Goldsmith 1982, p. 526
    228. ^ Roscoe & Schormlemmer 1894, p. 4
    229. ^ Glinka 1959, p. 76
    230. ^ Hérold 2006, pp. 149–150
    231. ^ a b c The Chemical News and Journal of Physical Science 1864
    232. ^ Oxford University Press 1989
    233. ^ Kemshead 1875, p. 13
    234. ^ Kendall 1811, pp. 298–303
    235. ^ Brande 1821, p. 5
    236. ^ Herzfeld 1927; Edwards 2000, pp. 100–03
    237. ^ Edwards & Sienko 1983, p. 693
    238. ^ Kubaschewski 1949, pp. 931–940
    239. ^ Remy 1956, p. 9
    240. ^ White 1962, p. 106: It makes a ringing sound when struck.
    241. ^ Johnson 1966, pp. 3–4
    242. ^ Horvath 1973, pp. 335–336
    243. ^ Myers 1979, p. 712
    244. ^ Rao & Ganguly 1986
    245. ^ Smith & Dwyer 1991, p. 65: The difference between melting point and boiling point.
    246. ^ a b Herman 1999, p. 702
    247. ^ Suresh & Koga 2001, pp. 5940–5944
    248. ^ a b Edwards 2010, pp. 941–965
    249. ^ Hill, Holman & Hulme 2017, p. 182: Atomic conductance is the electrical conductivity of one mole of a substance. It is equal to electrical conductivity divided by molar volume.
    250. ^ Povh & Rosin 2017, p. 131
    251. ^ Beach 1911
    252. ^ Stott 1956, pp. 100–102
    253. ^ Parish 1977, p. 178
    254. ^ Sanderson 1957, p. 229
    255. ^ Hare & Bache 1836, p. 310
    256. ^ Chambers 1743: "That which distinguishes metals from all other bodies ... is their heaviness ..."
    257. ^ Erman and Simon 1808
    258. ^ Hare 1836, p. 310
    259. ^ Edwards 2000, p. 85
    260. ^ Russell & Lee 2005, p. 466
    261. ^ Atkins et al. 2006, pp. 320–21
    262. ^ Zhigal'skii & Jones 2003, p. 66
    263. ^ Kneen, Rogers & Simpson 1972, pp. 218–219
    264. ^ Emsley 1971, p. 1
    265. ^ Jones 2010, p. 169
    266. ^ a b Johnson 1966, pp. 3–6, 15
    267. ^ Shkol'nikov 2010, p. 2127
    268. ^ Aylward & Findlay 2008, pp. 6–13; 126
    269. ^ Edelstein & Morrs 2009, p. 123
    270. ^ Arblaster JW (ed.) 2018, p. 269; Lavrukhina & Pozdnyakov 1970, p. 269
    271. ^ Duffus 2002, p. 798
    272. ^ a b Rahm, Zeng & Hoffmann 2019, p. 345
    273. ^ Hein & Arena 2011, pp. 228, 523; Timberlake 1996, pp. 88, 142; Kneen, Rogers & Simpson 1972, p. 263; Baker 1962, pp. 21, 194; Moeller 1958, pp. 11, 178
    274. ^ Goldwhite & Spielman 1984, p. 130
    275. ^ Oderberg 2007, p. 97
    276. ^ Bertomeu-Sánchez, Garcia-Belmar & Bensaude-Vincent 2002, pp. 248–249
    277. ^ Dupasquier 1844, pp. 66–67
    278. ^ Bache 1832, pp. 248–276
    279. ^ Renouf 1901, pp. 268
    280. ^ Vernon 2020, pp. 217–225
    281. ^ a b Welcher 2009, p. 3–32: "The elements change from ... metalloids, to moderately active nonmetals, to very active nonmetals, and to a noble gas."
    282. ^ Vernon 2020, pp. 224
    283. ^ MacKay, MacKay & Henderson 2002, pp. 195–196
    284. ^ Bynum, Browne & Porter 1981, p. 318
    285. ^ Tregarthen 2003, p. 10
    286. ^ Lewis 1993, pp. 28, 827
    287. ^ Lewis 1993, pp. 28, 813
    288. ^ a b c Rochow 1966, p. 4
    289. ^ Wiberg 2001, p. 780; Emsley 2011, p. 397; Rochow 1966, pp. 23, 84
    290. ^ Kneen, Rogers & Simpson 1972, pp. 321, 404, 436
    291. ^ Kneen, Rogers & Simpson 1972, p. 439
    292. ^ Kneen, Rogers & Simpson 1972, p. 465
    293. ^ Kneen, Rogers & Simpson 1972, p. 308
    294. ^ Wiberg 2001, pp. 505, 681, 781; Glinka 1958, p. 355
    295. ^ Godfrin & Lauter 1995, pp. 216‒218
    296. ^ Wiberg 2001, p. 416
    297. ^ Desai, James & Ho 1984, p. 1160; Matula 1979, p. 1260
    298. ^ Schaefer 1968, p. 76; Carapella 1968, pp. 29‒32
    299. ^ Keeler & Wothers 2013, p. 293
    300. ^ Kneen, Rogers & Simpson 1972, p. 264
    301. ^ Rayner-Canham 2018, p. 203
    302. ^ Mackin 2014, p. 80
    303. ^ Johnson 1966, pp. 105–108
    304. ^ Stein 1969, pp. 5396‒5397; Pitzer 1975, pp. 760‒761
    305. ^ Porterfield 1993, p. 336
    306. ^ Rochow 1966, p. 4; Atkins et al. 2006, pp. 8, 122–123
    307. ^ Wiberg 2001, p. 750
    308. ^ Sanderson 1967, p. 172; Mingos 2019, p. 27
    309. ^ House 2008, p. 441
    310. ^ Mingos 2019, p. 27; Sanderson 1967, p. 172
    311. ^ Wiberg 2001, p. 399
    312. ^ Kläning & Appelman 1988, p. 3760
    313. ^ a b Rao 2002, p. 22
    314. ^ Sidorov 1960, pp. 599‒603
    315. ^ McMillan 2006, p. 823
    316. ^ Wells 1984, p. 534
    317. ^ a b Puddephatt & Monaghan 1989, p. 59
    318. ^ King 1995, p. 182
    319. ^ Ritter 2011, p. 10
    320. ^ Yamaguchi & Shirai 1996, p. 3
    321. ^ Vernon 2020, p. 223
    322. ^ Vernon 2020, p. 220
    323. ^ Woodward et al. 1999, p. 134
    324. ^ Dalton 2019
    325. ^ Aylward & Findlay 2008, p. 132

    Bibliography

    External links

    • Media related to Nonmetals at Wikimedia Commons