A periodic table excerpt showing elements consistently classified as nonmetals (the noble gases plus fluorine, chlorine, bromine, iodine, oxygen, sulfur, selenium, nitrogen, phosphorus, carbon and hydrogen) and those sometimes classified as such (boron, silicon, germanium, arsenic, antimony, and tellurium). Nearby metals are aluminium, gallium, indium, thallium, tin, lead, bismuth, polonium, and astatine.

Periodic table excerpt (full table below) showing the elements considered nonmetals:
  usually to always listed as a nonmetal
  sometimes so listed

Lists of nonmetals often differ due to different definitions.[n 1]
_____
† While hydrogen (H) is typically positioned within group 1 of the periodic table, as depicted in the full table below, some representations may place it above fluorine (F) in group 17.[n 2]

‡ Theory and experimental evidence suggest astatine (At) is a metal;[4] it is treated as such in this article.

A nonmetal is a chemical element generally characterized by low density and high electronegativity (the ability of an atom in a molecule to attract electrons to itself). They range from colorless gases like hydrogen to shiny solids like carbon, as graphite. Nonmetals are often poor conductors of heat and electricity, and when in solid form tend to be brittle or crumbly due to the limited mobility of their electrons. In contrast, metals are good conductors and most can easily be flattened into sheets and drawn into wires because of the free movement of their electrons. While compounds of metals tend to be basic in nature those of nonmetals tend to be acidic.

Hydrogen and helium, which are nonmetals, together constitute about 99% of the observable ordinary matter in the universe by mass. Five nonmetallic elements—hydrogen, carbon, nitrogen, oxygen, and silicon—make up most of the Earth's crust, atmosphere, oceans and biosphere.

The distinct properties of nonmetallic elements allow for specific uses that metals cannot often achieve. For example, hydrogen, oxygen, carbon, and nitrogen, are vital to living organisms. Additionally, nonmetals are integral to various industries, such as electronics, energy storage, agriculture, and chemical production.

While the term "non-metallic" has origins dating back to 1566, there is no universally accepted definition of a nonmetal. This lack of clarity arises from elements with a marked mixture of metallic and nonmetallic properties. As a result, classifications vary, typically listing from 14 to 23 (or 24) elements as nonmetals.

Definition and applicable elements edit

 
Like carbon, arsenic (here sealed in a container to prevent tarnishing) vaporises rather than melts when heated. The vapor is lemon-yellow and smells like garlic.[5] The chemistry of arsenic is predominately nonmetallic in nature.[6]

A nonmetal is a chemical element generally characterized by low density and high electronegativity. They lack a preponderance of metallic properties, such as luster or shininess, malleability and ductility, good thermal and electrical conductivity, and the tendency to produce basic rather than acidic oxides when combined with oxygen.[7] There is no universally agreed definition of a nonmetal[8] resulting in variations among sources as to which elements are counted as such. The criteria applied depend on the properties viewed as most representative of nonmetallic or metallic character.[9][n 3]

While Steudel, in his 2020 textbook Chemistry of the Non-metals[10][n 4] listed twenty-three elements as nonmetals, any such list is open to debate and revision.[11] Among these elements, fourteen are commonly recognized as nonmetals namely hydrogen, oxygen, nitrogen, and sulfur, as well as the highly reactive halogens: fluorine, chlorine, bromine, and iodine.[11] The noble gases, including helium, neon, argon, krypton, xenon, and radon, are also definitively classified as nonmetals, as supported by Hawley's Condensed Chemical Dictionary.[11]

The classification of carbon, phosphorus, and selenium is not necessarily clear cut; while they are commonly deemed nonmetals, some sources have labelled them as metalloids.Cite error: A <ref> tag is missing the closing </ref> (see the help page). Nonetheless, their predominantly nonmetallic chemistry, characterized by weak acidity, also supports their classification as nonmetals.[12]

Of the 118 known chemical elements,[13] roughly 20% are classified as nonmetals.[14] Astatine, the fifth halogen, is often ignored on account of its rarity and intense radioactivity;[15] theory and experimental evidence suggest it is a metal.[4][n 5] The superheavy elements copernicium (element 112), flerovium (114), and oganesson (118) may turn out to be nonmetals. As of August 2023 their status has not been confirmed.[18]

General properties edit

Properties noted in this section refer to the elements in their most stable forms in ambient conditions unless otherwise mentioned

Physical edit

Variety in color and form
of some nonmetallic elements
Boron in its β-rhombohedral phase
Metallic appearance of carbon as graphite
Blue color of liquid oxygen
Pale yellow liquid fluorine in a cryogenic bath
Sulfur as a yellow powder
Liquid bromine at room temperature
Metallic appearance of iodine under white light
Liquefied xenon

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 6] The fluid nonmetals have very low densities, melting and boiling points, and are poor conductors of heat and electricity.[21] The solid elements have low densities, are brittle or crumbly with low mechanical and structural strength,[22] and are poor to good conductors of electricity.[n 7]

The diverse forms of nonmetals can be attributed to their varied internal structures and bonding arrangements. Nonmetals existing as discrete atoms like xenon, or as 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.[26] In contrast, nonmetals that form giant structures, such as chains of up to 1,000 atoms (selenium, for example),[27] sheets (carbon as graphite, for example),[28] or three-dimensional lattices (silicon, for example)[29] have higher melting and boiling points, and are all solids, as it takes more energy to overcome their stronger covalent bonds.[30] Nonmetals closer to the left side of the periodic table, or further down a column, often have some weak metallic interactions between their molecules, chains, or layers, consistent with their proximity to the metals; this occurs in boron,[31] carbon,[32] phosphorus,[33] arsenic,[34] selenium,[35] antimony,[36] tellurium[37] and iodine.[38]

Nonmetallic elements are either shiny, colored, or colorless. The shiny appearance 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.[39] The colored nonmetals (sulfur, fluorine, chlorine, bromine) absorb some colors (wavelengths) and transmit the complementary or opposite colors. In the case of chlorine, for example, Elliot writes that its "familiar yellow-green colour ... is due to a broad region of absorption in the violet and blue regions of the spectrum".[40][n 8] 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.[42]

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,[43] the electrons in nonmetals typically lack such mobility.[44] Among nonmetallic elements, good electrical and thermal conductivity is seen only in carbon, arsenic, and antimony.[n 9] Good thermal conductivity otherwise occurs only in boron, silicon, phosphorus, and germanium;[23] such conductivity is transmitted though vibrations of the crystalline lattices of these elements.[45] Moderate electrical conductivity is observed in boron, silicon, phosphorus, germanium, selenium, tellurium, and iodine.[n 10] Plasticity occurs under limited circumstances in carbon, as seen in exfoliated (expanded) graphite[47][48] and carbon nanotube wire,[49] in phosphorus as white phosphorus (soft as wax, pliable and can be cut with a knife, at room temperature),[50] in sulfur, when it exists as plastic sulfur,[51] and in selenium which can be drawn into wires from its molten state.[52]

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 free to move between atoms, and metallic properties are predicted. Otherwise nonmetallic properties are expected.[53]

Allotropes edit

 
Brownish crystals of buckminsterfullerene60), a semiconducting allotrope of carbon
Some chemistry-based typical
differences between metals and nonmetals[54]
Aspect Metals Nonmetals
Electronegativity Lower than nonmetals,
with some exceptions[55]
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[56]
In aqueous
solution
[57]
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.[58] For example, carbon, a versatile nonmetal, can manifest as graphite, diamond, and other forms, with graphite displaying relatively good electrical conductivity, while diamond is transparent and an extremely poor conductor of electricity.[59] Carbon further exists in several allotropic structures, including buckminsterfullerene,[60] and amorphous[61] and paracrystalline (mixed amorphous and crystalline)[62] variations. Iodine among the halogen nonmetals,[63] as well as the other unclassified nonmetals, and metalloids also have allotropic variations. Allotropic forms of the noble gases have not been observed.

Chemical edit

Nonmetals possess relatively high values of electronegativity[64] and tend to form acidic compounds. For example, solid nonmetals (including metalloids) react with nitric acid to produce either an acid, or an oxide with predominately acidic properties.[n 11]

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. Metals, on the other hand, follow a less rigorously predictive 18-electron rule.[67]

Furthermore, nonmetals typically exhibit higher ionization energies, electron affinities, and standard reduction potentials than metals. General, the higher these values are (including electronegativity) the more nonmetallic the element tends to be.[68]

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.[69] Consequently, there is a corresponding reduction in atomic radius[70] as the heightened nuclear charge draws the outer electrons closer to the nucleus core.[71] 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.[72]

The number of compounds formed by nonmetals is vast.[73] The first 10 places in a "top 20" table of elements most frequently encountered in 895,501,834 compounds, as listed in the Chemical Abstracts Service register for November 2, 2021, were occupied by nonmetals. Hydrogen, carbon, oxygen, and nitrogen collectively appeared in most (80%) of compounds. Silicon, a metalloid, ranked 11th. The highest-rated metal, with an occurrence frequency of 0.14%, was iron, in 12th place.[74] A few examples of nonmetal compounds are: boric acid (H
3
BO
3
), used in ceramic glazes;[75] selenocysteine (C
3
H
7
NO
2
Se
), the 21st amino acid of life;[76] phosphorus sesquisulfide (P4S3), found in strike anywhere matches;[77] and teflon ((C
2
F
4
)n), used to create non-stick coatings for pans and other cookware.[78]

Complications edit

 
Periodic table highlighting the first row of each block.[n 12] 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.
 
Electronegativity values of the group 16 chalcogen elements showing a W-shaped alternation or secondary periodicity going down the group

Adding complexity to the chemistry of the nonmetals are the anomalies observed in the first row of each periodic table block. These anomalies are especially prominent in hydrogen, boron (whether as a nonmetal or metalloid), carbon, nitrogen, oxygen, and fluorine. In later rows these anomalies manifest as secondary periodicity or non-uniform periodic trends within most of the p-block groups,[79] and unusual oxidation states in the heavier nonmetals.

First row anomaly edit

Starting with hydrogen, the first row anomaly primarily arises from the electron configurations of the elements concerned. Hydrogen is particularly notable for its diverse bonding behaviors. It most commonly forms covalent bonds, but it can also lose its single electron in an aqueous solution, leaving behind a bare proton with tremendous polarizing power.[80] Consequently, this proton can attach itself to the lone electron pair of an oxygen atom in a water molecule, laying the foundation for acid-base chemistry.[81] Moreover, a hydrogen atom in a molecule can form a second, albeit weaker, bond with an atom or group of atoms in another molecule. As Cressey explains, such bonding, "helps give snowflakes their hexagonal symmetry, binds DNA into a 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."[82]

Hydrogen and helium, as well as boron through neon, have unusually small atomic radii. This phenomenon arises because the 1s and 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.[83] 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.[84]

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.[85]

Secondary periodicity edit

A noticeable alternation in certain periodic trends becomes evident when descending groups 13 to 15, and to a lesser extent, groups 16 to 17.[86][n 13] Immediately after the first row of d-block metals, from scandium to zinc, the 3d electrons in the p-block elements—specifically, gallium (a metal), germanium, arsenic, selenium, and bromine—prove less effective at shielding the increasing positive nuclear charge. This same effect is observed with the emergence of 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.[88]

Unusual oxidation states edit

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 the lowest for their group (that is, 3, 2, 1, or 0), for example in phosphorus pentachloride (PCl5), sulfur hexafluoride (SF6), iodine heptafluoride (IF7), and xenon difluoride (XeF2).[89]

Types edit

 
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 14]

The classification of nonmetals can vary, with approaches ranging from as few as two types to as many as six or seven. For instance, the periodic table in the Encyclopædia Britannica recognizes noble gases, halogens, and other nonmetals, and splits the elements commonly recognized as metalloids between "other metals" and "other nonmetals".[101] 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.[102]

When traversing the periodic table from right to left, three or four types of nonmetals are more or less commonly discerned:

  • the relatively inert noble gases;[103]
  • a set of chemically strong halogen elements—fluorine, chlorine, bromine and iodine—sometimes referred to as nonmetal halogens[104] or halogen nonmetals[105] (as used here) or stable halogens;[106]
  • a set of unclassified nonmetals, encompassing elements like hydrogen, carbon, nitrogen, and oxygen, for which there is no widely recognized collective name;[n 16] and
  • the chemically weak nonmetallic metalloids[115] which are sometimes considered nonmetals and sometimes not.[n 17]

Since metalloids occupy a transition region or "frontier territory",[117] where metals meet nonmetals, their classification varies among authors. Some consider them distinct from both metals and nonmetals, while others classify them as nonmetals[118] or as a sub-class of nonmetals.[119] There are also authors who categorize certain metalloids as metals, such as arsenic and antimony, due to their similarities to heavy metals.[120][n 18] In this context metalloids are here treated as nonmetals, based on their chemical behavior,[115] for the sake of comparative analysis.

Aside from the metalloids, some boundary fuzziness and overlapping (as occurs with classification schemes generally),[121] can be discerned among the other types of nonmetals. Carbon, phosphorus, selenium, and iodine border the metalloids and show some metallic character, as does hydrogen. Among the noble gases, radon is the most metallic and begins to show some cationic behavior, which is unusual for a nonmetal.[122]

Noble gases edit

 
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.[103]

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.[123] As a consequence, they all exist as gases under standard conditions, even those with atomic masses surpassing many typically solid elements.[124]

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,[125] with a most of these compounds involving the combination of oxygen or fluorine with either krypton, xenon, or radon.[126]

In terms of the periodic table, an analogy can be drawn between the noble gases and noble metals such as platinum and gold, as they share a similar reluctance to combine with other elements.[127] As a further example, xenon, in the +8 oxidation state, forms a pale yellow explosive oxide,XeO4, while osmium, another noble metal, forms a yellow, strongly oxidizing oxide,[128] OsO4. Additionally, there are parallels in the formulas of the oxyfluorides: XeO2F4 and OsO2F4, and XeO3F2 and OsO3F2.[129]

About 1015 tonnes of noble gases are present in the Earth's atmosphere.[130] Additionally, natural gas can contain as much as 7% helium.[131] Radon diffuses out of rocks, where it forms during the natural decay sequence of uranium and thorium.[132] In the Earth's core there may be around 1013 tons of xenon, in the form of stable XeFe3 and XeNi3 intermetallic compounds. This could explain why "studies of the Earth's atmosphere have shown that more than 90% of the expected amount of Xe is depleted."[133]

Halogen nonmetals edit

 
A cluster of purple fluorite CaF
2
, the most common fluorine mineral, between two quartzes. It is the main source of fluorine for commercial uses.[134]

Although the halogen nonmetals are notably reactive and corrosive elements, they can also be found in everyday compounds like toothpaste (NaF); common table salt (NaCl); swimming pool disinfectant (NaBr); or food supplements (KI). The term "halogen" itself means "salt former".[135]

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[90] solid. Electrically, the first three elements function as insulators while iodine behaves as a semiconductor (along its planes).[136]

Chemically, the halogen nonmetals exhibit high ionization energies, electron affinities, and electronegativity values, and are mostly relatively strong oxidizing agents.[137] These characteristics contribute to their corrosive nature.[138] All four elements tend to form primarily ionic compounds with metals,[139] in contrast to the remaining nonmetals (except for oxygen) which tend to form primarily covalent compounds with metals.[n 19] The highly reactive and strongly electronegative nature of the halogen nonmetals epitomizes nonmetallic character.[143]

In terms of the periodic table, the highly nonmetallic halogens in group 17 find their counterparts in the highly reactive alkali metals, such as sodium and potassium, in group 1.[144] Further, and much like the halogen nonmetals, most of the alkali metals are known to form –1 anions, a characteristic seldom observed among metals.[145]

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.[146]

Metalloids edit

 
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.[147]

The six elements more commonly recognized as metalloids are boron, silicon, germanium, arsenic, antimony, and tellurium, all of which have a metallic appearance. 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.[1]

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 electronic structures of semimetals, although both have less stable semiconducting forms.[1]

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.[1]

In the periodic table, to the left of the weakly nonmetallic metalloids are an indeterminate set of weakly metallic metals including tin, lead and bismuth)[148] sometimes referred to as post-transition metals.[149] Dingle explains the situation this way:

... with 'no-doubt' metals on the far left of the table, and no-doubt non-metals on the far right ... the gap between the two extremes is bridged first by the poor (post-transition) metals, and then by the metalloids—which, perhaps by the same token, might collectively be renamed the 'poor non-metals'.[150]

The metalloids are commonly found combined with oxygen, sulfur, or, in the case of tellurium, gold or silver.[151] Boron is typically found in boron-oxygen borate minerals, including in volcanic spring waters. Silicon is present in silicon-oxygen mineral 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.[152]

Unclassified nonmetals edit

 
Selenium conducts electricity around 1,000 times better when light falls on it, a property used in light-sensing applications.[153]

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[154] and a semiconductor perpendicular to its planes;[155] phosphorus and selenium are semiconductors;[156] while hydrogen, nitrogen, oxygen, and sulfur are insulators.[n 20]

These elements are often considered too diverse to merit a collective classification,[158] and have been referred to as other nonmetals,[159] or simply as nonmetals, located between the metalloids and the halogens.[160] As a result, their chemistry is typically taught disparately, according to their respective periodic table groups:[161] 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 21]

Hydrogen, in particular, behaves in some respects like a metal and in others like a nonmetal.[163] Like a metal, it can initially lose its single electron,[164] it can substitute for alkali metals in typical alkali metal structures,[165] and it can form alloy-like hydrides, featuring metallic bonding, with certain transition metals.[166] Conversely, it behaves as an insulating diatomic gas, akin to a typical nonmetal, and during chemical reactions, it tends to ultimately attain the electron configuration of helium.[167] It accomplishes this by forming a covalent or ionic bonds[168] or, if it has lost its electron, by attaching itself to a lone pair of electrons.[169]

Some or all of these nonmetals share several properties. Being less reactive than the halogens,[170] they can occur naturally in the environment.[171] They have significant roles in biology[172] and geochemistry.[173] Collectively, their physical and chemical characteristics can be described as "moderately non-metallic".[173] However, they all have corrosive aspects. Hydrogen can corrode metals. Carbon corrosion can occur in fuel cells.[174] Acid rain is caused by dissolved nitrogen or sulfur. Oxygen causes iron to corrode via rust. White phosphorus, the most unstable form, ignites in air and leaves behind phosphoric acid residue.[175] Untreated selenium in soils can lead to the formation of corrosive hydrogen selenide gas.[176] When combined with metals, the unclassified nonmetals can form high-hardness (interstitial or refractory) compounds[177] due to their relatively small atomic radii and sufficiently low ionization energies.[173] 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]

In terms of the periodic table, there is a geographic analogy between the unclassified nonmetals and transition metals. The unclassified nonmetals are positioned between the strongly nonmetallic halogens on the right and the weakly nonmetallic metalloids on the left. Similarly, the transition metals occupy a region between the "virulent and violent" metals on the left side of the periodic table, and the "calm and contented" metals on the right. They effectively serve as a "transitional bridge" connecting these two regions.[180]

Unclassified nonmetals are typically found in elemental forms or in association with other elements:[151]

  • Hydrogen is present in the world's oceans as a component of water and occurs in natural gas as a component of methane and hydrogen sulfide.[181]
  • Carbon can be found in limestone, dolomite, and marble, as carbonates.[182] Additionally, carbon exists as graphite, primarily occurring in metamorphic silicate rocks,[183] resulting from the compression and heating of sedimentary carbon compounds.[184]
  • Oxygen is found in the atmosphere; in the oceans as a component of water; and in the Earth's crust as oxide minerals.[185]
  • Phosphorus minerals are widespread, typically appearing as phosphorus-oxygen phosphates.[186]
  • Elemental sulfur can be found in or near hot springs and volcanic regions around the world, while sulfur minerals are common and are often found as sulfides or oxygen-sulfur sulfates.[187]
  • Selenium is found in metal sulfide ores, where it may partially replace sulfur. In rare instances, elemental selenium can also be found.[188]

Prevalence and access edit

Abundance edit

Approximate nonmetal composition
of the Earth and its biomass, by weight[189]
Domain Main components Next most
abundant
Crust O 61%, Si 20% H 2.9%
Atmosphere N 78%, O 21% Ar 0.5%
Hydrosphere O 66.2%, H 33.2% Cl 0.3%
Biomass O 63%, C 20%, H 10% N 3.0%

Hydrogen and helium dominate the universe, comprising an estimated 99% of all ordinary matter and over 99.9% of all atoms.[190] Oxygen, the next most abundant element, constitutes around 0.1% of the universe's composition.[191] Less than 5% of the universe consists of ordinary matter, which includes stars, planets, and living entities, while the majority remains enigmatic, comprising dark energy and dark matter, both yet poorly understood.[192]

Five nonmetals—hydrogen, carbon, nitrogen, oxygen, and silicon—play crucial roles in shaping Earth's composition. These elements form the fundamental building blocks of the Earth's crust, atmosphere, hydrosphere, and biomass. Their relative proportions, as outlined in the accompanying table, underscore their fundamental significance in the terrestrial environment.

Extraction edit

 
Germanium occurs in some zinc-copper-lead ore bodies, in quantities sufficient to justify extraction.[193] In 2022, the 99.999% pure form was priced at US$1300 per kilogram.[194]

Nonmetals, and metalloids, are extracted in their raw forms from:[171]

  • brine—chlorine, bromine, iodine;
  • liquid air—nitrogen, oxygen, neon, argon, krypton, xenon;
  • minerals—boron (borate minerals); carbon (coal; diamond; graphite); fluorine (fluorite); silicon (silica); phosphorus (phosphates); antimony (stibnite, tetrahedrite); iodine (in sodium iodate and sodium iodide);
  • natural gas—hydrogen, helium, sulfur; and
  • ores, as processing byproducts—germanium (zinc ores); arsenic (copper and lead ores); selenium, tellurium (copper ores); and radon (uranium-bearing ores).

Cost edit

Day to day costs will vary depending on purity, quantity,[n 22] market conditions, and supplier surcharges.[197]

Based on the available literature as of April 2023, the cited costs of most nonmetals are less than the US$0.74 per gram cost of silver.[198] The exceptions are boron, phosphorus, germanium, xenon, and radon (notionally):

  • Boron costs around $25 per gram for 99.7% pure polycrystalline chunks with a particle size of about 1 cm.[199] Earlier, in 1997, boron was quoted at $280 per gram for polycrystalline 4-to-6-mm-diameter rods of 99.999% purity,[200] about 10 times the then $28.35 per gram cost of gold.[201]
  • In 2020, phosphorus in its most-stable black form could "cost up to $1,000 per gram",[202] more than 15 times the cost of gold, whereas ordinary red phosphorus, in 2017, was priced at about $3.40 per kilogram.[203] Researchers hoped to be able to reduce the cost of black phosphorus to as low as $1 per gram.[202]
  • Germanium and xenon cost about $1.30 and $7.60 per gram.[204]
  • Up to 2013, radon was available from the National Institute of Standards and Technology for $1,636 per 0.2 ml unit of issue, equivalent to about $86,000,000 per gram, with no indication of a discount for bulk quantities.[205]

Uses edit

 
A high-voltage circuit-breaker employing sulfur hexafluoride (SF6) as its inert (air replacement) interrupting medium[206]

Nonmetallic elements display unique properties that enable a wide range of applications, often unachievable solely with metallic elements. A remarkable aspect of nonmetals is their ubiquitous presence in living organisms, where hydrogen, oxygen, carbon, and nitrogen serve as the foundational building blocks of life. Additionally, these elements play pivotal roles in various industries, encompassing cutting-edge electronics, advanced energy storage systems, agriculture, and large-scale chemical production. A few examples include:

  • Nitrogen plays a vital role in fertilizer production, contributing to enhanced crop growth and increased agricultural productivity. Its low-temperature characteristics find utility in cryogenic applications, including the preservation of biological samples and the freezing of food.[208]
  • Oxygen is indispensable for human respiration in life support systems, and it is employed in medical contexts to aid patients with respiratory challenges. It also supports combustion and finds application in various industrial processes, including metal smelting and waste incineration.[209]
  • Sulfur is an essential component in the production of sulfuric acid, one of the most extensively utilized industrial chemicals. Additionally, it plays a key role in the synthesis of various organic compounds. Sulfur compounds are critically involved in the vulcanization process, enhancing the strength and elasticity of rubber products.[211]

Various subsets of nonmetals find common applications in fields such as (inert) air replacements, dyestuffs, flame retardants or extinguishers, household items, lasers and lighting, mineral acids, plug-in hybrid vehicles, and welding gases.[171][212] Metalloids, to the extent that they exhibit metallic properties, have specialized applications that include oxide glasses and alloying components, among others.[213]

History, background, and taxonomy edit

Discovery edit

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 23] Radon is the most recently uncovered nonmetal, with its detection occurring at the end of the 19th century.[171]

The isolation of nonmetallic elements depended on a range of chemical and physical techniques. These methods encompassed spectroscopy, fractional distillation, radiation detection, electrolysis, ore acidification, displacement reactions, combustion, and controlled heating processes. While some nonmetals were naturally occurring as free elements, others required intricate extraction procedures:

  • The noble gases, renowned for their low reactivity, were first identified through unusual and mundane methods. Helium was initially detected by its distinctive yellow line in the solar corona spectrum. Subsequently, it was observed escaping as bubbles when uranite UO2 was dissolved in acid. Neon, argon, krypton, and xenon were obtained through the fractional distillation of air. The discovery of radon occurred three years after Henri Becquerel's pioneering research on radiation in 1896.[215]
  • The isolation of halogen nonmetals from their halides involved techniques including electrolysis, acid addition, or displacement. These efforts were not without peril, as some chemists tragically lost their lives in their pursuit of isolating fluorine.[216]
  • The unclassified nonmetals have a diverse history. Hydrogen was discovered and first described in 1671 as the product of the reaction between iron filings and dilute acids. Carbon was found naturally in forms like charcoal, soot, graphite, and diamond. Nitrogen was discovered by examining air after carefully removing oxygen. Oxygen itself was obtained by heating mercurous oxide. Phosphorus was derived from the heating of ammonium sodium hydrogen phosphate (Na(NH4)HPO4), a compound found in urine.[217] Sulfur occurred naturally as a free element, simplifying its isolation. Selenium,[n 24] was first identified as a residue in sulfuric acid.[219]
  • Metalloids were commonly isolated by heating of their oxides ((boron, silicon, arsenic, tellurium) or a sulfide (germanium).[171] Antimony, in an unusual twist, was found in its native form and could also be obtained through heating of its sulfide compound.[220]

Origin and use of the term edit

 
An extract from the English translation of Lavoisier's Traité élémentaire de chimie (1789),[221] 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 Meteorology, the ancient Greek philosopher Aristotle categorized substances found within the Earth into two distinct groups: metals and "fossiles".[n 25] The latter category included various minerals such as realgar, ochre, ruddle, sulfur, cinnabar, and other substances that he referred to as "stones which cannot be melted".[222]

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.[223]

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.[224]

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.[225]

The pivotal moment in the systematic classification of chemical elements, distinguishing between metallic and nonmetallic substances, came in 1789 with the groundbreaking work of Antoine Lavoisier. Lavoisier, a French chemist, published the first modern list of chemical elements in his revolutionary[226] work Traité élémentaire de chimie. In this work he categorized elements into distinct groups, including gases, metallic substances, nonmetallic substances, and earths (heat-resistant oxides).[227] 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.[228]

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"[229][230] to describe nonmetallic elements, noting their ability to form negatively charged ions with oxygen in aqueous solutions.[231][232] While Berzelius' terminology gained significant acceptance,[229] it later faced criticism from some who found it counterintuitive,[232] misapplied,[233] or even invalid.[234][235] In 1864, reports indicated that the term "metalloids" was still endorsed by leading authorities,[236] but there were reservations about its appropriateness. The idea of designating elements like arsenic as metalloids had been considered.[236] By as early as 1866, some authors began preferring the term "nonmetal" over "metalloid" to describe nonmetallic elements.[237] In 1875, Kemshead 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.[238]

Suggested distinguishing criteria edit

Some single properties used to distinguish between
metals and nonmetals classified by type and date of source
Physical Chemical

Electron related

In 1809, the British chemist and inventor Humphry Davy made a groundbreaking discovery that reshaped the understanding of metals and nonmetals.[261] 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.[262] Sodium and potassium, on the contrary, floated on water. Nevertheless, their classification as metals was firmly established by their distinct chemical properties.[263]

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, listed by type and the year of their mention.

One of the most commonly recognized (though somewhat unreliable) properties used in this context is the change in electrical conductivity with temperature. Typically, metals exhibit an increase in electrical conductivity as temperature decreases, while nonmetals display the opposite trend.[251] However, there are exceptions that challenge this generalization. For example, plutonium, carbon, arsenic, and antimony defy the norm. Plutonium's electrical conductivity increases when heated within a specific temperature range of −175 to +125 °C.[264] Similarly, despite its common classification as a nonmetal, carbon also experiences an increase in electrical conductivity when subjected to heating.[265] Arsenic and antimony, which are occasionally classified as nonmetals, exhibit behavior similar to carbon, highlighting the complexity of the distinction between metals and nonmetals.[266]

Kneen and colleagues 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.[9]

However, Emsley pointed out the complexity of this task, asserting that no single property alone can unequivocally assign elements to either the metal or nonmetal category.[267] Furthermore, Jones emphasized that classification systems typically rely on more than two attributes to define distinct types.[268]

Metals and nonmetals sorted by
density and electronegativity (EN)[n 27]
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

An approach to distinguishing between metallic and nonmetallic properties was suggested by Johnson, emphasizing the significance of physical properties, while acknowledging the potential need for other properties in certain ambiguous cases. His observations highlighted several key distinctions:[273]

  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 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.

Hein and Arena observed that nonmetals generally have low densities and high electronegativity,[64] which is consistent with the data presented in the table. Nonmetallic elements are predominantly located in the top left quadrant of this table, where both density and electronegativity values are relatively high. In contrast, the other three quadrants are primarily occupied by metals.

While some authors choose to further subdivide elements into metals, metalloids, and nonmetals, Oderberg disagrees with this approach. He argues that according to the principles of categorization, anything not classified as a metal should be considered a nonmetal.[274]

Development of types edit

 
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,[275] established a basic taxonomy of nonmetals to aid in the study of these elements. He wrote:[276]

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.[277] The boroid nonmetals eventually evolved into the metalloids, with this classification beginning as early as 1864.[236] The noble gases were also identified as a distinct group among the nonmetals, dating back to as early as 1900.[278]

Comparison of selected properties edit

The two tables in this section present a selection of physical and chemical properties of metals,[n 28] metalloids, unclassified nonmetals, halogen nonmetals, and noble gases, based on the elements' most stable forms in ambient conditions.

The purpose is to show the left-to-right progression of metallic-to-nonmetallic character across the periodic table.[279]

The dashed lines encircling 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 edit

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

Property Metals Metalloids Unc. nonmetals Halogen nonmetals Noble gases
Form and heft[280] 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[281] He, Ne lighter than air[282]
Appearance lustrous[21] lustrous[283]
  • ◇ lustrous: C, P, Se[284]
  • ◇ colorless: H, N, O[285]
  • ◇ colored: S[286]
  • ◇ colored: F, Cl, Br[287]
  • ◇ lustrous: I[1]
colorless[288]
Elasticity mostly malleable and ductile[21] (Hg is liquid) brittle[283] C, black P, S, Se brittle; all four have less stable non-brittle forms[289][n 29] iodine is brittle[291] not applicable
Electrical conductivity good[n 30]
  • ◇ moderate: B, Si, Ge, Te
  • ◇ good: As, Sb[n 31]
  • ◇ poor: H, N, O, S
  • ◇ moderate: P, Se
  • ◇ good: C[n 32]
  • ◇ poor: F, Cl, Br
  • ◇ moderate: I[n 33]
poor[n 34]
Electronic structure[295] 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 edit

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

Property Metals Metalloids Unc. nonmetals Halogen nonmetals Noble gases
General chemical behavior
weakly nonmetallic[n 35] moderately nonmetallic[298] strongly nonmetallic[299]
  • ◇ inert to nonmetallic[300]
  • ◇ Rn shows some cationic behavior[301]
Oxides basic; some amphoteric or acidic[302] amphoteric or weakly acidic[303][n 36] acidic[n 37] or neutral[n 38] acidic[n 39] metastable XeO3 is acidic;[308] stable XeO4 strongly so[309]
few glass formers[n 40] all glass formers[311] some glass formers[n 41] no glass formers reported no glass formers reported
ionic, polymeric, layer, chain, and molecular structures[313] polymeric in structure[314]
  • ◇ mostly molecular[314]
  • ◇ C, P, S, Se have at least one polymeric form
  • ◇ mostly molecular
  • ◇ iodine has at least one polymeric form, I2O5[315]
  • ◇ mostly molecular
  • XeO2 is polymeric[316]
Compounds with metals alloys[21] or intermetallic compounds[317] tend to form alloys or intermetallic compounds[318]
  • ◇ salt-like to covalent: H†, C, N, P, S, Se[319]
  • ◇ mainly ionic: O[320]
mainly ionic[139] simple compounds in ambient conditions not known[n 42]
Ionization energy (kJ mol−1)‡
[322]
  • ◇ low to high
  • ◇ 376 to 1,007
  • ◇ average 643
  • ◇ moderate
  • ◇ 762 to 947
  • ◇ average 833
  • ◇ moderate to high
  • ◇ 941 to 1,402
  • ◇ average 1,152
  • ◇ high
  • ◇ 1,008 to 1,681
  • ◇ average 1,270
  • ◇ high to very high
  • ◇ 1,037 to 2,372
  • ◇ average 1,589
Electronegativity (Pauling)[n 43]
[324]
  • ◇ low to high
  • ◇ 0.7 to 2.54
  • ◇ average 1.5
  • ◇ moderately high
  • ◇ 1.9 to 2.18
  • ◇ average 2.05
  • ◇ moderately high to high
  • ◇ 2.19 to 3.44
  • ◇ average 2.65
  • ◇ high
  • ◇ 2.66 to 3.98
  • ◇ average 3.19
  • ◇ high (Rn) to very high
  • ◇ ca. 2.43 to 4.7
  • ◇ average 3.3

† Hydrogen can also form alloy-like hydrides[166]
‡ The labels low, moderate, high, and very high are arbitrarily based on the value spans listed in the table

Notes edit

  1. ^ For example, of authors who recognize metalloids as a distinct type of element, about a quarter include selenium.[1]
  2. ^ Hydrogen has historically been placed over one or more of lithium, boron,[2] carbon, or fluorine; or over no group at all; or over all main groups simultaneously, and therefore may or may not be adjacent to other nonmetals.[3]
  3. ^ Metallic or nonmetallic character is usually taken to be indicated by one property rather than two or more properties.
  4. ^ Steudel's monograph is an updated translation of the fifth German edition of 2013, incorporating the literature up to Spring 2019.
  5. ^ When synthesized in 1940 the discoverers of astatine considered it to be a metal.[16] Subsequently it was reported to show both metallic and nonmetallic properties.[17] In 2013, on the basis of relativistic studies, astatine was predicted to be a metal.[4] For a summary of the properties of astatine see the Metalloid article.
  6. ^ Solid iodine has a silvery metallic appearance under white light at room temperature.[19] It volatizes at ordinary and higher temperatures, passing from solid to gas; its vapours are violet-colored.[20]
  7. ^ The solid nonmetals have electrical conductivity values ranging from 10−18 S•cm−1 for sulfur[23] to 3 × 104 in graphite[24] or 3.9 × 104 for arsenic;[25] cf. 0.69 × 104 for manganese to 63 × 104 for silver, metals both.[23]
  8. ^ 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.[41]
  9. ^ 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.[23] Electrical conductivity values of metals range from 0.69 S•cm−1 × 104 for manganese to 63 × 104 for silver; cf. carbon 3 × 104,[24] arsenic 3.9 × 104 and antimony 2.3 × 104.[23]
  10. ^ These elements being semiconductors.[46]
  11. ^ Acids are formed by boron, phosphorus, selenium, arsenic, iodine;[65] oxides by carbon, silicon, germanium, sulfur, antimony, and tellurium.[66]
  12. ^ 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.
  13. ^ The net result is an even-odd difference between periods (except in the 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.[87]
  14. ^ 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.[90] The remaining reactive nonmetallic elements have giant covalent structures, but for H which is a diatomic gas.[91]

    The single dagger nonmetals N, S and iodine are somewhat hobbled as "strong" nonmetals.


    While N has a high electronegativity, it is a reluctant anion former,[92] and a pedestrian oxidizing agent unless combined with a more active nonmetal like O or F.[93]


    S reacts in the cold with alkalic and post-transition metals, and Cu, Ag and Hg,[94] 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.[95]
    Iodine is sufficiently corrosive to cause lesions resembling thermal burns, if handled without suitable protection,[96] and tincture of iodine will smoothly dissolve Au.[97] That said, while "F, Cl and Br will all oxidize Fe2+ (aq) to Fe3+(aq) ... iodine ... is such a [relatively] weak oxidizing agent that it cannot remove electrons from Fe(II) ions to form Fe(III) ions."[98] Thus, for the reaction X2 + 2e → 2X(aq) the reduction potentials are F +2.87 V; Cl +1.36; Br +1.09; I +0.54. Here Fe3+ + e → Fe3+ +0.77.[99] 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.[100]
  15. ^ 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.
  16. ^ Varying configurations of these nonmetals have been referred to as, for example, basic nonmetals,[107] bioelements,[108] central nonmetals,[109] CHNOPS,[110] essential elements,[111] "non-metals",[112][n 15] orphan nonmetals,[113] or redox nonmetals.[114]
  17. ^ 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".[116]
  18. ^ 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".[121]
  19. ^ Metal oxides are usually ionic.[140] On the other hand, oxides of metals with high oxidation states are usually either polymeric or covalent.[141] A polymeric oxide has a linked structure composed of multiple repeating units.[142]
  20. ^ 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.[157]
  21. ^ 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.[162]
  22. ^ For example, as at April 2023, the commercial price of silicon was $4 per pound or $0.0088 per gram.[195] On the other hand, the price quoted for a 335 gram sample of silicon for hobbyists and science enthusiasts was about $57, or 0.170 per gram, or about 20 times the commercial price.[196]
  23. ^ How helium acquired the -ium suffix is explained in the following passage by its discoverer, William Lockyer: "I took upon myself the responsibility of coining the word helium ... I did not know whether the substance ... was a metal like calcium or a gas like hydrogen, but I did know that it behaved like hydrogen [being found in the sun] and that hydrogen, as Dumas had stated, behaved as a metal".[214]
  24. ^ Berzelius, who discovered selenium, thought it had the properties of a metal, combined with the properties of sulfur.[218]
  25. ^ Not to be confused with the modern usage of fossil to refer to the preserved remains, impression, or trace of any once-living thing
  26. ^ The Goldhammer-Herzfeld ratio is roughly equal to the cube of the atomic radius divided by the molar volume.[242] 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.[243]
  27. ^ (a) Up to element 99 (Es), with the values taken from Aylward and Findlay.[269]
    (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.[270]
    (c) The density values used for At and Fr are theoretical estimates.[271]
    (d) Bjerrum classified "heavy metals" as those metals with densities above 7 g/cm3.[272]
    (e) Vernon specified a minimum electronegativity of 1.9 for the metalloids, on the revised Pauling scale.[1]
  28. ^ Metals are included for reference
  29. ^ Carbon as exfoliated (expanded) graphite,[47][290] and as carbon nanotube wire;[49] phosphorus as white phosphorus (soft as wax, pliable and can be cut with a knife, at room temperature);[50] sulfur as plastic sulfur;[51] and selenium as selenium wires.[52]
  30. ^ Metals have electrical conductivity values of from 6.9×103 S•cm−1 for manganese to 6.3×105 for silver.[292]
  31. ^ Metalloids have electrical conductivity values of from 1.5×10−6 S•cm−1 for boron to 3.9×104 for arsenic.[293]
  32. ^ Unclassified nonmetals have electrical conductivity values of from ca. 1×10−18 S•cm−1 for the elemental gases to 3×104 in graphite.[294]
  33. ^ 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.[294][136]
  34. ^ The elemental gases have electrical conductivity values of ca. 1×10−18 S•cm−1.[294]
  35. ^ They always give "compounds less acidic in character than the corresponding compounds of the [typical] nonmetals".[283]
  36. ^ Arsenic trioxide reacts with sulfur trioxide, forming arsenic "sulfate" As2(SO4)3.[304]
  37. ^ NO
    2
    , N
    2
    O
    5
    , SO
    3
    , SeO
    3
    strongly so[305]
  38. ^ (H2O, CO, NO, N2O); CO and N2O are "formally the anhydrides of formic and hyponitrous acid, respectively viz. CO + H2O → H2CO2 (HCOOH, formic acid); N2O + H2O → H2N2O2 (hyponitrous acid)".[306]
  39. ^ ClO
    2
    , Cl
    2
    O
    7
    , I
    2
    O
    5
    strongly so[307]
  40. ^ V; Mo, W; Al, In, Tl; Sn, Pb; Bi[310]
  41. ^ P, S, Se;[310] CO2 forms a glass at 40 GPa[312]
  42. ^ 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.[321]
  43. ^ Values for the noble gases are from Rahm, Zeng and Hoffmann.[323]

References edit

Citations edit

  1. ^ a b c d e f Vernon 2013
  2. ^ Luchinskii & Trifonov 1981, pp. 200–220
  3. ^ Rayner-Canham 2020, p. 212
  4. ^ a b c Restrepo et al. 2006, p. 411; Thornton & Burdette 2010, p. 86; Hermann, Hoffmann & Ashcroft 2013, pp. 11604‒1‒11604‒5
  5. ^ Parkes & Mellor 1943, p. 740
  6. ^ Pascoe 1982, p. 3
  7. ^ Glinka 1958, p. 77; Oxtoby, Gillis & Butler 2015, p. I.23
  8. ^ Godovikov & Nenasheva 2020, p. 4; Sanderson 1957, p. 229; Morely & Muir 1892, p. 241
  9. ^ a b Kneen, Rogers & Simpson 1972, pp. 218–219
  10. ^ Steudel 2020, p. 43
  11. ^ a b c Larrañaga, Lewis & Lewis 2016, p. 988
  12. ^ Vernon 2020, p. 220; Rochow 1966, p. 4
  13. ^ IUPAC Periodic Table of the Elements
  14. ^ Johnson 2007, p. 13
  15. ^ Bodner & Pardue 1993, p. 354; Cherim 1971, p. 98
  16. ^ Vasáros & Berei 1985, p. 109
  17. ^ Nefedov et al. 1968, p. 87
  18. ^ Mewes et al. 2019; Smits et al. 2020; Florez et al. 2022
  19. ^ Koenig 1962, p. 108
  20. ^ Tidy 1887, pp. 107–108
  21. ^ a b c d Kneen, Rogers & Simpson 1972, pp. 261–264
  22. ^ Phillips 1973, p. 7
  23. ^ a b c d e Aylward & Findlay 2008, pp. 6–12
  24. ^ a b Jenkins & Kawamura 1976, p. 88
  25. ^ Carapella 1968, p. 30
  26. ^ Zumdahl & DeCoste 2010, pp. 455, 456, 469, A40; Earl & Wilford 2021, p. 3-24
  27. ^ Still 2016, p. 120
  28. ^ Wiberg 2001, pp. 780
  29. ^ Wiberg 2001, pp. 824, 785
  30. ^ Earl & Wilford 2021, p. 3-24
  31. ^ Siekierski & Burgess 2002, p. 86
  32. ^ Charlier, Gonze & Michenaud 1994
  33. ^ 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).
  34. ^ Wiberg 2001, pp. 742
  35. ^ Evans 1966, pp. 124–25
  36. ^ Wiberg 2001, pp. 758
  37. ^ Stuke 1974, p. 178; Donohue 1982, pp. 386–87; Cotton et al. 1999, p. 501
  38. ^ 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 ..."
  39. ^ Wiberg 2001, p. 416; Wiberg is here referring to iodine.
  40. ^ Elliot 1929, p. 629
  41. ^ Fox 2010, p. 31
  42. ^ Wibaut 1951, p. 33: "Many substances ...are colourless and therefore show no selective absorption in the visible part of the spectrum."
  43. ^ Kneen, Rogers & Simpson 1972, pp. 85–86, 237
  44. ^ Salinas 2019, p. 379
  45. ^ Yang 2004, p. 9
  46. ^ Wiberg 2001, pp. 416, 574, 681, 824, 895, 930; Siekierski & Burgess 2002, p. 129
  47. ^ a b Chung 1987
  48. ^ Godfrin & Lauter 1995
  49. ^ a b Janas, Cabrero-Vilatela & Bulmer 2013
  50. ^ a b Faraday 1853, p. 42; Holderness & Berry 1979, p. 255
  51. ^ a b Partington 1944, p. 405
  52. ^ a b Regnault 1853, p. 208
  53. ^ Edwards 2000, pp. 100, 102–103; Herzfeld 1927, pp. 701–705
  54. ^ Kneen, Rogers & Simpson 1972, pp. 263‒264
  55. ^ Langley & Hattori 2014, p. 214
  56. ^ a b Abbott 1966, p. 18
  57. ^ Brown et al. 2014, p. 237
  58. ^ Barton 2021, p. 200
  59. ^ Borg & Dienes 1992, p. 26
  60. ^ Wiberg 2001, p. 796
  61. ^ Shang et al. 2021
  62. ^ Tang et al. 2021
  63. ^ Shanabrook, Lannin & Hisatsune 1981, pp. 130–133
  64. ^ a b Hein & Arena 2013, pp. 226, G-6
  65. ^ Lidin 1996, pp. 22, 29; 322, 165; 381, 173–174; 12, 147; 157 [B; P; Se; As; I]; Housecroft & Sharpe 2008, p. 472 [I]
  66. ^ Lidin 1996, pp. 52, 58; 386; 140; 361, 365; 372, 376; 403 [C; Si; Ge; S; Sb; Te]; Rochow 1973, p. 1338 [Si]; Sanderson 1967, p. 172 [Ge]; Shkol'nikov 2010, p. 2127 [Sb]; Wiberg 2001, pp. 592 [Te]
  67. ^ Matson & Orbaek 2013, p. 85
  68. ^ Yoder, Suydam & Snavely 1975, p. 58
  69. ^ Young et al. 2018, p. 753
  70. ^ Brown et al. 2014, p. 227
  71. ^ Siekierski & Burgess 2002, pp. 21, 133, 177
  72. ^ Moore 2016; Burford, Passmore & Sanders 1989, p. 54
  73. ^ King & Caldwell 1954, p. 17; Brady & Senese 2009, p. 69
  74. ^ Chemical Abstracts Service 2021
  75. ^ Emsley 2011, pp. 81
  76. ^ Cockell 2019, p. 210
  77. ^ Scott 2014, p. 3
  78. ^ Emsley 2011, p. 184
  79. ^ Kneen, Rogers & Simpson 1972, pp. 226, 360
  80. ^ Lee 1996, p. 240
  81. ^ Greenwood & Earnshaw 2002, p. 43
  82. ^ Cressey 2010
  83. ^ Siekierski & Burgess 2002, pp. 24–25
  84. ^ Siekierski & Burgess 2002, p. 23
  85. ^ Petruševski & Cvetković 2018; Grochala 2018
  86. ^ Siekierski & Burgess 2002, pp. 52, 101, 111, 124, 194
  87. ^ Scerri 2020, pp. 407–420
  88. ^ Greenwood & Earnshaw 2002, pp. 27, 1232, 1234
  89. ^ Cox 2004, p. 146
  90. ^ a b Vernon 2013, p. 1706
  91. ^ Wiberg 2001, passim
  92. ^ Vernon 2020, p. 222
  93. ^ Atkins & Overton 2010, pp. 377, 389
  94. ^ Moody 1991, p. 391
  95. ^ Rodgers 2012, p. 504; Wulfsberg 2000, p. 726
  96. ^ Stellman 1998, chapter 104–211
  97. ^ Nakao 1992, p. 426–427
  98. ^ Hill & Holman 2000, p. 196
  99. ^ Wiberg 2001, pp. 1761–1762
  100. ^ Young 2006, p. 1285
  101. ^ Encyclopædia Britannica 2021
  102. ^ Royal Society of Chemistry 2021
  103. ^ a b Matson & Orbaek 2013, p. 203
  104. ^ Chambers & Holliday 1982, pp. 273–274; Bohlmann 1992, p. 213; Jentzsch 2015, p. 247
  105. ^ Kernion 2019, p. 191; Cao et al. 2021, pp. 20–21; Hussain et al. 2023
  106. ^ Vassilakis, Kalemos & Mavridis 2014, p. 1; Hanley & Koga 2018, p. 24; Kaiho 2017, ch. 2, p. 1
  107. ^ Williams 2007, pp. 1550–1561: H, C, N, P, O, S
  108. ^ Wächtershäuser 2014, p. 5: H, C, N, P, O, S, Se
  109. ^ Hengeveld & Fedonkin, pp. 181–226: C, N, P, O, S
  110. ^ Wakeman 1899, p. 562
  111. ^ Fraps 1913, p. 11: H, C, Si, N, P, O, S, Cl
  112. ^ Parameswaran at al. 2020, p. 210: H, C, N, P, O, S, Se
  113. ^ Knight 2002, p. 148: H, C, N, P, O, S, Se
  114. ^ Fraústo da Silva & Williams 2001, p. 500: H, C, N, O, S, Se
  115. ^ a b Bailar et al. 1989, p. 742
  116. ^ Tshitoyan et al. 2019, pp. 95–98
  117. ^ Russell & Lee 2005, p. 419
  118. ^ Hampel & Hawley 1976, p. 174;
  119. ^ Goodrich 1844, p. 264; The Chemical News 1897, p. 189; Hampel & Hawley 1976, p. 191; Lewis 1993, p. 835; Hérold 2006, pp. 149–50
  120. ^ Tyler 1948, p. 105; Reilly 2002, pp. 5–6
  121. ^ a b Jones 2010, pp. 169–71
  122. ^ Stein 1983, p. 165
  123. ^ Jolly 1966, p. 20
  124. ^ Clugston & Flemming 2000, pp. 100–101, 104–105, 302
  125. ^ Maosheng 2020, p. 962
  126. ^ Mazej 2020
  127. ^ Wiberg 2001, p. 1131
  128. ^ Thompson 2004
  129. ^ Vernon 2020, p. 229
  130. ^ Cox 2000, pp. 258–259; Möller 2003, p. 173; Trenberth & Smith 2005, p. 864
  131. ^ Emsley 2011, p. 220
  132. ^ Emsley 2011, p. 440
  133. ^ Zhu et al. 2014, pp. 644–648
  134. ^ Emsley 2011, p. 181
  135. ^ Wiberg 2001, pp. 4022
  136. ^ a b Greenwood & Earnshaw 2002, p. 804
  137. ^ Rudolph 1973, p. 133: "Oxygen and the halogens in particular ... are therefore strong oxidizing agents."
  138. ^ Daniel & Rapp 1976, p. 55
  139. ^ a b Cotton et al. 1999, p. 554
  140. ^ Woodward et al. 1999, pp. 133–194
  141. ^ Phillips & Williams 1965, pp. 478–479
  142. ^ Moeller et al. 2012, p. 314
  143. ^ Lanford 1959, p. 176
  144. ^ Rayner-Canham 2020, p. 92, 139
  145. ^ Massey 2000, p. 113
  146. ^ Schmedt, Mangstl & Kraus 2012, p. 7847‒7849
  147. ^ Bailar, Moeller & Kleinberg 1965, p. 477; Mee 1964, p. 153
  148. ^ Masterton, Hurley & Neth 2011, p. 38
  149. ^ McCue 1963, p. 264
  150. ^ Dingle 2017, p. 101
  151. ^ a b Cox 1997, pp. 130–132; Emsley 2011, passim
  152. ^ Hurlbut 1961, p. 132
  153. ^ Emsley 2011, p. 478
  154. ^ Greenwood & Earnshaw 2002, p. 277
  155. ^ Atkins et al. 2006, p. 320
  156. ^ Greenwood & Earnshaw 2002, p. 482; Berger 1997, p. 86
  157. ^ Moss 1952, pp. 180, 202
  158. ^ Cite error: The named reference Cao2021 was invoked but never defined (see the help page).
  159. ^ Challoner 2014, p. 5; Government of Canada 2015; Gargaud et al. 2006, p. 447
  160. ^ Crichton 2012, p. 6; Scerri 2013; Los Alamos National Laboratory 2021
  161. ^ Vernon 2020, p. 218
  162. ^ Wulfsberg 2000, pp. 273–274, 620
  163. ^ Seese & Daub 1985, p. 65
  164. ^ MacKay, MacKay & Henderson 2002, p. 209
  165. ^ Cousins, Davidson & García-Vivó 2013, pp. 11809–11811
  166. ^ a b Cao et al. 2021, p. 4
  167. ^ Liptrot 1983, p. 161
  168. ^ Wiberg 2001, pp. 255–257
  169. ^ Scott & Kanda 1962, p. 153
  170. ^ Taylor 1960, p. 316
  171. ^ a b c d e Emsley 2011, passim
  172. ^ 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."
  173. ^ a b c Cao et al. 2021, p. 20
  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. ^ Atkins 2001, pp. 24–25
  181. ^ National Center for Biotechnology Information 2021
  182. ^ Emsley 2011, p. 113
  183. ^ Greenwood & Earnshaw 2002, p. 270–271
  184. ^ Khan 2001, p. 59
  185. ^ Emsley 2011, pp. 376, 380, 640
  186. ^ Cox 1997, pp. 130; Emsley 2011, p. 393
  187. ^ Cox 1997, pp. 130; Emsley 2011, pp. 515–516, 518
  188. ^ Boyd 2011, p. 570
  189. ^ Nelson 1987, p. 732: crust, atmosphere, hydrosphere; Fortescue 2012, pp. 56, 65: biomass
  190. ^ MacKay, MacKay & Henderson 2002, p. 200
  191. ^ Cox 1997, pp. 17, 19
  192. ^ Ostriker & Steinhardt 2001, pp. 46‒53; Zhu 2020, p. 27
  193. ^ Höll et al. 2007
  194. ^ U.S. Geological Survey 2023, p. 78
  195. ^ U.S. Geological Survey 2023, p. 153
  196. ^ Billing Metals & Manufacturing
  197. ^ U.S. Geological Survey 2023, p. 72, 86, 147
  198. ^ U.S. Geological Survey 2023, pp. 32, 34, 78, 82, 84, 91, 156, 158, 160, 170, 176; Kopteva, Kalimullin & Tcvetkov 2021, p. 1; Oztemel, Salt & Salt 2022, p. 5; Neice & Zornow 2016, p. 1268; Howe-Grant 1995, p. 17; Dalakov, Neuber & Herzog R 2020; Boysen, Cristóbal & Hilbig 2020, pp. 475–489; Gardner & Menon 2018, p. 455; Rajarathnam & Assallo 2016, p. 3; Xia, Ning & Zhu 2020
  199. ^ Chand, Kumar & Bhumla 2022, p. 22058
  200. ^ Berger 1997, p. 42
  201. ^ U.S. Geological Survey 1998
  202. ^ a b Boise State University 2020
  203. ^ Hu, Shen & Yu 2017, p. 595
  204. ^ Gardner & Menon 2018, p. 454; U.S. Geological Survey 2023, p. 78
  205. ^ National Institute of Standards and Technology 2013
  206. ^ Bolin 2017, p. 2-1
  207. ^ Emsley 2011, pp. 114–115
  208. ^ Emsley 2011, pp. 363–364
  209. ^ Emsley 2011, pp. 376–377; 379
  210. ^ Emsley 2011, pp. 484–489
  211. ^ Emsley 2011, pp. 511, 516–517
  212. ^ Bhuwalka et al. 2021, pp. 10097–10107
  213. ^ Gaffney & Marley 2017, p. 27
  214. ^ Labinger 2019, p. 305
  215. ^ Emsley 2011, pp. 42–43, 219–220, 263–264, 341, 441–442, 596, 609
  216. ^ Emsley 2011, pp. 84, 128, 180–181, 247
  217. ^ Cook 1923, p. 124
  218. ^ Weeks 1945, p. 161
  219. ^ Emsley 2011, pp. 113, 363, 378, 477, 514–515
  220. ^ Weeks 1945, p. 22; Emsley 2011, p. 40
  221. ^ Lavoisier 1790, p. 175
  222. ^ Jordan 2016
  223. ^ Stillman 1924, p. 213
  224. ^ de L'Aunay 1566, p. 7
  225. ^ Lémery 1699, p. 118; Dejonghe 1998, p. 329
  226. ^ Strathern 2000, p. 239
  227. ^ Criswell p. 1140
  228. ^ Salzberg 1991, p. 204
  229. ^ a b Goldsmith 1982, p. 526
  230. ^ Berzelius 1811, p. 258
  231. ^ Partington 1964, p. 168
  232. ^ a b Bache 1832, p. 250
  233. ^ Roscoe & Schormlemmer 1894, p. 4
  234. ^ Glinka 1959, p. 76
  235. ^ Hérold 2006, pp. 149–150
  236. ^ a b c The Chemical News and Journal of Physical Science 1864
  237. ^ Oxford English Dictionary 1989, "nonmetal"
  238. ^ Kemshead 1875, p. 13
  239. ^ Kendall 1811, pp. 298–303
  240. ^ Brande 1821, p. 5
  241. ^ Edwards & Sienko 1983, pp. 691–96
  242. ^ Edwards & Sienko 1983, p. 693
  243. ^ Herzfeld 1927; Edwards 2000, pp. 100–03
  244. ^ Kubaschewski 1949, pp. 931–940
  245. ^ Remy 1956, p. 9
  246. ^ White 1962, p. 106: It makes a ringing sound when struck.
  247. ^ Johnson 1966, pp. 3–4
  248. ^ Horvath 1973, pp. 335–336
  249. ^ Rao & Ganguly 1986
  250. ^ Smith & Dwyer 1991, p. 65: The difference between melting point and boiling point.
  251. ^ a b Herman 1999, p. 702
  252. ^ 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.
  253. ^ Suresh & Koga 2001, pp. 5940–5944
  254. ^ Johnson 2007, pp. 15–16
  255. ^ a b Edwards 2010, pp. 941–965
  256. ^ Povh & Rosin 2017, p. 131
  257. ^ Beach 1911
  258. ^ Stott 1956, pp. 100–102
  259. ^ Parish 1977, p. 178
  260. ^ Sanderson 1957, p. 229
  261. ^ Hare & Bache 1836, p. 310
  262. ^ Chambers 1743: "That which distinguishes metals from all other bodies ... is their heaviness ..."
  263. ^ Edwards 2000, p. 85
  264. ^ Russell & Lee 2005, p. 466
  265. ^ Atkins et al. 2006, pp. 320–21
  266. ^ Zhigal'skii & Jones 2003, p. 66
  267. ^ Emsley 1971, p. 1
  268. ^ Jones 2010, p. 169
  269. ^ Aylward & Findlay 2008, pp. 6–13; 126
  270. ^ Edelstein & Morrs 2009, p. 123
  271. ^ Arblaster JW (ed.) 2018, p. 269; Lavrukhina & Pozdnyakov 1970, p. 269
  272. ^ Bjerrum 1936
  273. ^ Johnson 1966, pp. 3–5, 15
  274. ^ Oderberg 2007, p. 97
  275. ^ Bertomeu-Sánchez, Garcia-Belmar & Bensaude-Vincent 2002, pp. 248–249
  276. ^ Dupasquier 1844, pp. 66–67
  277. ^ Bache 1832, pp. 248–276
  278. ^ Renouf 1901, pp. 268
  279. ^ Vernon 2020, pp. 217–225
  280. ^ Tregarthen 2003, p. 10
  281. ^ Lewis 1993, pp. 28, 827
  282. ^ Lewis 1993, pp. 28, 813
  283. ^ a b c Rochow 1966, p. 4
  284. ^ Wiberg 2001, p. 780; Emsley 2011, p. 397; Rochow 1966, pp. 23, 84
  285. ^ Kneen, Rogers & Simpson 1972, pp. 321, 404, 436
  286. ^ Kneen, Rogers & Simpson 1972, p. 439
  287. ^ Kneen, Rogers & Simpson 1972, p. 465
  288. ^ Kneen, Rogers & Simpson 1972, p. 308
  289. ^ Wiberg 2001, pp. 505, 681, 781; Glinka 1958, p. 355
  290. ^ Godfrin & Lauter 1995, pp. 216‒218
  291. ^ Wiberg 2001, p. 416
  292. ^ Desai, James & Ho 1984, p. 1160; Matula 1979, p. 1260
  293. ^ Schaefer 1968, p. 76; Carapella 1968, pp. 29‒32
  294. ^ a b c Bogoroditskii & Pasynkov 1967, p. 77; Jenkins & Kawamura 1976, p. 88
  295. ^ Keeler & Wothers 2013, p. 293
  296. ^ Kneen, Rogers & Simpson 1972, p. 264
  297. ^ Rayner-Canham 2018, p. 203
  298. ^ Welcher 2001, p. 3–32: "The elements change from ... metalloids, to moderately active nonmetals, to very active nonmetals, and to a noble gas."
  299. ^ Mackin 2014, p. 80
  300. ^ Johnson 1966, pp. 105–108
  301. ^ Stein 1969, pp. 5396‒5397; Pitzer 1975, pp. 760‒761
  302. ^ Porterfield 1993, p. 336
  303. ^ Rochow 1966, p. 4; Atkins et al. 2006, pp. 8, 122–123
  304. ^ Wiberg 2001, p. 750
  305. ^ Sanderson 1967, p. 172; Mingos 2019, p. 27
  306. ^ House 2008, p. 441
  307. ^ Mingos 2019, p. 27; Sanderson 1967, p. 172
  308. ^ Wiberg 2001, p. 399
  309. ^ Kläning & Appelman 1988, p. 3760
  310. ^ a b Rao 2002, p. 22
  311. ^ Sidorov 1960, pp. 599‒603
  312. ^ McMillan 2006, p. 823
  313. ^ Wells 1984, p. 534
  314. ^ a b Puddephatt & Monaghan 1989, p. 59
  315. ^ King 1995, p. 182
  316. ^ Ritter 2011, p. 10
  317. ^ Yamaguchi & Shirai 1996, p. 3
  318. ^ Vernon 2020, p. 223
  319. ^ Vernon 2020, p. 220
  320. ^ Woodward et al. 1999, p. 134
  321. ^ Dalton 2019
  322. ^ Aylward & Findlay 2008, p. 132
  323. ^ Rahm, Zeng & Hoffmann 2019, p. 345
  324. ^ Aylward & Findlay 2008, p. 126

Bibliography edit

  • Abbott D 1966, An Introduction to the Periodic Table, J. M. Dent & Sons, London
  • Arblaster JW (ed.) 2018, Selected Values of the Crystallographic Properties of Elements, ASM International, Materials Park, Ohio, ISBN 978-1-62708-154-2
  • Atkins PA 2001, The Periodic Kingdom: A Journey Into the Land of the Chemical Elements, Phoenix, London, ISBN 978-1-85799-449-0
  • Atkins PA et al. 2006, Shriver & Atkins' Inorganic Chemistry, 4th ed., Oxford University Press, Oxford, ISBN 978-0-7167-4878-6
  • Atkins PA & Overton T 2010, Shriver & Atkins' Inorganic Chemistry, 5th ed., Oxford University Press, Oxford, ISBN 978-0-19-923617-6
  • Aylward G and Findlay T 2008, SI Chemical Data, 6th ed., John Wiley & Sons Australia, Milton, ISBN 978-0-470-81638-7
  • Bache AD 1832, An essay on chemical nomenclature, prefixed to the treatise on chemistry; by J. J. Berzelius, American Journal of Science, vol. 22, pp. 248–277
  • Bailar JC, Moeller T & Kleinberg J 1965, University Chemistry, DC Heath, Boston
  • Bailar JC et al. 1989, Chemistry, 3rd ed., Harcourt Brace Jovanovich, San Diego, ISBN 978-0-15-506456-0
  • Barton AFM 2021, States of Matter, States of Mind, CRC Press, Boca Raton, ISBN 978-0-7503-0418-4
  • Beach FC (ed.) 1911, The Americana: A universal reference library, vol. XIII, Mel–New, Metalloid, Scientific American Compiling Department, New York
  • Benner SA, Ricardo A & Carrigan MA 2018, "Is there a common chemical model for life in the universe?", in Cleland CE & Bedau MA (eds.), The Nature of Life: Classical and Contemporary Perspectives from Philosophy and Science, Cambridge University Press, Cambridge, ISBN 978-1-108-72206-3
  • Berger LI 1997, Semiconductor Materials, CRC Press, Boca Raton, ISBN 978-0-8493-8912-2
  • Bertomeu-Sánchez JR, Garcia-Belmar A & Bensaude-Vincent B 2002, "Looking for an order of things: Textbooks and chemical classifications in nineteenth century France", Ambix, vol. 49, no. 3, doi:10.1179/amb.2002.49.3.227
  • Berzelius JJ 1811, 'Essai sur la nomenclature chimique', Journal de Physique, de Chimie, d'Histoire Naturelle, vol. LXXIII, pp. 253‒286
  • Bhuwalka et al. 2021, "Characterizing the changes in material use due to vehicle electrification", Environmental Science & Technology, vol. 55, no. 14, pp. 10097–10107, doi:10.1021/acs.est.1c00970
  • Billing Metals & Manufacturing, Silicon, Large Collectors sample. Element 14., accessed May 2, 2023
  • Bjerrum N 1936, Bjerrum's Inorganic Chemistry, Heinemann, London
  • Bodner GM & Pardue HL 1993, Chemistry, An Experimental Science, John Wiley & Sons, New York, ISBN 0-471-59386-9
  • Bogoroditskii NP & Pasynkov VV 1967, Radio and Electronic Materials, Iliffe Books, London
  • Bohlmann R 1992, "Synthesis of halides", in Winterfeldt E (ed.), Heteroatom manipulation, Pergamon Press, Oxford, ISBN 978-0-08-091249-3
  • Boise State University 2020, "Cost-effective manufacturing methods breathe new life into black phosphorus research", Micron School of Materials Science and Engineering, accessed July 9, 2021
  • Bolin P 2017, "Gas-insulated substations", in McDonald JD (ed.), Electric Power Substations Engineering, 3rd, ed., CRC Press, Boca Raton, FL, ISBN 978-1-4398-5638-3
  • Borg RG & Dienes GJ 1992, The Physical Chemistry of Solids, Academic Press, Boston, ISBN 978-0-12-118420-9
  • Boyd R 2011, "Selenium stories", Nature Chemistry, vol. 3, doi:10.1038/nchem.1076
  • Boysen B, Cristóbal J & Hilbig J 2020, "Economic and environmental assessment of water reuse in industrial parks: case study based on a Model Industrial Park", Journal of Water Reuse and Desalination, vol. 10, no. 4, pp. 475–489, doi:10.2166/wrd.2020.034
  • Brady JE & Senese F 2009, Chemistry: The study of Matter and its Changes, 5th ed., John Wiley & Sons, New York, ISBN 978-0-470-57642-7
  • Brande WT 1821, A Manual of Chemistry, vol. II, John Murray, London
  • Brown TL et al. 2014, Chemistry: The Central Science, 3rd ed., Pearson Australia: Sydney, ISBN 978-1-4425-5460-3
  • Burford N, Passmore J & Sanders JCP 1989, "The preparation, structure, and energetics of homopolyatomic cations of groups 16 (the chalcogens) and 17 (the halogens)", in Liebman JF & Greenberg A (eds.), From atoms to polymers: isoelectronic analogies, VCH, New York, ISBN 978-0-89573-711-3
  • Cao C et al. 2021, "Understanding periodic and non-periodic chemistry in periodic tables", Frontiers in Chemistry, vol. 8, no. 813, doi:10.3389/fchem.2020.00813
  • Carapella SC 1968, "Arsenic" in Hampel CA (ed.), The Encyclopedia of the Chemical Elements, Reinhold, New York
  • Carmalt CJ & Norman NC 1998, 'Arsenic, antimony and bismuth: Some general properties and aspects of periodicity', in Norman NC (ed.), Chemistry of Arsenic, Antimony and Bismuth, Blackie Academic & Professional, London, pp. 1–38, ISBN 0-7514-0389-X
  • Challoner J 2014, The Elements: The New Guide to the Building Blocks of our Universe, Carlton Publishing Group, ISBN 978-0-233-00436-5
  • Chambers E 1743, in "Metal", Cyclopedia: Or an Universal Dictionary of Arts and Sciences (etc.), vol. 2, D Midwinter, London
  • Chambers C & Holliday AK 1982, Inorganic Chemistry, Butterworth & Co., London, ISBN 978-0-408-10822-5
  • Chand H, Kumar A & Bhumla P 2022, "Scalable production of ultrathin boron nanosheets from a low-cost precursor", Advanced Materials Interfaces, vol. 9, no. 2, doi:10.1002/admi.202200508
  • Charlier J-C, Gonze X, Michenaud J-P 1994, First-principles study of the stacking effect on the electronic properties of graphite(s), Carbon, vol. 32, no. 2, pp. 289–99, doi:10.1016/0008-6223(94)90192-9
  • Chemical Abstracts Service 2021, CAS REGISTRY database as of November 2, Case #01271182
  • Cherim SM 1971, Chemistry for Laboratory Technicians, Saunders, Philadelphia, ISBN 978-0-7216-2515-7
  • Chung DD 1987, "Review of exfoliated graphite", Journal of Materials Science, vol. 22, doi:10.1007/BF01132008
  • Clugston MJ & Flemming R 2000, Advanced Chemistry, Oxford University Press, Oxford, ISBN 978-0-19-914633-8
  • Cockell C 2019, The Equations of Life: How Physics Shapes Evolution, Atlantic Books, London, ISBN 978-1-78649-304-0
  • Cook CG 1923, Chemistry in Everyday Life: With Laboratory Manual, D Appleton, New York
  • Cotton A et al. 1999, Advanced Inorganic Chemistry, 6th ed., Wiley, New York, ISBN 978-0-471-19957-1
  • Cousins DM, Davidson MG & García-Vivó D 2013, "Unprecedented participation of a four-coordinate hydrogen atom in the cubane core of lithium and sodium phenolates", Chemical Communications, vol. 49, doi:10.1039/C3CC47393G
  • Cox AN (ed.) 2000, Allen's Astrophysical Quantities, 4th ed., AIP Press, New York, ISBN 978-0-387-98746-0
  • Cox PA 1997, The Elements: Their Origins, Abundance, and Distribution, Oxford University Press, Oxford, ISBN 978-0-19-855298-7
  • Cox T 2004, Inorganic Chemistry, 2nd ed., BIOS Scientific Publishers, London, ISBN 978-1-85996-289-3
  • Crawford FH 1968, Introduction to the Science of Physics, Harcourt, Brace & World, New York
  • Crichton R 2012, Biological Inorganic Chemistry: A New Introduction to Molecular Structure and Function, 2nd ed., Elsevier, Amsterdam, ISBN 978-0-444-53783-6
  • Cressey D 2010, "Chemists re-define hydrogen bond", Nature newsblog, accessed August 23, 2017
  • Criswell B 2007, "Mistake of having students be Mendeleev for just a day", Journal of Chemical Education, vol. 84, no. 7, pp. 1140–1144, doi:10.1021/ed084p1140
  • Dalakov P, Neuber E & Herzog R 2020, "Innovative neon refrigeration unit operating down to 30 K", MATEC Web of Conferences, vol. 324, doi:10.1051/matecconf/202032401003
  • Dalton L 2019, "Argon reacts with nickel under pressure-cooker conditions", Chemical & Engineering News, accessed November 6, 2019
  • Daniel PL & Rapp RA 1976, "Halogen corrosion of metals", in Fontana MG & Staehle RW (eds.), Advances in Corrosion Science and Technology, Springer, Boston, doi:10.1007/978-1-4615-9062-0_2
  • de L'Aunay L 1566, Responce au discours de maistre Iacques Grevin, docteur de Paris, qu'il a escript contre le livre de maistre Loys de l'Aunay, medecin en la Rochelle, touchant la faculté de l'antimoine (Response to the Speech of Master Jacques Grévin,... Which He Wrote Against the Book of Master Loys de L'Aunay,... Touching the Faculty of Antimony), De l'Imprimerie de Barthelemi Berton, La Rochelle
  • Dejonghe L 1998, "Zinc–lead deposits of Belgium", Ore Geology Reviews, vol. 12, no. 5, 329–354, doi:10.1016/s0169-1368(98)00007-9
  • Desai PD, James HM & Ho CY 1984, "Electrical resistivity of aluminum and manganese", Journal of Physical and Chemical Reference Data, vol. 13, no. 4, doi:10.1063/1.555725
  • Dingle A 2017, The Elements: An Encyclopedic Tour of the Periodic Table, Quad Books, Brighton, ISBN 978-0-85762-505-2
  • Donohue J 1982, The Structures of the Elements, Robert E. Krieger, Malabar, Florida, ISBN 978-0-89874-230-5
  • 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
  • Dupasquier A 1844, Traité élémentaire de chimie industrielle, Charles Savy Juene, Lyon
  • Earl B & Wilford D 2021, Cambridge O Level Chemistry, Hodder Education, London, ISBN 978-1-3983-1059-9
  • Edelstein NM & Morrs LR 2009, "Chemistry of the actinide elements", in Nagy S (ed.), Radiochemistry and Nuclear Chemistry: Volume II, Encyclopedia of Life Support Systems, EOLSS Publishers, Oxford, pp. 118–176, ISBN 978-1-84826-577-6
  • Edwards PP 2000, "What, why and when is a metal?", in Hall N (ed.), The New Chemistry, Cambridge University, Cambridge, pp. 85–114, ISBN 978-0-521-45224-3
  • Edwards PP et al. 2010, "... a metal conducts and a non-metal doesn’t", Philosophical Transactions of the Royal Society A, 2010, vol, 368, no. 1914, doi:10.1098/rsta.2009.0282
  • 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, doi:10.1021/ed060p691, PMID 25666074
  • Elliot A 1929, "The absorption band spectrum of chlorine", Proceedings of the Royal Society A, vol. 123, no. 792, pp. 629–644, doi:10.1098/rspa.1929.0088
  • Emsley J 1971, The Inorganic Chemistry of the Non-metals, Methuen Educational, London, ISBN 978-0-423-86120-4
  • Emsley J 2011, Nature's Building Blocks: An A–Z Guide to the Elements, Oxford University Press, Oxford, ISBN 978-0-19-850341-5
  • Encyclopædia Britannica 2021, Periodic table, accessed September 21, 2021
  • Evans RC 1966, An Introduction to Crystal Chemistry, 2nd ed., Cambridge University, Cambridge
  • Faraday M 1853, The Subject Matter of a Course of Six Lectures on the Non-metallic Elements, (arranged by John Scoffern), Longman, Brown, Green, and Longmans, London
  • Florez et al. 2022, From the gas phase to the solid state: The chemical bonding in the superheavy element flerovium, The Journal of Chemical Physics, vol. 157, 064304, doi:10.1063/5.0097642
  • Fortescue JAC 2012, Environmental Geochemistry: A Holistic Approach, Springer-Verlag, New York, ISBN 978-1-4612-6047-9
  • Fox M 2010, Optical Properties of Solids, 2nd ed., Oxford University Press, New York, ISBN 978-0-19-957336-3
  • Fraps GS 1913, Principles of Agricultural Chemistry, The Chemical Publishing Company, Easton, PA
  • Fraústo da Silva JJR & Williams RJP 2001, The Biological Chemistry of the Elements: The Inorganic Chemistry of Life, 2nd ed., Oxford University Press, Oxford, ISBN 978-0-19-850848-9
  • Gaffney J & Marley N 2017, General Chemistry for Engineers, Elsevier, Amsterdam, ISBN 978-0-12-810444-6
  • Gardner AJ & Menon DK 2018, "Moving to human trials for argon neuroprotection in neurological injury: A narrative review", British Journal of Anaesthesia, vol. 120, no. 4, pp. 453–468, doi:10.1016/j.bja.2017.10.017
  • Gargaud M et al. (eds.) 2006, Lectures in Astrobiology, vol. 1, part 1: The Early Earth and Other Cosmic Habitats for Life, Springer, Berlin, ISBN 978-3-540-29005-6
  • Glinka N 1958, General chemistry, Sobolev D (trans.), Foreign Languages Publishing House, Moscow
  • Glinka N 1959, General chemistry, Foreign Languages Publishing House, Moscow
  • Godfrin H & Lauter HJ 1995, "Experimental properties of 3He adsorbed on graphite", in Halperin WP (ed.), Progress in Low Temperature Physics, volume 14, Elsevier Science B.V., Amsterdam, ISBN 978-0-08-053993-5
  • Godovikov AA & Nenasheva N 2020, Structural-chemical Systematics of Minerals, 3rd ed., Springer, Cham, Switzerland, ISBN 978-3-319-72877-3
  • Goldsmith RH 1982, 'Metalloids', Journal of Chemical Education, vol. 59, no. 6, pp. 526–527, doi:10.1021/ed059p526
  • Goodrich BG 1844, A Glance at the Physical Sciences, Bradbury, Soden & Co., Boston
  • Government of Canada 2015, Periodic table of the elements, accessed August 30, 2015
  • Greenwood NN & Earnshaw A 2002, Chemistry of the Elements, 2nd ed., Butterworth-Heinemann, ISBN 978-0-7506-3365-9
  • Grochala W 2018, "On the position of helium and neon in the Periodic Table of Elements", Foundations of Chemistry, vol. 20, pp. 191–207, doi:10.1007/s10698-017-9302-7
  • Hampel CA & Hawley GG 1976, Glossary of Chemical Terms, Van Nostrand Reinhold, New York, ISBN 978-0-442-23238-2
  • Hanley JJ & Koga KT 2018, "Halogens in terrestrial and cosmic geochemical systems: Abundances, geochemical behaviours, and analytical methods" in The Role of Halogens in Terrestrial and Extraterrestrial Geochemical Processes: Surface, Crust, and Mantle, Harlov DE & Aranovich L (eds.), Springer, Cham, ISBN 978-3-319-61667-4
  • Hare RA & Bache F 1836, Compendium of the Course of Chemical Instruction in the Medical Department of the University of Pennsylvania, 3rd ed., JG Auner, Philadelphia
  • Hein M & Arena S 2013, Foundations of College Chemistry, John Wiley & Sons, ISBN 978-1-118-29823-7
  • Hengeveld R & Fedonkin MA 2007, "Bootstrapping the energy flow in the beginning of life", Acta Biotheoretica, vol. 55, doi:10.1007/s10441-007-9019-4
  • Herman ZS 1999, "The nature of the chemical bond in metals, alloys, and intermetallic compounds, according to Linus Pauling", in Maksić, ZB, Orville-Thomas WJ (eds.), 1999, Pauling's Legacy: Modern Modelling of the Chemical Bond, Elsevier, Amsterdam, doi:10.1016/S1380-7323(99)80030-2
  • Hermann A, Hoffmann R & Ashcroft NW 2013, "Condensed astatine: Monatomic and metallic", Physical Review Letters, vol. 111, 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, doi:10.1016/j.crci.2005.10.002
  • Herzfeld K 1927, "On atomic properties which make an element a metal", Physical Review, vol. 29, no. 5, doi:10.1103/PhysRev.29.701
  • Hill G & Holman J 2000, Chemistry in Context, 5th ed., Nelson Thornes, Cheltenham, ISBN 0-17-448307-4
  • Hill G, Holman J & Hulme PG 2017, Chemistry in Context, 7th ed., Oxford University Press, Oxford, ISBN 978-0-19-839618-5
  • Holderness A & Berry M 1979, Advanced Level Inorganic Chemistry, 3rd ed., Heinemann Educational Books, London, ISBN 978-0-435-65435-1
  • Höll, Kling & Schroll E 2007, "Metallogenesis of germanium—A review", Ore Geology Reviews, vol. 30, nos. 3–4, pp. 145–180, doi:10.1016/j.oregeorev.2005.07.034
  • Horvath AL 1973, "Critical temperature of elements and the periodic system", Journal of Chemical Education, vol. 50, no. 5, doi:10.1021/ed050p335
  • House JE 2008, Inorganic Chemistry, Elsevier, Amsterdam, ISBN 978-0-12-356786-4
  • Housecroft CE & Sharpe AG 2008, Inorganic Chemistry, 3rd ed., Prentice-Hall, Harlow, ISBN 978-0-13-175553-6
  • Howe-Grant MI (ed.) 1995, Fluorine Chemistry: A Comprehensive Treatment, John Wiley and Sons, New York, p. 17, ISBN 978-0-471-12031-5
  • Hu Z, Shen Z & Yu JC 2017, "Phosphorus containing materials for photocatalytic hydrogen evolution", Green Chemistry, vol. 19, no. 3, pp. 588–613, doi:10.1039/C6GC02825J
  • Hurlbut Jr CS 1961, Manual of Mineralogy, 15th ed., John Wiley & Sons, New York
  • Hussain et al. 2023, "Tuning the electronic properties of molybdenum di-sulphide monolayers via doping using first-principles calculations", Physica Scripta, vol. 98, no. 2, doi:10.1088/1402-4896/acacd1
  • IUPAC Periodic Table of the Elements, accessed October 11, 2021
  • Janas D, Cabrero-Vilatela, A & Bulmer J 2013, "Carbon nanotube wires for high-temperature performance", Carbon, vol. 64, pp. 305–314, doi:10.1016/j.carbon.2013.07.067
  • Jenkins GM & Kawamura K 1976, Polymeric Carbons—Carbon Fibre, Glass and Char, Cambridge University Press, Cambridge, ISBN 978-0-521-20693-8
  • Jentzsch AV & Matile S 2015, "Anion transport with halogen bonds", in Metrangolo P & Resnati G (eds.), Halogen Bonding I: Impact on Materials Chemistry and Life Sciences, Springer, Cham, ISBN 978-3-319-14057-5
  • Johnson D (ed.) 2007, Metals and Chemical Change, RSC Publishing, Cambridge, ISBN 978-0-85404-665-2
  • Johnson RC 1966, Introductory Descriptive Chemistry, WA Benjamin, New York
  • 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 978-0-521-19436-5
  • Jordan JM 2016 " 'Ancient episteme' and the nature of fossils: a correction of a modern scholarly error", History and Philosophy of the Life Sciences, vol. 38, no, 1, pp. 90–116, doi:10.1007/s40656-015-0094-6
  • Kaiho T 2017, Iodine Made Simple, CRC Press, e-book, doi:10.1201/9781315158310
  • Keeler J & Wothers P 2013, Chemical Structure and Reactivity: An Integrated Approach, Oxford University Press, Oxford, ISBN 978-0-19-960413-5
  • Kemshead WB 1875, Inorganic chemistry, William Collins, Sons, & Company, London
  • Kendall EA 1811, Pocket Encyclopædia, 2nd ed., vol. III, Longman, Hurst, Rees, Orme, and Co., London
  • Kernion MC & Mascetta JA 2019, Chemistry: The Easy Way, 6th ed., Kaplan, New York, ISBN 978-1-4380-1210-0
  • Khan N 2001, An Introduction to Physical Geography, Concept Publishing, New Delhi, ISBN 978-81-7022-898-1
  • King RB 1994, Encyclopedia of Inorganic Chemistry, vol. 3, John Wiley & Sons, New York, ISBN 978-0-471-93620-6
  • King RB 1995, Inorganic Chemistry of Main Group Elements, VCH, New York, ISBN 978-1-56081-679-9
  • King GB & Caldwell WE 1954, The Fundamentals of College Chemistry, American Book Company, New York
  • Kläning UK & Appelman EH 1988, "Protolytic properties of perxenic acid", Inorganic Chemistry, vol. 27, no. 21, doi:10.1021/ic00294a018
  • Kneen WR, Rogers MJW & Simpson P 1972, Chemistry: Facts, Patterns, and Principles, Addison-Wesley, London, ISBN 978-0-201-03779-1
  • Knight J 2002, Science of Everyday Things: Real-life chemistry, Gale Group, Detroit, ISBN 9780787656324
  • Koenig SH 1962, in Proceedings of the International Conference on the Physics of Semiconductors, held at Exeter, July 16−20, 1962, The Institute of Physics and the Physical Society, London
  • Kopteva A, Kalimullin L & Tcvetkov P 2021, "Prospects and obstacles for green hydrogen production in Russia", Energies, vol. 14, no. 3, pp. 1–21, doi:10.3390/en14030718
  • Kosanke et al. 2012, Encyclopedic Dictionary of Pyrotechnics (and Related Subjects), Part 3 – P to Z, Pyrotechnic Reference Series No. 5, Journal of Pyrotechnics, Whitewater, Colorado, ISBN 978-1-889526-21-8
  • Kubaschewski O 1949, "The change of entropy, volume and binding state of the elements on melting", Transactions of the Faraday Society, vol. 45, doi:10.1039/TF9494500931
  • Kugler HK & Keller C (eds) 1985, Gmelin Handbook of Inorganic and Organometallic chemistry, 8th ed., 'At, Astatine', system no. 8a, Springer-Verlag, Berlin, ISBN 3-540-93516-9
  • Labinger JA 2019, "The history (and pre-history) of the discovery and chemistry of the noble gases", in Giunta CJ, Mainz VV & Girolami GS (eds.), 150 Years of the Periodic Table: A Commemorative Symposium, Springer Nature, Cham, Switzerland, ISBN 978-3-030-67910-1
  • Lanford OE 1959, Using Chemistry, McGraw-Hill, New York
  • Langley RH & Hattori H 2014, 1,001 Practice Problems: Chemistry For Dummies, John Wiley & Sons, Hoboken, NJ, ISBN 978-1-118-54932-2
  • Larrañaga MD, Lewis RJ & Lewis RA 2016, Hawley's Condensed Chemical Dictionary, 16th ed., Wiley, Hoboken, New York, ISBN 978-1-118-13515-0
  • Lavoisier A 1790, Elements of Chemistry, R Kerr (trans.), William Creech, Edinburgh
  • Lavrukhina AK & Pozdnyakov AA 1970, Analytical Chemistry of Technetium, Promethium, Astatine, and Francium, R Kondor, trans., Ann Arbor–Humphrey Science Publishers, Ann Arbor, ISBN 978-0-250-39923-9
  • Lee JD 1996, Concise Inorganic Chemistry, 5th ed., Blackwell Science, Oxford, ISBN 978-0-632-05293-6
  • Lémery N 1714, Traité universel des drogues simples, mises en ordre alphabetique, L d'Houry, Paris, p. 118
  • Lewis RJ 1993, Hawley's Condensed Chemical Dictionary, 12th ed., Van Nostrand Reinhold, New York, ISBN 978-0-442-01131-4
  • Lidin RA 1996, Inorganic Substances Handbook, Begell House, New York, ISBN 978-0-8493-0485-9
  • Liptrot GF 1983, Modern Inorganic Chemistry, 4th ed., Bell & Hyman, ISBN 978-0-7135-1357-8
  • Los Alamos National Laboratory 2021, Periodic Table of Elements: A Resource for Elementary, Middle School, and High School Students, accessed September 19, 2021
  • Luchinskii GP & Trifonov DN 1981, "Some problems of chemical elements classification and the structure of the periodic system", in Uchenie o Periodichnosti. Istoriya i Sovremennoct, (Russian) Nauka, Moscow
  • MacKay KM, MacKay RA & Henderson W 2002, Introduction to Modern Inorganic Chemistry, 6th ed., Nelson Thornes, Cheltenham, ISBN 978-0-7487-6420-4
  • Mackin M 2014, Study Guide to Accompany Basics for Chemistry, Elsevier Science, Saint Louis, ISBN 978-0-323-14652-4
  • Maosheng M 2020, "Noble gases in solid compounds show a rich display of chemistry with enough pressure", Frontiers in Chemistry, vol. 8, doi:10.3389/fchem.2020.570492
  • Massey AG 2000, Main Group Chemistry, 2nd ed., John Wiley & Sons, Chichester, ISBN 978-0-471-49039-5
  • Masterton W, Hurley C & Neth E 2011, Chemistry: Principles and Reactions, 7th ed., Brooks/Cole, Belmont, California, ISBN 978-1-111-42710-8
  • Matson M & Orbaek AW 2013, Inorganic Chemistry for Dummies, John Wiley & Sons: Hoboken, ISBN 978-1-118-21794-8
  • Matula RA 1979, "Electrical resistivity of copper, gold, palladium, and silver", Journal of Physical and Chemical Reference Data, vol. 8, no. 4, doi:10.1063/1.555614
  • Mazej Z 2020, "Noble-gas chemistry more than half a century after the first report of the noble-gas compound", Molecules, vol. 25, no. 13, doi:10.3390/molecules25133014, PMID 32630333, PMC 7412050
  • McCue JJ 1963, World of Atoms: An Introduction to Physical Science, Ronald Press, New York
  • McMillan P 2006, "A glass of carbon dioxide", Nature, vol. 441, doi:10.1038/441823a
  • Mee AJ 1964, Physical Chemistry, Aldine Publishing, Chicago
  • Messler Jr RW 2011, The Essence of Materials for Engineers, Jones and Bartlett Learning, Sudbury, Massachusetts, ISBN 978-0-7637-7833-0
  • Mewes et al. 2019, Copernicium: A relativistic noble liquid, Angewandte Chemie International Edition, vol. 58, pp. 17964–17968, doi:10.1002/anie.201906966
  • Mingos DMP 2019, "The discovery of the elements in the Periodic Table", in Mingos DMP (ed.), The Periodic Table I. Structure and Bonding, Springer Nature, Cham, doi:10.1007/978-3-030-40025-5
  • Moeller T et al. 2012, Chemistry: With Inorganic Qualitative Analysis, Academic Press, New York, ISBN 978-0-12-503350-3
  • Möller D 2003, Luft: Chemie, Physik, Biologie, Reinhaltung, Recht, Walter de Gruyter, Berlin, ISBN 978-3-11-016431-2
  • Moody B 1991, Comparative Inorganic Chemistry, 3rd ed., Edward Arnold, London, ISBN 978-0-7131-3679-1
  • Moore JT 2016, Chemistry for Dummies, 2nd ed., ch. 16, Tracking periodic trends, John Wiley & Sons: Hoboken, ISBN 978-1-119-29728-4
  • Morita A 1986, 'Semiconducting black phosphorus', Journal of Applied Physics A, vol. 39, no. 4, pp. 227–42, doi:10.1007/BF00617267
  • Morely HF & Muir MM 1892, Watt's Dictionary of Chemistry, vol. 3, Longman's Green, and Co., London
  • Moss, TS 1952, Photoconductivity in the Elements, Butterworths Scientific, London
  • Nakao Y 1992, "Dissolution of noble metals in halogen–halide–polar organic solvent systems", Journal of the Chemical Society, Chemical Communications, no. 5, doi:10.1039/C39920000426
  • National Center for Biotechnology Information 2021, "PubChem compound summary for CID 402, Hydrogen sulfide", accessed August 31, 2021
  • National Institute of Standards and Technology 2013, SRM 4972 – Radon-222 Emanation Standard, accessed August 1, 2021
  • Nefedov VD et al. 1968, 'Astatine', Russian Chemical Reviews, vol. 37, no. 2, pp. 87–98, doi:10.1070/rc1968v037n02abeh0
  • Neice AE & Zornow MH 2016, "Editorial: Xenon anaesthesia for all, or only a select few?", Anaesthesia, vol. 71, no. 11, pp. 1259–1272 (1268), doi:10.1111/anae.13569
  • Nelson PG 1987, "Important elements", Journal of Chemical Education, vol. 68, no. 9, doi:10.1021/ed068p732
  • Oderberg DS 2007, Real Essentialism, Routledge, New York, ISBN 978-1-134-34885-5
  • Ostriker JP & Steinhardt PJ 2001, "The quintessential universe", Scientific American, vol. 284, no. 1, pp. 46–53 PMID 11132422, doi:10.1038/scientificamerican0101-46
  • Oxtoby DW, Gillis HP & Butler LJ 2015, Principles of Modern Chemistry, 8th ed., Cengage Learning, Boston, ISBN 978-1-305-07911-3
  • Oztemel BH, Salt I, Salt Y 2022, "Carbon dioxide utilization: Process simulation of synthetic fuel production from flue gases", Chemical Industry and Chemical Engineering Quarterly, vol. 28, no. 4, doi:10.2298/CICEQ211025005B
  • Parameswaran P et al. 2020, "Phase evolution and characterization of mechanically alloyed hexanary Al16.6Mg16.6Ni16.6Cr16.6Ti16.6Mn16.6 high entropy alloy", Metal Powder Report, vol. 75, no. 4, doi:10.1016/j.mprp.2019.08.001
  • Parish RV 1977, The Metallic Elements, Longman, London, ISBN 978-0-582-44278-8
  • Parkes GD & Mellor JW 1943, Mellor's Nodern Inorganic Chemistry, Longmans, Green and Co., London
  • Partington JR 1944, A Text-book of Inorganic Chemistry, 5th ed., Macmillan & Co., London
  • Partington JR 1964, A history of chemistry, vol. 4, Macmillan, London
  • Pascoe KJ 1982, An Introduction to the Properties of Engineering Materials, 3rd ed., Von Nostrand Reinhold (UK), Wokingham, Berkshire, ISBN 978-0-442-30233-7
  • Petruševski VM & Cvetković J 2018, "On the 'true position' of hydrogen in the Periodic Table", Foundations of Chemistry, vol. 20, pp. 251–260, doi:10.1007/s10698-018-9306-y
  • Phillips CSG & Williams RJP 1965, Inorganic Chemistry, vol. 1, Principles and non-metals, Clarendon Press, Oxford
  • Phillips JC 1973, "The chemical structure of solids", in Hannay NB (ed.), Treatise on Solid State Chemistry, vol. 1, Plenum Press, New York, pp. 1–42, ISBN 978-1-4684-2663-2
  • Pitzer K 1975, "Fluorides of radon and elements 118", Journal of the Chemical Society, Chemical Communications, no. 18, doi:10.1039/C3975000760B
  • Porterfield WW 1993, Inorganic Chemistry, Academic Press, San Diego, ISBN 978-0-12-562980-5
  • Povh B & Rosina M 2017, Scattering and Structures: Essentials and Analogies in Quantum Physics, 2nd ed., Springer, Berlin, doi:10.1007/978-3-662-54515-7
  • Powell P & Timms P 1974, The Chemistry of the Non-Metals, Chapman and Hall, London, ISBN 978-0-412-12200-2
  • Puddephatt RJ & Monaghan PK 1989, The Periodic Table of the Elements, 2nd ed., Clarendon Press, Oxford, ISBN 978-0-19-855516-2
  • Rahm M, Zeng T & Hoffmann R 2019, "Electronegativity seen as the ground-state average valence electron binding energy", Journal of the American Chemical Society, vol. 141, no. 1, pp. 342−351, doi:10.1021/jacs.8b10246
  • Rajarathnam GP & Assallo AM 2016, The Zinc/bromine Flow Battery: Materials Challenges and Practical Solutions for Technology Advancement, Springer, Singapore, p. 3, ISBN 978-981-287-645-4
  • Rao KY 2002, Structural Chemistry of Glasses, Elsevier, Oxford, ISBN 978-0-08-043958-7
  • Rao CNR & Ganguly PA 1986, "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
  • Rayner-Canham G 2018, "Organizing the transition metals", in Scerri E & Restrepo G (Ed's.) Mendeleev to Oganesson: A multidisciplinary perspective on the periodic table, Oxford University, New York, ISBN 978-0-190-668532
  • Rayner-Canham G 2020, The Periodic Table: Past, Present and Future, World Scientific, New Jersey, ISBN 978-981-121-850-7
  • Regnault MV 1853, Elements of Chemistry, vol. 1, 2nd ed., Clark & Hesser, Philadelphia
  • Reilly C 2002, Metal Contamination of Food, Blackwell Science, Oxford, ISBN 978-0-632-05927-0
  • Remy H 1956, Treatise on Inorganic Chemistry, Anderson JS (trans.), Kleinberg J (ed.), vol. II, Elsevier, Amsterdam
  • Renouf E 1901, "Lehrbuch der Anorganischen Chemie", Science, vol. 13, no. 320, doi:10.1126/science.13.320.268
  • Restrepo G, Llanos EJ & Mesa H 2006, "Topological space of the chemical elements and its properties", Journal of Mathematical Chemistry, vol. 39, doi:10.1007/s10910-005-9041-1
  • Ritter SK 2011, "The case of the missing xenon", Chemical & Engineering News, vol. 89, no. 9, ISSN 0009-2347
  • Rochow EG 1966, The Metalloids, DC Heath and Company, Boston
  • Rochow EG 1973, "Silicon", in Bailar JC et al. (eds.), Comprehensive Inorganic Chemistry, vol. 1, Pergamon Press, Oxford, ISBN 978-0-08-015655-2
  • Rodgers GE 2012, Descriptive Inorganic, Coordination, and Solid State Chemistry, 3rd ed., Brooks/Cole, Belmont, California, ISBN 978-0-8400-6846-0
  • Roscoe HE & Schorlemmer FRS 1894, A treatise on chemistry: Volume II: The metals, D Appleton, New York
  • Royal Society of Chemistry 2021, Periodic Table: Non-metal, accessed September 3, 2021
  • Rudolph J 1973, Chemistry for the Modern Mind, Macmillan, New York
  • Russell AM & Lee KL 2005, Structure-Property Relations in Nonferrous Metals, Wiley-Interscience, New York, ISBN 0-471-64952-X
  • Salinas JT 2019 Exploring Physical Science in the Laboratory, Moreton Publishing, Englewood, Colorado, ISBN 978-1-61731-753-8
  • Salzberg HW 1991, From Caveman to Chemist: Circumstances and Achievements, American Chemical Society, Washington, DC, ISBN 0-8412-1786-6
  • Sanderson RT 1957, "An electronic distinction between metals and nonmetals", Journal of Chemical Education, vol. 34, no. 5, doi:10.1021/ed034p229
  • Sanderson RT 1967, Inorganic Chemistry, Reinhold, New York
  • Scerri E (ed.) 2013, 30-Second Elements: The 50 Most Significant Elements, Each Explained In Half a Minute, Ivy Press, London, ISBN 978-1-84831-616-4
  • Scerri E 2020, The Periodic Table: Its Story and Its Significance, Oxford University Press, New York, ISBN 978-0-19091-436-3
  • Schaefer JC 1968, "Boron" in Hampel CA (ed.), The Encyclopedia of the Chemical Elements, Reinhold, New York
  • Schlager N & Lauer J (eds.) 2000, Science and Its Times: 1700–1799, volume 4 of Science and Its Times: Understanding the Social Significance of Scientific Discovery, Gale Group, ISBN 978-0-7876-3932-7
  • Schmedt auf der Günne J, Mangstl M & Kraus F 2012, "Occurrence of difluorine F2 in nature—In situ proof and quantification by NMR spectroscopy", Angewandte Chemie International Edition, vol. 51, no. 31, doi:10.1002/anie.201203515
  • Scott D 2014, Around the World in 18 Elements, Royal Society of Chemistry, e-book, ISBN 978-1-78262-509-4
  • Scott EC & Kanda FA 1962, The Nature of Atoms and Molecules: A General Chemistry, Harper & Row, New York
  • Seese WS & Daub GH 1985, Basic Chemistry, 4th ed., Prentice-Hall, Englewood Cliffs, NJ, ISBN 978-0-13-057811-2
  • Segal BG 1989, Chemistry: Experiment and Theory, 2nd ed., John Wiley & Sons, New York, ISBN 0-471-84929-4
  • Shanabrook BV, Lannin JS & Hisatsune IC 1981, "Inelastic light scattering in a onefold-coordinated amorphous semiconductor", Physical Review Letters, vol. 46, no. 2, 12 January, doi:10.1103/PhysRevLett.46.130
  • Shang et al. 2021, "Ultrahard bulk amorphous carbon from collapsed fullerene", Nature, vol. 599, pp. 599–604, doi:10.1038/s41586-021-03882-9
  • Shkol’nikov EV 2010, "Thermodynamic characterization of the amphoterism of oxides M2O3 (M = As, Sb, Bi) and their hydrates in aqueous media, Russian Journal of Applied Chemistry, vol. 83, no. 12, pp. 2121–2127, doi:10.1134/S1070427210120104
  • Sidorov TA 1960, "The connection between structural oxides and their tendency to glass formation", Glass and Ceramics, vol. 17, no. 11, doi:10.1007BF00670116
  • Siekierski S & Burgess J 2002, Concise Chemistry of the Elements, Horwood Press, Chichester, ISBN 978-1-898563-71-6
  • Smith A & Dwyer C 1991, Key Chemistry: Investigating Chemistry in the Contemporary World: Book 1: Materials and Everyday Life, Melbourne University Press, Carlton, Victoria, ISBN 978-0-522-84450-4
  • Smits et al. 2020, "Oganesson: A noble gas element that is neither noble nor a gas", Angewandte Chemie International Edition, vol. 59, pp. 23636–23640, doi:10.1002/anie.202011976
  • Stein L 1969, "Oxidized radon in halogen fluoride solutions", Journal of the American Chemical Society, vol. 19, no. 19, doi:10.1021/ja01047a042
  • Stein L 1983, "The chemistry of radon", Radiochimica Acta, vol. 32, doi:10.1524/ract.1983.32.13.163
  • Stellman JM (ed.) 1998, Encyclopaedia of Occupational Health and Safety, vol. 4, 4th ed., International Labour Office, Geneva, ISBN 978-92-2-109817-1
  • Steudel R 1977, Chemistry of the Non-metals: With an Introduction to atomic Structure and Chemical Bonding, Walter de Gruyter, Berlin, ISBN 978-3-11-004882-7
  • Steudel R & Eckert B 2003, "Solid sulfur allotropes", in Steudel R (ed.), Elemental Sulfur and Sulfur-rich Compounds I, Springer-Verlag, Berlin, ISBN 978-3-540-40191-9
  • Steudel R 2020, Chemistry of the Non-metals: Syntheses - Structures - Bonding - Applications, in collaboration with D Scheschkewitz, Berlin, Walter de Gruyter, doi:10.1515/9783110578065
  • Still B 2016 The Secret Life of the Periodic Table, Cassell, London, ISBN 978-1-84403-885-5
  • Stillman JM 1924, The Story of Early Chemistry, D. Appleton, New York
  • Stott RWA 1956, Companion to Physical and Inorganic Chemistry, Longmans, Green and Co, London
  • Stuke J 1974, "Optical and electrical properties of selenium", in Zingaro RA & Cooper WC (eds.), Selenium, Van Nostrand Reinhold, New York, pp. 174
  • Strathern P 2000, Mendeleyev's dream: The Quest for the Elements, Hamish Hamilton, London, ISBN 978-0-8412-1786-7
  • Suresh CH & Koga NA 2001, "A consistent approach toward atomic radii”, Journal of Physical Chemistry A, vol. 105, no. 24. doi:10.1021/jp010432b
  • Tang et al. 2021, "Synthesis of paracrystalline diamond", Nature, vol. 599, pp. 605–610, doi:10.1038/s41586-021-04122-w
  • 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. 867–7, doi:10.1016/0038-1098(84)90441-1
  • Taylor MD 1960, First Principles of Chemistry, Van Nostrand, Princeton
  • The Chemical News and Journal of Physical Science 1864, "Notices of books: Manual of the Metalloids", vol. 9, p. 22
  • The Chemical News and Journal of Physical Science 1897, "Notices of books: A Manual of Chemistry, Theoretical and Practical", by WA Tilden", vol. 75, pp. 188–189
  • Thompson M 2004, Osmium tetroxide (OsO4), Molecule of the Month, (May), doi:10.6084/m9.figshare.5437084
  • Thornton BF & Burdette SC 2010, "Finding eka-iodine: Discovery priority in modern times", Bulletin for the history of chemistry, vol. 35, no. 2, accessed September 14, 2021
  • Tidy CM 1887, Handbook of Modern Chemistry, 2nd ed., Smith, Elder & Co., London
  • Tregarthen L 2003, Preliminary Chemistry, Macmillan Education: Melbourne, ISBN 978-0-7329-9011-4
  • Trenberth KE & Smith L 2005, "The mass of the atmosphere: A constraint on global analyses", Journal of Climate, vol. 18, no. 6, doi:10.1175/JCLI-3299.1
  • Tshitoyan et al. 2019, "Unsupervised word embeddings capture latent knowledge from materials science literature", Nature, vol. 571, doi:10.1038/s41586-019-1335-8
  • Tyler PM 1948, From the Ground Up: Facts and Figures of the Mineral Industries of the United States, McGraw-Hill, New York
  • U.S. Geological Survey 1998, Mineral Commodity Summaries, U.S. Geological Survey, accessed 27 August 2023
  • U.S. Geological Survey 2023, Mineral Commodity Summaries, U.S. Geological Survey
  • Vasáros L & Berei K 1985, 'General Properties of Astatine', pp. 107–28, in Kugler & Keller
  • Vassilakis AA, Kalemos A & Mavridis A 2014, "Accurate first principles calculations on chlorine fluoride ClF and its ions ClF±", Theoretical Chemistry Accounts, vol. 133, no. 1436, doi:10.1007/s00214-013-1436-7
  • Vernon R 2013, "Which elements are metalloids?", Journal of Chemical Education, vol. 90, no. 12, 1703‒1707, doi:10.1021/ed3008457
  • Vernon R 2020, "Organising the metals and nonmetals", Foundations of Chemistry, vol. 22, doi:10.1007/s10698-020-09356-6 (open access)
  • Wächtershäuser G 2014, "From chemical invariance to genetic variability", in Weigand W and Schollhammer P (eds.), Bioinspired Catalysis: Metal Sulfur Complexes, Wiley-VCH, Weinheim, doi:10.1002/9783527664160.ch1
  • Wakeman TH 1899, "Free thought—Past, present and future", Free Thought Magazine, vol. 17
  • Wasewar KL 2021, "Intensifying approaches for removal of selenium", in Devi et al. (eds.), Selenium contamination in water, John Wiley & Sons, Hoboken, pp. 319–355, ISBN 978-1-119-69354-3
  • Weeks ME 1945, Discovery of the Elements, 5th ed., Journal of Chemical Education, Easton, Pennsylvania
  • Welcher SH 2001, High marks: Regents Chemistry Made Easy, 2nd ed., High Marks Made Easy, New York, ISBN 978-0-9714662-4-1
  • Wells AF 1984, Structural Inorganic Chemistry, 5th ed., Clarendon Press, Oxford, ISBN 978-0-19-855370-0
  • White JH 1962, Inorganic Chemistry: Advanced and Scholarship Levels, University of London Press, London
  • Wibaut P 1951, Organic Chemistry, Elsevier Publishing Company, New York
  • Wiberg N 2001, Inorganic Chemistry, Academic Press, San Diego, ISBN 978-0-12-352651-9
  • Williams RPJ 2007, "Life, the environment and our ecosystem", Journal of Inorganic Biochemistry, vol. 101, nos. 11–12, doi:10.1016/j.jinorgbio.2007.07.006
  • Woodward et al. 1999, "The electronic structure of metal oxides", In Fierro JLG (ed.), Metal Oxides: Chemistry and Applications, CRC Press, Boca Raton, ISBN 1-4200-2812-X
  • Wulfsberg G 1987, Principles of Descriptive Chemistry, Brooks/Cole, Belmont CA, ISBN 978-0-534-07494-4
  • Wulfsberg G 2000, Inorganic Chemistry, University Science Books, Sausalito, California, ISBN 978-1-891389-01-6
  • Xia G-J, Ning Z-X & Zhu X-M 2020, “Effect of low-frequency oscillation on plasma focusing in krypton hall thruster", Journal of Propulsion and Power, vol. 36, no. 1, pp. Journal of Propulsion and Power, doi:10.2514/1.B37599
  • Yamaguchi M & Shirai Y 1996, "Defect structures", in Stoloff NS & Sikka VK (eds.), Physical Metallurgy and Processing of Intermetallic Compounds, Chapman & Hall, New York, ISBN 978-1-4613-1215-4
  • Yang J 2004, "Theory of thermal conductivity", in Tritt TM (ed.), Thermal Conductivity: Theory, Properties, and Applications, Kluwer Academic/Plenum Publishers, New York, pp. 1–20, ISBN 978-0-306-48327-1,
  • Yoder CH, Suydam FH & Snavely FA 1975, Chemistry, 2nd ed, Harcourt Brace Jovanovich, New York, ISBN 978-0-15-506470-6
  • Young JA 2006, "Iodine", Journal of Chemical Education, vol. 83, no. 9, doi:10.1021/ed083p1285
  • Young et al. 2018, General Chemistry: Atoms First, Cengage Learning: Boston, ISBN 978-1-337-61229-6
  • Zhao J, Tu Z & Chan SH 2021, "Carbon corrosion mechanism and mitigation strategies in a proton exchange membrane fuel cell (PEMFC): A review", Journal of Power Sources, vol. 488, #229434, doi:10.1016/j.jpowsour.2020.229434
  • Zhigal'skii GP & Jones BK 2003, The Physical Properties of Thin Metal Films, Taylor & Francis, London, ISBN 978-0-415-28390-8
  • Zhu W 2020, Chemical Elements In Life, World Scientific, Singapore, ISBN 978-981-121-032-7
  • Zhu et al. 2014, "Reactions of xenon with iron and nickel are predicted in the Earth's inner core", Nature Chemistry, vol. 6, doi:10.1038/nchem.1925, PMID 24950336
  • Zumdahl SS & DeCoste DJ 2010, Introductory Chemistry: A Foundation, 7th ed., Cengage Learning, Mason, Ohio, ISBN 978-1-111-29601-8