Download as pdf or txt
Download as pdf or txt
You are on page 1of 217

Nonmetal

Nonmetals (and metalloids) in the periodic table:

   Metalloid    Reactive nonmetal    Noble gas


Metalloids are included in the legend as they generally
behave chemically like nonmetals and are sometimes
counted as such.

Apart from hydrogen, nonmetals are located in the p-


block. Helium, as an s-block element, would normally
be placed next to hydrogen and above beryllium.

However, since it is a noble gas, it is instead placed


above neon (in the p-block).

In chemistry, a nonmetal (or non-metal) is


a chemical element that mostly lacks the
characteristics of a metal. Physically, a
nonmetal tends to have a relatively low
melting point, boiling point, and density. A
nonmetal is typically brittle when solid and
usually has poor thermal conductivity and
electrical conductivity. Chemically,
nonmetals tend to have relatively high
ionization energy, electron affinity, and
electronegativity. They gain or share
electrons when they react with other
elements and chemical compounds.
Seventeen elements are generally
classified as nonmetals: most are gases
(hydrogen, helium, nitrogen, oxygen,
fluorine, neon, chlorine, argon, krypton,
xenon and radon); one is a liquid
(bromine); and a few are solids (carbon,
phosphorus, sulfur, selenium, and iodine).
Metalloids such as boron, silicon, and
germanium are sometimes counted as
nonmetals.

The nonmetals are divided into two


categories reflecting their relative
propensity to form chemical compounds:
reactive nonmetals and noble gases. The
reactive nonmetals vary in their
nonmetallic character. The less
electronegative of them, such as carbon
and sulfur, mostly have weak to
moderately strong nonmetallic properties
and tend to form covalent compounds
with metals. The more electronegative of
the reactive nonmetals, such as oxygen
and fluorine, are characterised by stronger
nonmetallic properties and a tendency to
form predominantly ionic compounds with
metals. The noble gases are distinguished
by their great reluctance to form
compounds with other elements.

The distinction between categories is not


absolute. Boundary overlaps, including
with the metalloids, occur as outlying
elements in each category show or begin
to show less-distinct, hybrid-like, or
atypical properties.

Although five times more elements are


metals than nonmetals, two of the
nonmetals—hydrogen and helium—make
up over 99 percent of the observable
universe.[1] Another nonmetal, oxygen,
makes up almost half of the Earth's crust,
oceans, and atmosphere.[2] Living
organisms are composed almost entirely
of nonmetals: hydrogen, oxygen, carbon,
and nitrogen.[3] Nonmetals form many
more compounds than metals.[4]

Definition and applicable


elements
There is Nonmetals in the periodic table
no
rigorous
definition of a nonmetal. Broadly, any
element lacking a preponderance of
metallic properties can be regarded as a
nonmetal.

The elements generally classified as


nonmetals include one element in group 1
(hydrogen); one in group 14 (carbon); two
in group 15 (nitrogen and phosphorus);
three in group 16 (oxygen, sulfur and
selenium); most of group 17 (fluorine,
chlorine, bromine and iodine); and all of
group 18 (with the possible exception of
oganesson).
As there is no widely agreed definition of a
nonmetal, elements in the periodic table
vicinity of where the metals meet the
nonmetals are inconsistently classified by
different authors. Elements sometimes
also classified as nonmetals are the
metalloids boron (B), silicon (Si),
germanium (Ge), arsenic (As), antimony
(Sb), tellurium (Te), and astatine (At).[5]
The nonmetal selenium (Se) is sometimes
instead classified as a metalloid,
particularly in environmental chemistry.[6]
Properties

“ The marvelous variety and infinite subtlety of


the non-metallic elements, their compounds,
structures and reactions, is not sufficiently
acknowledged in the current teaching of
chemistry. ”

JJ Zuckerman and FC Nachod


In Steudel's Chemistry of the non-metals
(1977, preface)

Nonmetals show more variability in their


properties than do metals.[7] These
properties are largely determined by the
interatomic bonding strengths and
molecular structures of the nonmetals
involved, both of which are subject to
variation as the number of valence
electrons in each nonmetal varies. Metals,
in contrast, have more homogenous
structures and their properties are more
easily reconciled.[8]

Physically, they largely exist as diatomic or


monatomic gases, with the remainder
having more substantial (open-packed)
forms, unlike metals, which are nearly all
solid and close-packed. If solid, they have
a submetallic appearance (with the
exception of sulfur) and are mostly brittle,
as opposed to metals, which are lustrous,
and generally ductile or malleable; they
usually have lower densities than metals;
are mostly poorer conductors of heat and
electricity; and tend to have significantly
lower melting points and boiling points
than those of metals.
Scatter plot of electronegativity values and
standard electrode potentials of chemically active

nonmetallic elements, showing a rough correlation


between the two properties. The higher the
standard electrode potential, the greater is the
capacity to act as an oxidizing agent.[9] The chart
shows that oxygen and the nonmetallic halogens
are the strongest oxidising agents and that, for the
most part the elements commonly recognised as
most part, the elements commonly recognised as
metalloids are the weakest. The electrode
potentials are for the reduction of the elements to
monatomic anions (X→X−; X = F, Cl, Br, I, or H) or to
their protonated forms (for example, O2→H2O;
N2→NH3.[10]
Hydrogen and nitrogen have anomalous standard
electrode potentials due to their reluctance to form
anions.
A broad progression in nonmetallic character is
seen, with the metalloids in the lower left, and

oxygen and the nonmetallic halogens in the upper


right.
Trend lines are shown with and without the
anomalous hydrogen and nitrogen values. The R2
values show how close each trend line fits its data
points. Values range from 0.0 (indicating no fit) and
( fi)
1.0 (a very good fit).

Chemically the nonmetals mostly have


high ionisation energies, high electron
affinities (nitrogen and the noble gases
have negative electron affinities) and high
electronegativity values[n 1] noting that, in
general, the higher an element's ionisation
energy, electron affinity, and
electronegativity, the more nonmetallic
that element is.[11] Nonmetals (including –
to a limited extent – xenon and probably
radon) usually exist as anions or
oxyanions in aqueous solution; they
generally form ionic or covalent
compounds when combined with metals
(unlike metals, which mostly form alloys
with other metals); and have acidic oxides
whereas the common oxides of nearly all
metals are basic.

Complicating the chemistry of the


nonmetals is the first row anomaly seen
particularly in hydrogen, (boron), carbon,
nitrogen, oxygen and fluorine; and the
alternation effect seen in (arsenic),
selenium and bromine.[12] The first row
anomaly largely arises from the electron
configurations of the elements concerned.

Hydrogen is noted for the different ways it


forms bonds. It most commonly forms
covalent bonds.[13] It can lose its single
valence electron in aqueous solution,
leaving behind a bare proton with
tremendous polarising power. This
subsequently attaches itself to the lone
electron pair of an oxygen atom in a water
molecule, thereby forming the basis of
acid-base chemistry.[14] Under certain
conditions a hydrogen atom in a molecule
can form a second, weaker, bond with an
atom or group of atoms in another
molecule. 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."[15]

From (boron) to neon, since the 2p


subshell has no inner analogue and
experiences no electron repulsion effects
it consequently has a relatively small
radius, unlike the 3p, 4p and 5p subshells
of heavier elements[16] (a similar effect is
seen in the 1s elements, hydrogen and
helium). Ionisation energies and
electronegativities among these elements
are consequently higher than would
otherwise be expected, having regard to
periodic trends. The small atomic radii of
carbon, nitrogen, and oxygen facilitates
the formation of triple or double bonds.[17]
The larger atomic radii, which enable
higher coordination numbers, and lower
electronegativities, which better tolerate
higher positive charges, of the heavier
group 15–18 nonmetals means they are
able to exhibit valences other than the
lowest for their group (that is, 3, 2, 1, or 0)
for example in PCl5, SF6, IF7, and XeF2.[18]
Period four elements immediately after the
first row of the transition metals, such as
selenium and bromine, have unusually
small atomic radii because the 3d
electrons are not effective at shielding the
increased nuclear charge, and smaller
atomic size correlates with higher
electronegativity.[19]

Categories
Immediately to the left of most nonmetals
on the periodic table are metalloids such
as boron, silicon, and germanium, which
generally behave chemically like
nonmetals,[20] and are included here for
comparative purposes. In this sense they
can be regarded as the most metallic of
nonmetallic elements.
Based on shared attributes, the nonmetals
can be divided into the two categories of
reactive nonmetal, and noble gas. The
metalloids and the two nonmetal
categories then span a progression in
chemical nature from weakly nonmetallic,
to moderately nonmetallic, to strongly
nonmetallic (oxygen and the four
nonmetallic halogens), to almost inert.
Analogous categories occur among the
metals in the form of the weakly metallic
(the post-transition metals), the
moderately metallic (most of the transition
metals), the strongly metallic (the alkali
metal and alkaline earth metals, and the
lanthanides and actinides), and the
relatively inert (the noble transition
metals).

As with categorisation schemes generally,


there is some variation and overlapping of
properties within and across each
category. One or more of the metalloids
are sometimes classified as nonmetals.[5]
Among the reactive nonmetals, carbon,
phosphorus, selenium, and iodine—which
border the metalloids—show some
metallic character, as does hydrogen.
Among the noble gases, radon is the most
metallic and begins to show some cationic
behaviour, which is unusual for a
nonmetal.[21]

Metalloid …

The seven Metalloids in the periodic table


metalloids
are boron
(B), silicon (Si), germanium (Ge), arsenic
(As), antimony (Sb), tellurium (Te), and
astatine (At). On a standard periodic table,
they occupy a diagonal area in the p-block
extending from boron at the upper left to
astatine at lower right, along the dividing
line between metals and nonmetals shown
on some periodic tables. They are called
metalloids mainly in light of their physical
resemblance to metals.

While they each have a metallic


appearance, they are brittle and only fair
conductors of electricity. Boron, silicon,
germanium, tellurium are semiconductors.
Arsenic and antimony have the electronic
band structures of semimetals although
both have less stable semiconducting
allotropes. Astatine has been predicted to
have a metallic crystalline structure.
Electronegativity values
of metalloids and nonmetals[n 2]
1 2 13 14 15 16 17 18
Noble
gases
H Reactive He
1   
2.2 nonmetals (5.5)
B C N O F Ne

2.04 2.55 3.04 3.44 3.98 (4.84)
Si P S Cl Ar

1.9 2.19 2.58 3.16 (3.2)
Ge As Se Br Kr

2.01 2.18 2.55 2.96 (2.94)
5  Sb Te I Xe
2.05 2.1 2.66 (2.4)
At Rn
6  Metalloids
2.2 (2.06)
Electronegativity (EN) gives some indication of
nonmetallic character. The metalloids have
uniformly moderate values (1.8–2.2). Among the
reactive nonmetals, hydrogen (2.2) and
phosphorus (2.19) have moderate values but they
each have higher ionisation energies than the
metalloids, and are very rarely classed as such.
Oxygen and the nonmetallic halogens have
uniformly high EN values; nitrogen has a high EN
but a marginally negative electron affinity that
makes it a reluctant anion former.[n 3] The noble
gases have some of the highest ENs but their
complete valence shells and sizeably negative
electron affinities render them chemically inert to a
large degree.

Chemically the metalloids generally


behave like (weak) nonmetals. They have
moderate ionisation energies, low to high
electron affinities, moderate
electronegativity values, are poor to
moderately strong oxidising agents, and
demonstrate a tendency to form alloys
with metals.

Reactive nonmetal …
T Reactive nonmetals in the periodic table
he
re
active nonmetals have a diverse range of
individual physical and chemical
properties. In periodic table terms they
largely occupy a position between the
weakly nonmetallic metalloids to the left
and the noble gases to the right.

Physically, five are solids, one is a liquid


(bromine), and five are gases. Of the
solids, graphite carbon, selenium, and
iodine are metallic-looking, whereas S8
sulfur has a pale-yellow appearance.
Ordinary white phosphorus has a
yellowish-white appearance but the black
allotrope, which is the most stable form of
phosphorus, has a metallic-looking
appearance. Bromine is a reddish-brown
liquid. Of the gases, fluorine and chlorine
are coloured pale yellow, and yellowish
green. Electrically, most are insulators
whereas graphite is a semimetal and black
phosphorus, selenium, and iodine are
semiconductors.
Chemically, they tend to have moderate to
high ionisation energies, electron affinities,
and electronegativity values, and be
relatively strong oxidising agents.
Collectively, the highest values of these
properties are found among oxygen and
the nonmetallic halogens. Manifestations
of this status include oxygen's major
association with the ubiquitous processes
of corrosion and combustion, and the
intrinsically corrosive nature of the
nonmetallic halogens. All five of these
nonmetals exhibit a tendency to form
predominately ionic compounds with
metals whereas the remaining nonmetals
tend to form predominately covalent
compounds with metals.

Noble gas …

Six Noble gases in the periodic table


nonmeta
ls are
categorised as noble gases: helium (He),
neon (Ne), argon (Ar), krypton (Kr), xenon
(Xe), and the radioactive radon (Rn). In
periodic table terms they occupy the
outermost right column. They are called
noble gases in light of their
characteristically very low chemical
reactivity.

They have very similar properties, all being


colorless, odorless, and nonflammable.
With their closed valence shells the noble
gases have feeble interatomic forces of
attraction resulting in very low melting and
boiling points.[22] That is why they are all
gases under standard conditions, even
those with atomic masses larger than
many normally solid elements.[23]

Chemically, the noble gases have relatively


high ionization energies, negative electron
affinities, and relatively high
electronegativities. Compounds of the
noble gases number less than half a
thousand, with most of these occurring via
oxygen or fluorine combining with either
krypton, xenon or radon.

The status of the period 7 congener of the


noble gases, oganesson (Og), is not
known—it may or may not be a noble gas.
It was originally predicted to be a noble
gas[24] but may instead be a fairly reactive
solid with an anomalously low first
ionisation potential, and a positive electron
affinity, due to relativistic effects.[25] On the
other hand, if relativistic effects peak in
period 7 at element 112, copernicium,
oganesson may turn out to be a noble gas
after all,[26] albeit more reactive than either
xenon or radon. While oganesson could be
expected to be the most metallic of the
group 18 elements, credible predictions on
its status as either a metal or a nonmetal
(or a metalloid) appear to be absent.

Alternative categories …
Nonmetal categorisation and alternat

Reactive nonmetal

H, C, N, P, O, S, Se, F, Cl, Br, I

Other nonmetal
(1)
H, C, N, P, O, S, (Se)

Solid Liquid
(2)
C, P, S, Se, I, At Br

(3) Electronegative Very electrone


nonmetal nonmetal
H, C, P, S, Se, I N, O, F, Cl, Br

Diatomic eleme
Polyatomic
element
(4)

C, P, S, Se H, N, O, F, Cl, Br

Hydrogen Nonmetal
(5)
H C, N, P, O, S, Se

(6) Metalloid Intermediate


nonmetal

B, Si, Ge, As, Sb, H, C, N, P, S, Se


Te

Hydrogen Metalloid Quintessential


nonmetal
(7)
B, Si, Ge, As, Sb,
H C, N, P, O, S, Se
Te, Po

The nonmetals are sometimes instead


divided according to either (1) the relative
homogeneity of the halogens; (2) physical
form; (3) electronegativity; (4) molecular
structure; (5) the peculiar nature of
hydrogen, and the relative homogeneity of
the halogens; (6) their analogous
categories among the metals; or (7) the
uniqueness of hydrogen and the treatment
of the metalloids as nonmetallic
analogues of the post-transition metals.

In scheme (1), the halogens are in a


category of their own; astatine is classed
as a nonmetal, rather than a metalloid; and
the remaining nonmetals are referred to as
other nonmetals.[27] If selenium is counted
as a metalloid rather than another
nonmetal, the resulting set of less active
nonmetals (H, C, N, P, O, S) are sometimes
instead referred to or categorised as
organogens,[28] CHONPS elements[29] or
biogens.[30] Collectively these six
nonmetals comprise the bulk of life on
Earth;[31] a rough estimate of the
composition of the biosphere is
C1450H3000O1450N15P1S1.[32]

In scheme (2), the nonmetals can simply


be divided based on their physical forms
at room temperature and pressure. The
fluid nonmetals (bromine and the gaseous
nonmetals) have the highest ionisation
energy and electronegativity values among
the elements, with the exception of
hydrogen which tends to be anomalous in
whichever category it is placed in. The
solid nonmetals are collectively the most
metallic of the nonmetallic elements, apart
from the metalloids.

In scheme (3), the nonmetals are divided


based on a loose correlation between
electronegativity and oxidizing power.[33]
Very electronegative nonmetals have
electronegativity values over 2.8;
electronegative nonmetals have values of
1.9 to 2.8.

In scheme (4), the nonmetals are


distinguished based on the molecular
structures of their most
thermodynamically stable forms in
ambient conditions.[34] Polyatomic
nonmetals form structures or molecules in
which each atom has two or three nearest
neighbours (Cx, P4, S8, Sex); diatomic
nonmetals form molecules in which each
atom has one nearest neighbour (H2, N2,
O2, F2, Cl2, Br2, I2); and the monatomic
noble gases exist as isolated atoms (He,
Ne, Ar, Kr, Xe, Rn) with no fixed nearest
neighbour. This gradual reduction in the
number of nearest neighbours
corresponds (approximately) to a
reduction in metallic character. A similar
progression is seem among the metals.
Metallic bonding tends to involve close-
packed centrosymmetric structures with a
high number of nearest neighbours. Post-
transition metals and metalloids,
sandwiched between the true metals and
the nonmetals, tend to have more complex
structures with an intermediate number of
nearest neighbours.

In scheme (5), hydrogen is placed by itself


on account of it being "so different from all
other elements".[35] The remaining
nonmetals are divided into nonmetals,
halogens, and noble gases, with the
unnamed category being distinguished by
including nonmetals with relatively strong
interatomic bonding, and the metalloids
being effectively treated as a third super-
category alongside metals and nonmetals.
Scheme 6: Metals and nonmetals
Noble gases
       
He, Ne, Ar, Kr, Xe, Rn
Active metals Corrosive nonmetals
Groups 1–3, Ln, An, (Al) O, F, Cl, Br, I
Intermediate
Transition metals
nonmetals
Most of them
H, C, N, P, S, Se
Frontier metals Metalloids
(Al) Ag, Sn, Bi etc B, Si, Ge, As, Sb, Te
Noble metals
Ru, Rh, Pd, Os, Ir, Pt, Au

In scheme (6) the nonmetals are divided


into four classes that complement a four-
fold division of the metals, with the noble
metals treated as a subset of the
transition metals. The metalloids are
treated as chemically weak nonmetals, in
a manner analogous to their chemically
weak frontier metal counterparts.[36]

In scheme (7), hydrogen is again placed by


itself on account of its uniqueness. The
remaining nonmetals are divided into
metalloids, quintessential nonmetals,
halogens, and noble gases. Since the
metalloids abut the post-transition or
"poor" metals, they might be renamed the
"poor non-metals".[37]

Comparison of properties
Characteristic and other properties of
metalloids, reactive nonmetals, and noble
gases are summarized in the following
table. Metalloids have been included in
light of their generally nonmetallic
chemistry. Physical properties are listed in
loose order of ease of determination;
chemical properties run from general to
specific, and then to descriptive.
Some properties of metalloids, reactive
nonmetals, and noble gases
Physical Reactive
property Metalloid nonmeta

Form solid solid: C, P, S


Se, I
liquid: Br
gaseous: H
O, F, Cl

Appearance metallic metallic,


coloured, o
translucent

Elasticity brittle brittle if sol


Atomic close-packed* polyatomic
structure or polyatomic P, S, Se
diatomic: H
O, F, Cl, Br, I

Bulk 12*, 6, 4, 3, or 3, 2, or 1
coordination 2
number

Allotropes most form known for C


O, S, Se

Electrical moderate poor to


conductivity moderate

Volatility low: B, Si, Ge, low: C


Sb, Te moderate: P
moderate: As, Se, Br, I
At? high: H, N, O
Cl

Electronic metallic* to semimetall


structure semiconductor semicondu
or insulator

Outer s and p 3–7 1, 4–7


electrons

Crystal structure rhombohedral:  


B, As, Sb cubic: P, O,
cubic: Si, Ge, hexagonal:
At? C, N, Se
hexagonal: Te orthorhomb
S, Cl, Br, I

Chemical Metalloid Reactive


property nonmeta

General nonmetallic to incipient


chemical metallic
behaviour

Ionization low moderate t


energy high

Electron affinity low to high moderate t


high
(exception:
is negative)

Electronegativity moderate: moderate t


Si < Ge ≈ B ≈  high:
Sb < Te < As ≈  P < Se ≈ C
At < I < Br < N <
< O < F

Non-zero • negative oxidation states


oxidation states known for all, but for H this
an unstable state
• positive oxidation states
known for all but F, and only
exceptionally for O
• from −5 for B to +7 for Cl,
and At

Oxidising power low low to high


(exception: At
is moderate)

Catenation marked marked


tendency tendency: C
S, Se
less tenden
H, N, O, F, C
I

Compounds tend to form mainly


with metals alloys or inter- covalent: H
metallic N, P, S, Se
compounds mainly ionic
F, Cl, Br, I

Oxides • polymeric in • C, P, S, Se,


structure[39] and I are
• B, Si, Ge, As, known in at
Sb, Te[40] are least one
glass formers polymeric f
• tend to be • P, S, Se are
amphoteric or glass forme
weakly CO2 forms
acidic[41][42] glass at
40 GPa
• acidic, or
neutral (H2O
CO, NO, N2O

Sulfates most form some form

*Bulk astatine has been predicted to have a metallic

face-centred cubic structure


† Hydrogen can also form alloy-like hydrides

Properties of nonmetals (and


metalloids) by Group
Abbreviations used in this section are: AR
Allred-Rochow; CN coordination number;
and MH Moh's hardness

Group 1 …

Hydrogen in an electrical discharge tube


Hydrogen is a colourless, odourless, and
comparatively unreactive diatomic gas
with a density of 8.988 × 10−5 g/cm3 and
is about 14 times lighter than air. It
condenses to a colourless liquid
−252.879 °C and freezes into an ice- or
snow-like solid at −259.16 °C. The solid
form has a hexagonal crystalline structure
and is soft and easily crushed. Hydrogen
is an insulator in all of its forms. It has a
high ionisation energy (1312.0 kJ/mol),
moderate electron affinity (73 kJ/mol), and
moderate electronegativity (2.2). Hydrogen
is a poor oxidising agent (H2 + 2e− → 2H–
= –2.25 V at pH 0). Its chemistry, most of
which is based around its tendency to
acquire the electron configuration of the
noble gas helium, is largely covalent in
nature, noting it can form ionic hydrides
with highly electropositive metals, and
alloy-like hydrides with some transition
metals. The common oxide of hydrogen
(H2O) is a neutral oxide.[n 4]

Group 13 …
Boron

Boron is a lustrous, barely reactive solid


with a density 2.34 g/cm3 (cf. aluminium
2.70), and is hard (MH 9.3) and brittle. It
melts at 2076 °C (cf. steel ~1370 °C) and
boils at 3927 °C. Boron has a complex
rhombohedral crystalline structure (CN
5+). It is a semiconductor with a band gap
of about 1.56 eV. Boron has a moderate
ionisation energy (800.6 kJ/mol), low
electron affinity (27 kJ/mol), and moderate
electronegativity (2.04). Being a metalloid,
most of its chemistry is nonmetallic in
nature. Boron is a poor oxidizing agent
(B12 + 3e → BH3 = –0.15 V at pH 0). While
it bonds covalently in nearly all of its
compounds, it can form intermetallic
compounds and alloys with transition
metals of the composition MnB, if n > 2.
The common oxide of boron (B2O3) is
weakly acidic.
Group 14 …

Carbon, as graphite

Carbon (as graphite, its most


thermodynamically stable form) is a
lustrous and comparatively unreactive
solid with a density of 2.267 g/cm3, and is
soft (MH 0.5) and brittle. It sublimes to
vapour at 3642 C°. Carbon has a
hexagonal crystalline structure (CN 3). It is
a semimetal in the direction of its planes,
with an electrical conductivity exceeding
that of some metals, and behaves as a
semiconductor in the direction
perpendicular to its planes. It has a high
ionisation energy (1086.5 kJ/mol),
moderate electron affinity (122 kJ/mol),
and high electronegativity (2.55). Carbon
is a poor oxidising agent (C + 4e− → CH4 =
0.13 V at pH 0). Its chemistry is largely
covalent in nature, noting it can form salt-
like carbides with highly electropositive
metals. The common oxide of carbon
(CO2) is a medium-strength acidic oxide.

Silicon has a blue-grey metallic lustre.

Silicon is a metallic-looking relatively


unreactive solid with a density of
2.3290 g/cm3, and is hard (MH 6.5) and
brittle. It melts at 1414 °C (cf. steel
~1370 °C) and boils at 3265 °C. Silicon
has a diamond cubic structure (CN 4). It is
a semiconductor with a band gap of about
1.11 eV. Silicon has a moderate ionisation
energy (786.5 kJ/mol), moderate electron
affinity (134 kJ/mol), and moderate
electronegativity (1.9). It is a poor
oxidising agent (Si + 4e → Si4 = –0.147 at
pH 0). As a metalloid the chemistry of
silicon is largely covalent in nature, noting
it can form alloys with metals such as iron
and copper. The common oxide of silicon
(SiO2) is weakly acidic.
Germanium

Germanium is a shiny, mostly unreactive


grey-white solid with a density of
5.323 g/cm3 (about two-thirds that of
iron), and is hard (MH 6.0) and brittle. It
melts at 938.25 °C (cf. silver 961.78 °C)
and boils at 2833 °C. Germanium has a
diamond cubic structure (CN 4). It is a
semiconductor with a band gap of about
0.67 eV. Germanium has a moderate
ionisation energy (762 kJ/mol), moderate
electron affinity (119 kJ/mol), and
moderate electronegativity (2.01). It is a
poor oxidising agent (Ge + 4e → GeH4 = –
0.294 at pH 0). As a metalloid the
chemistry of germanium is largely
covalent in nature, noting it can form
alloys with metals such as aluminium and
gold. Most alloys of germanium with
metals lack metallic or semimetallic
conductivity. The common oxide of
germanium (GeO2) is amphoteric.

Group 15 …

Liquid nitrogen

Nitrogen is a colourless, odourless, and


relatively inert diatomic gas with a density
of 1.251 × 10−3 g/cm3 (marginally heavier
than air). It condenses to a colourless
liquid at −195.795 °C and freezes into an
ice- or snow-like solid at −210.00 °C. The
solid form (density 0.85 g/cm3; cf. lithium
0.534) has a hexagonal crystalline
structure and is soft and easily crushed.
Nitrogen is an insulator in all of its forms.
It has a high ionisation energy
(1402.3 kJ/mol), low electron affinity (–
6.75 kJ/mol), and high electronegativity
(3.04). The latter property manifests in the
capacity of nitrogen to form usually strong
hydrogen bonds, and its preference for
forming complexes with metals having low
electronegativities, small cationic radii,
and often high charges (+3 or more).
Nitrogen is a poor oxidising agent (N2 +
6e− → 2NH3 = −0.057 V at pH 0). Only
when it is in a positive oxidation state, that
is, in combination with oxygen or fluorine,
are its compounds good oxidising agents,
for example, 2NO3− → N2 = 1.25 V. Its
chemistry is largely covalent in nature;
anion formation is energetically
unfavourable owing to strong inter
electron repulsions associated with having
three unpaired electrons in its outer
valence shell, hence its negative electron
affinity. The common oxide of nitrogen
(NO) is weakly acidic. Many compounds of
nitrogen are less stable than diatomic
nitrogen, so nitrogen atoms in compounds
seek to recombine if possible and release
energy and nitrogen gas in the process,
which can be leveraged for explosive
purposes.
Phosphorus, as black phosphorus

Phosphorus in its most


thermodynamically stable black form, is a
lustrous and comparatively unreactive
solid with a density of 2.69 g/cm3, and is
soft (MH 2.0) and has a flaky
comportment. It sublimes at 620 °C. Black
phosphorus has an orthorhombic
crystalline structure (CN 3). It is a
semiconductor with a band gap of 0.3 eV.
It has a high ionisation energy
(1086.5 kJ/mol), moderate electron affinity
(72 kJ/mol), and moderate
electronegativity (2.19). In comparison to
nitrogen, phosphorus usually forms weak
hydrogen bonds, and prefers to form
complexes with metals having high
electronegativities, large cationic radii, and
often low charges (usually +1 or +2.
Phosphorus is a poor oxidising agent (P4 +
3e− → PH3– = −0.046 V at pH 0 for the
white form, −0.088 V for the red). Its
chemistry is largely covalent in nature,
noting it can form salt-like phosphides
with highly electropositive metals.
Compared to nitrogen, electrons have
more space on phosphorus, which lowers
their mutual repulsion and results in anion
formation requiring less energy. The
common oxide of phosphorus (P2O5) is a
medium-strength acidic oxide.

White phosphorus, stored under water to prevent its


oxidation[44]
When assessing periodicity in the
properties of the elements it needs to be
borne in mind that the quoted properties
of phosphorus tend to be those of its least
stable white form rather than, as is the
case with all other elements, the most
stable form. White phosphorus is the most
common, industrially important, and easily
reproducible allotrope. For those reasons
it is the standard state of the element.
Paradoxically, it is also thermodynamically
the least stable, as well as the most
volatile and reactive form. It gradually
changes to red phosphorus. This
transformation is accelerated by light and
heat, and samples of white phosphorus
almost always contain some red
phosphorus and, accordingly, appear
yellow. For this reason, white phosphorus
that is aged or otherwise impure is also
called yellow phosphorus. When exposed
to oxygen, white phosphorus glows in the
dark with a very faint tinge of green and
blue. It is highly flammable and pyrophoric
(self-igniting) upon contact with air. White
phosphorus has a density of 1.823 g/cm3,
is soft (MH 0.5) as wax, pliable and can be
cut with a knife. It melts at 44.15 °C and, if
heated rapidly, boils at 280.5 °C; it
otherwise remains solid and transforms to
violet phosphorus at 550 °C. It has a body-
centred cubic structure, analogous to that
of manganese, with unit cell comprising
58 P4 molecules. It is an insulator with a
band gap of about 3.7 eV.

Arsenic, sealed in a container to prevent tarnishing


Arsenic is a grey, metallic looking solid
which is stable in dry air but develops a
golden bronze patina in moist air, which
blackens on further exposure. It has a
density of 5.727 g/cm3, and is brittle and
moderately hard (MH 3.5; more than
aluminium; less than iron). Arsenic
sublimes at 615 °C. It has a rhombohedral
polyatomic crystalline structure (CN 3).
Arsenic is a semimetal, with an electrical
conductivity of around 3.9 × 104 S•cm−1
and a band overlap of 0.5 eV. It has a
moderate ionisation energy (947 kJ/mol),
moderate electron affinity (79 kJ/mol), and
moderate electronegativity (2.18). Arsenic
is a poor oxidising agent (As + 3e → AsH3
= –0.22 at pH 0). As a metalloid, its
chemistry is largely covalent in nature,
noting it can form brittle alloys with
metals, and has an extensive
organometallic chemistry. Most alloys of
arsenic with metals lack metallic or
semimetallic conductivity. The common
oxide of arsenic (As2O3) is acidic but
weakly amphoteric.
Antimony, showing its brilliant lustre

Antimony is a silver-white solid with a blue


tint and a brilliant lustre. It is stable in air
and moisture at room temperature.
Antimony has a density of 6.697 g/cm3,
and is moderately hard (MH 3.0; about the
same as copper). It has a rhombohedral
crystalline structure (CN 3). Antimony
melts at 630.63 °C and boils at 1635 °C. It
is a semimetal, with an electrical
conductivity of around 3.1 × 104 S•cm−1
and a band overlap of 0.16 eV. Antimony
has a moderate ionisation energy
(834 kJ/mol), moderate electron affinity
(101 kJ/mol), and moderate
electronegativity (2.05). It is a poor
oxidising agent (Sb + 3e → SbH3 = –0.51
at pH 0). As a metalloid, its chemistry is
largely covalent in nature, noting it can
form alloys with one or more metals such
as aluminium, iron, nickel, copper, zinc, tin,
lead and bismuth, and has an extensive
organometallic chemistry. Most alloys of
antimony with metals have metallic or
semimetallic conductivity. The common
oxide of antimony (Sb2O3) is amphoteric.

Group 16 …

Liquid oxygen (boiling)


“ In the United States alone, more than $10
billion is lost each year to corrosion…Much of
this corrosion is the rusting of iron and steel…
The oxidizing agent causing all of this
corrosion is usually oxygen. ”

MD Joesten, L Hogg, and ME Castellion


In The world of chemistry (2007, p. 217)

Oxygen is a colourless, odourless, and


unpredictably reactive diatomic gas with a
gaseous density of 1.429 × 10−3 g/cm3
(marginally heavier than air). It is generally
unreactive at room temperature. Thus,
sodium metal will "retain its metallic lustre
for days in the presence of absolutely dry
air and can even be melted (m.p. 97.82 °C)
in the presence of dry oxygen without
igniting".[45] On the other hand, oxygen can
react with many inorganic and organic
compounds either spontaneously or under
the right conditions,[46] (such as a flame or
a spark) [or ultra-violet light?]. It
condenses to pale blue liquid −182.962 °C
and freezes into a light blue solid at
−218.79 °C. The solid form (density
0.0763 g/cm3) has a cubic crystalline
structure and is soft and easily crushed.
Oxygen is an insulator in all of its forms. It
has a high ionisation energy
(1313.9 kJ/mol), high electron affinity
(141 kJ/mol), and high electronegativity
(3.44). Oxygen is a strong oxidising agent
(O2 + 4e → 2H2O = 1.23 V at pH 0). Metal
oxides are largely ionic in nature.[47]
Sulfur

Sulfur is a bright-yellow moderately


reactive[48] solid. It has a density of
2.07 g/cm3 and is soft (MH 2.0) and
brittle. It melts to a light yellow liquid
95.3 °C and boils at 444.6 °C. Sulfur has
an abundance on earth one-tenth that of
oxygen. It has an orthorhombic polyatomic
(CN 2) crystalline structure, and is brittle.
Sulfur is an insulator with a band gap of
2.6 eV, and a photoconductor meaning its
electrical conductivity increases a million-
fold when illuminated. Sulfur has a
moderate ionisation energy
(999.6 kJ/mol), moderate electron affinity
(200 kJ/mol), and high electronegativity
(2.58). It is a poor oxidising agent (S8 + 2e−
→ H2S = 0.14 V at pH 0). The chemistry of
sulfur is largely covalent in nature, noting it
can form ionic sulfides with highly
electropositive metals. The common oxide
of sulfur (SO3) is strongly acidic.
Selenium

Selenium is a metallic-looking, moderately


reactive[48] solid with a density of
4.81 g/cm3 and is soft (MH 2.0) and
brittle. It melts at 221 °C to a black liquid
and boils at 685 °C to a dark yellow
vapour. Selenium has a hexagonal
polyatomic (CN 2) crystalline structure. It
is a semiconductor with a band gap of
1.7 eV, and a photoconductor meaning its
electrical conductivity increases a million-
fold when illuminated. Selenium has a
moderate ionisation energy
(941.0 kJ/mol), high electron affinity
(195 kJ/mol), and high electronegativity
(2.55). It is a poor oxidising agent (Se +
2e− → H2Se = −0.082 V at pH 0). The
chemistry of selenium is largely covalent
in nature, noting it can form ionic
selenides with highly electropositive
metals. The common oxide of selenium
(SeO3) is strongly acidic.

Tellurium

Tellurium is a silvery-white, moderately


reactive,[48] shiny solid, that has a density
of 6.24 g/cm3 and is soft (MH 2.25) and
brittle. It is the softest of the commonly
recognised metalloids. Tellurium reacts
with boiling water, or when freshly
precipitated even at 50 °C, to give the
dioxide and hydrogen: Te + 2 H2O → TeO2
+ 2 H2. It has a melting point of 450 °C and
a boiling point of 988 °C. Tellurium has a
polyatomic (CN 2) hexagonal crystalline
structure. It is a semiconductor with a
band gap of 0.32 to 0.38 eV. Tellurium has
a moderate ionisation energy
(869.3 kJ/mol), high electron affinity
(190 kJ/mol), and moderate
electronegativity (2.1). It is a poor
oxidising agent (Te + 2e− → H2Te =
−0.45 V at pH 0). The chemistry of
tellurium is largely covalent in nature,
noting it has an extensive organometallic
chemistry and that many tellurides can be
regarded as metallic alloys. The common
oxide of tellurium (TeO2) is amphoteric.

Group 17 …
Liquid fluorine, in a cryogenic bath

Fluorine is an extremely toxic and reactive


pale yellow diatomic gas that, with a
gaseous density of 1.696 × 10−3 g/cm3, is
about 40% heavier than air. Its extreme
reactivity is such that it was not isolated
(via electrolysis) until 1886 and was not
isolated chemically until 1986. Its
occurrence in an uncombined state in
nature was first reported in 2012, but is
contentious. Fluorine condenses to a pale
yellow liquid at −188.11 °C and freezes
into a colourless solid[45] at −219.67 °C.
The solid form (density 1.7 g/cm−3) has a
cubic crystalline structure and is soft and
easily crushed. Fluorine is an insulator in
all of its forms. It has a high ionisation
energy (1681 kJ/mol), high electron
affinity (328 kJ/mol), and high
electronegativity (3.98). Fluorine is a
powerful oxidising agent (F2 + 2e → 2HF =
2.87 V at pH 0); "even water, in the form of
steam, will catch fire in an atmosphere of
fluorine".[49] Metal fluorides are generally
ionic in nature.
Chlorine gas

Chlorine is an irritating green-yellow


diatomic gas that is extremely reactive,
and has a gaseous density of 3.2 ×
10−3 g/cm3 (about 2.5 times heavier than
air). It condenses at −34.04 °C to an
amber-coloured liquid and freezes at
−101.5 °C into a yellow crystalline solid.
The solid form (density 1.9 g/cm−3) has an
orthorhombic crystalline structure and is
soft and easily crushed. Chlorine is an
insulator in all of its forms. It has a high
ionisation energy (1251.2 kJ/mol), high
electron affinity (349 kJ/mol; higher than
fluorine), and high electronegativity (3.16).
Chlorine is a strong oxidising agent (Cl2 +
2e → 2HCl = 1.36 V at pH 0). Metal
chlorides are largely ionic in nature. The
common oxide of chlorine (Cl2O7) is
strongly acidic.
Liquid bromine

Bromine is a deep brown diatomic liquid


that is quite reactive, and has a liquid
density of 3.1028 g/cm3. It boils at 58.8 °C
and solidifies at −7.3 °C to an orange
crystalline solid (density 4.05 g/cm−3). It is
the only element, apart from mercury,
known to be a liquid at room temperature.
The solid form, like chlorine, has an
orthorhombic crystalline structure and is
soft and easily crushed. Bromine is an
insulator in all of its forms. It has a high
ionisation energy (1139.9 kJ/mol), high
electron affinity (324 kJ/mol), and high
electronegativity (2.96). Bromine is a
strong oxidising agent (Br2 + 2e → 2HBr =
1.07 V at pH 0). Metal bromides are largely
ionic in nature. The unstable common
oxide of bromine (Br2O5) is strongly acidic.
Iodine crystals

Iodine, the rarest of the nonmetallic


halogens, is a metallic looking solid that is
moderately reactive, and has a density of
4.933 g/cm3. It melts at 113.7 °C to a
brown liquid and boils at 184.3 °C to a
violet-coloured vapour. It has an
orthorhombic crystalline structure with a
flaky habit. Iodine is semiconductor in the
direction of its planes, with a band gap of
about 1.3 eV and a conductivity of 1.7 ×
10−8 S•cm−1 at room temperature. This is
higher than selenium but lower than boron,
the least electrically conducting of the
recognised metalloids. Iodine is an
insulator in the direction perpendicular to
its planes. It has a high ionisation energy
(1008.4 kJ/mol), high electron affinity
(295 kJ/mol), and high electronegativity
(2.66). Iodine is a moderately strong
oxidising agent (I2 + 2e → 2I− = 0.53 V at
pH 0). Metal iodides are predominantly
ionic in nature. The only stable oxide of
iodine (I2O5) is strongly acidic.

Astatine is expected to have properties


intermediate between iodine, a nonmetal
with incident metallic properties, and
tennessine, which is predicted to be a
metal. Astatine has not so far been
synthesised in sufficient quantities to
enable a determination of its bulk
properties. A macro-sized sample of
astatine would vaporise itself due to
radioactive heating; it is not known if such
a phenomenon could be prevented with
sufficient cooling. Many of the properties
of astatine have nevertheless been
predicted. It is expected to have a metallic
appearance, a density of 6.35±0.15 g/cm3,
a melting point of 302 °C, a boiling point of
337 °C(?), and a face-centred cubic
crystalline structure. It has a moderate
ionisation energy (899.003 kJ/mol), and is
expected to have a high electron affinity
(222 kJ/mol), and moderate
electronegativity (2.2). Astatine is a weak
oxidizing agent (At + e → At− = 0.3 V at
pH 0).

Group 18 …

Liquified helium

Helium has a density of 1.785 ×


10−4 g/cm3 (cf. air 1.225 × 10−3 g/cm3),
liquifies at −268.928 °C, and cannot be
solidified at normal pressure. It has the
lowest boiling point of all of the elements.
Liquid helium exhibits super-fluidity,
superconductivity, and near-zero viscosity;
its thermal conductivity is greater than
that of any other known substance (more
than 1,000 times that of copper). Helium
can only be solidified at −272.20 °C under
a pressure of 2.5 MPa. It has a very high
ionisation energy (2372.3 kJ/mol), low
electron affinity (estimated at −50 kJ/mol),
and very high electronegativity (5.5 AR).
No normal compounds of helium have so
far been synthesised.

Neon in an electrical discharge tube

Neon has a density of 9.002 × 10-4 g/cm3,


liquifies at −245.95 °C, and solidifies at
−248.45 °C. It has the narrowest liquid
range of any element and, in liquid form,
has over 40 times the refrigerating
capacity of liquid helium and three times
that of liquid hydrogen. Neon has a very
high ionisation energy (2080.7 kJ/mol),
low electron affinity (estimated at
−120 kJ/mol), and very high
electronegativity (4.84 AR). It is the least
reactive of the noble gases; no normal
compounds of neon have so far been
synthesised.
A small piece of rapidly melting solid argon

Argon has a density of 1.784 ×


10−3 g/cm3, liquifies at −185.848 °C, and
solidifies at −189.34 °C. Although non-
toxic, it is 38% denser than air and
therefore considered a dangerous
asphyxiant in closed areas. It is difficult to
detect because (like all the noble gases) it
is colourless, odourless, and tasteless.
Argon has a high ionisation energy
(1520.6 kJ/mol), low electron affinity
(estimated at −96 kJ/mol), and high
electronegativity (3.2 AR). One interstitial
compound of argon, Ar1C60 is a stable
solid at room temperature.

A Kr-shaped krypton discharge tube

Krypton has a density of 3.749 ×


10−3 g/cm3, liquifies at −153.415 °C, and
solidifies at −157.37 °C. It has a high
ionisation energy (1350.8 kJ/mol), low
electron affinity (estimated at −60 kJ/mol),
and high electronegativity (2.94 AR).
Krypton can be reacted with fluorine to
form the difluoride, KrF2. The reaction of
KrF2 with B(OTeF5)3 produces an unstable
compound, Kr(OTeF5)2, that contains a
krypton-oxygen bond.

Pressurized xenon gas encapsulated in an acrylic


cube
cube

Xenon has a density of 5.894 ×


10−3 g/cm3, liquifies at −161.4 °C, and
solidifies at −165.051 °C. It is non-toxic,
and belongs to a select group of
substances that penetrate the blood–brain
barrier, causing mild to full surgical
anesthesia when inhaled in high
concentrations with oxygen. Xenon has a
high ionisation energy (1170.4 kJ/mol),
low electron affinity (estimated at
−80 kJ/mol), and high electronegativity
(2.4 AR). It forms a relatively large number
of compounds, mostly containing fluorine
or oxygen. An unusual ion containing
xenon is the tetraxenonogold(II) cation,
AuXe2+
4 , which contains Xe–Au bonds.
This ion occurs in the compound
AuXe4(Sb2F11)2, and is remarkable in
having direct chemical bonds between two
notoriously unreactive atoms, xenon and
gold, with xenon acting as a transition
metal ligand. The compound Xe2Sb2F11
contains a Xe–Xe bond, the longest
element-element bond known (308.71 pm
= 3.0871 Å). The most common oxide of
xenon (XeO3) is strongly acidic.

Radon, which is radioactive, has a density


of 9.73 × 10−3 g/cm3, liquifies at −61.7 °C,
and solidifies at −71 °C. It has a high
ionisation energy (1037 kJ/mol), low
electron affinity (estimated at −70 kJ/mol),
and moderate electronegativity (2.06 AR).
The only confirmed compounds of radon,
which is the rarest of the naturally
occurring noble gases, are the difluoride
RnF2, and trioxide, RnO3. It has been
reported that radon is capable of forming
a simple Rn2+ cation in halogen fluoride
solution, which is highly unusual behaviour
for a nonmetal, and a noble gas at that.
Radon trioxide (RnO3) is expected to be
acidic.

Oganesson, the heaviest element on the


periodic table, has only recently been
synthesized. Owing to its short half-life, its
chemical properties have not yet been
investigated. Due to the significant
relativistic destabilisation of the 7p3/2
orbitals, it is expected to be significantly
reactive and behave more similarly to the
group 14 elements, as it effectively has
four valence electrons outside a pseudo-
noble gas core. Its boiling point is
expected to be about 80±30 °C, so that it is
probably neither noble nor a gas; as a
liquid it is expected to have a density of
about 5 g/cm3. It is expected to have a
barely positive electron affinity (estimated
as 5 kJ/mol) and a moderate ionisation
energy of about 860 kJ/mol, which is
rather low for a nonmetal and close to
those of the metalloids tellurium and
astatine. The oganesson fluorides OgF2
and OgF4 are expected to show significant
ionic character, suggesting that
oganesson may have at least incipient
metallic properties. The oxides of
oganesson, OgO and OgO2, are predicted
to be amphoteric.

Cross-cutting relationships …
Periodic table extract showing some relationships
among the nonmetals. The dashed line around H
denotes that H is normally positioned on the far left of
the periodic table, above Li in group 1. The red arrows
denote that, as with the metalloids, the most stable
forms of C, P, Se, and I each have a metallic
appearance. The white arrow denotes that N, S, and Br
are a gas, solid, and liquid, respectively. That leaves
the triangle of O, F, and Cl representing the most
corrosive nonmetals. Not shown here are At (a
metalloid, predicted to be a post-transition metal), Rn
(a noble gas, showing incipient metallic behavior), and
Og (possibly a metalloid).

Some pairs of nonmetals show additional


relationships, beyond those associated
with group membership.

H and C. Hydrogen in group 1, and carbon


in group 14, show some out-of-group
similarities.[50] These include proximity in
ionization energies, electron affinities and
electronegativity values; half-filled valence
shells; and correlations between the
chemistry of H–H and C–H bonds.
H and N. Both are relatively unreactive
colourless diatomic gases, with
comparably high ionization energies
(1312.0 and 1402.3 kJ/mol), each having
half-valence subshells, 1s and 2p
respectively. Like the reactive azide N3−
anion, inter-electron repulsions in the H−
hydride anion (with its single nuclear
charge) make ionic hydrides highly
reactive. Unusually for nonmetals, the two
elements are known in cationic forms. In
water the H+ "cation" exists as an H13O6+
ion, with a delocalised proton in a central
OHO group.[51] Nitrogen forms an N5+
pentazenium cation; bulk quantities of the
salt N5+SbF6− can be prepared.
Coincidentally, the NH4+ ammonium cation
behaves in many respects as an alkali
metal anion.[52]

C and N. With nitrogen, carbon forms an


extensive series of nitride compounds
including those with high N:C ratios, and
with structures that are simple (CN12);
chain-like (C6N2 for example); graphitic
(linked C6N7 units); fullerenic (C48N12) or
polymeric (C3N3 units). Most of the
compounds prepared to date also contain
quantities of hydrogen.[53]

C and P. Carbon and phosphorus represent


an example of a less-well known diagonal
relationship, especially in organic
chemistry. "Spectacular" evidence of this
relationship was provided in 1987 with the
synthesis of a ferrocene-like molecule in
which six of the carbon atoms were
replaced by phosphorus atoms.[54] Further
illustrating the theme is the "extraordinary"
similarity between low coordinate
phosphorus compounds and unsaturated
carbon compounds, and related research
into organophosphorus chemistry.[55] In
2020, the first compound containing three
carbon atoms and one phosphorus
arranged in a tetrahedron, tri-tert-butyl
phosphatetrahedrane, (PC3)(C4H9)3 was
synthesised. While plain all-carbon
tetrahedrane (CH)4 has never been
isolated, phosphorus was selected in light
of its capacity to form tetrahedral
molecules, and the similarity of some of
its properties to those of carbon.[56]

N and P. Like nitrogen, the chemistry of


phosphorus is that of the covalent bond;
the two nonmetals rarely form anions.
Despite them being in the same group, and
the composition of some of their
compounds resembling one another, the
individual chemistries of nitrogen and
phosphorus are very different.[57] That
said, the two elements form an extensive
series of phosphorus–nitrogen
compounds having chain, ring and cage
structures; the P–N repeat unit in these
structures bears a strong resemblance to
the S–N repeat unit found in the wide
range of sulfur–nitrogen compounds,
discussed next.[58]

N and S. Nitrogen and sulfur have a less-


well known diagonal relationship,
manifested in like charge densities and
electronegativities (the latter are identical
if only the p electrons are counted; see
Hinze and Jaffe 1962) especially when
sulfur is bonded to an electron-
withdrawing group. They are able to form
an extensive series of seemingly
interchangeable sulfur nitrides, the most
famous of which, polymeric sulfur nitride,
is metallic, and a superconductor below
0.26 K. The aromatic nature of the S3N22+
ion, in particular, serves as an "exemplar"
of the similarity of electronic energies
between the two nonmetals.[54]

P and S (Se). Phosphorus reacts with


sulfur and selenium (and oxygen) to form
a large number of compounds. These
compounds are characterized by structural
analogies derived from the white
phosphorus P4 tetrahedron.[59]

S and Se Commonalties between sulfur


and selenium are abundantly obvious. For
example, selenium is found in metal
sulfide ores, where it partially replaces
sulfur; both elements are photoconductors
—their electrical conductivities increase by
up to six orders of magnitude when
exposed to light.[60]
O and F. Fluorine and oxygen share the
ability to often bring out the highest
oxidation states among the elements.

O and Cl. "Chlorination reactions have


many similarities to oxidation reactions.
They tend not to be limited to
thermodynamic equilibrium and often go
to complete chlorination. The reactions
are often highly exothermic. Chlorine, like
oxygen, forms flammable mixtures with
organic compounds."[61]
I and Xe. The chemistry of iodine in its
oxidation states of +1, +3, +5, and +7 is
analogous to that of xenon in an
immediately higher oxidation state.

Allotropes

Some allotropes of carbon


Many nonmetals have less stable
allotropes, with either nonmetallic or
metallic properties. Graphite, the standard
state of carbon, has a lustrous appearance
and is a fairly good electrical conductor.
The diamond allotrope of carbon is clearly
nonmetallic, however, being translucent
and having a relatively poor electrical
conductivity. Carbon is also known in
several other allotropic forms, including
semiconducting buckminsterfullerene
(C60). Nitrogen can form gaseous
tetranitrogen (N4), an unstable polyatomic
molecule with a lifetime of about one
microsecond.[62] Oxygen is a diatomic
molecule in its standard state; it also
exists as ozone (O3), an unstable
nonmetallic allotrope with a half-life of
around half an hour.[63] Phosphorus,
uniquely, exists in several allotropic forms
that are more stable than that of its
standard state as white phosphorus (P4).
The red and black allotropes are probably
the best known; both are semiconductors.
Phosphorus is also known as
diphosphorus (P2), an unstable diatomic
allotrope.[64] Sulfur has more allotropes
than any other element;[65] all of these,
except plastic sulfur (a metastable ductile
mixture of allotropes)[66] have nonmetallic
properties. Selenium has several
nonmetallic allotropes, all of which are
much less electrically conducting than its
standard state of grey "metallic"
selenium.[67] Iodine is also known in a
semiconducting amorphous form.[68]
Under sufficiently high pressures, just over
half of the nonmetals, starting with
phosphorus at 1.7 GPa,[69] have been
observed to form metallic allotropes.

Most metalloids, like the less


electronegative nonmetals, form
allotropes. Boron is known in several
crystalline and amorphous forms. The
discovery of a quasispherical allotropic
molecule borospherene (B40) was
announced in July 2014. Silicon was most
recently known only in its crystalline and
amorphous forms. Silicene, a two-
dimensional allotrope of silicon, with a
hexagonal honeycomb structure similar to
that of graphene, was observed in 2010.
The synthesis of an orthorhombic
allotrope Si24, was subsequently reported
in 2014. At pressure of ~10–11 GPa,
germanium transforms to a metallic phase
with the same tetragonal structure as tin;
when decompressed—and depending on
the speed of pressure release—metallic
germanium forms a series of allotropes
that are metastable at ambient condition.
Germanium also forms a graphene
analogue, germanene. Arsenic and
antimony form several well known
allotropes (yellow, grey, and black).
Tellurium is known only in its crystalline
and amorphous forms; astatine is not
known to have any allotropes.

Abundance and extraction

Fluorite, a source of fluorine


Hydrogen and helium are estimated to
make up approximately 99 per cent of all
ordinary matter in the universe. Less than
five percent of the Universe is believed to
be made of ordinary matter, represented
by stars, planets and living beings. The
balance is made of dark energy and dark
matter, both of which are poorly
understood at present.[70]

Hydrogen, carbon, nitrogen, and oxygen


constitute the great bulk of the Earth's
atmosphere, oceans, crust, and biosphere;
the remaining nonmetals have
abundances of 0.5 per cent or less. In
comparison, 35 per cent of the crust is
made up of the metals sodium,
magnesium, aluminium, potassium and
iron; together with a metalloid, silicon. All
other metals and metalloids have
abundances within the crust, oceans or
biosphere of 0.2 per cent or less.[71]

Nonmetals, and metalloids, in their


elemental forms are extracted from:[72]
brine: Cl, Br, I; liquid air: N, O, Ne, Ar, Kr, Xe;
minerals: B (borate minerals); C (coal;
diamond; graphite); F (fluorite); Si (silica) P
(phosphates); Sb (stibnite, tetrahedrite); I
(in sodium iodate NaIO3 and sodium
iodide NaI); natural gas: H, He, S; and from
ores, as processing byproducts: Ge (zinc
ores); As (copper and lead ores); Se, Te
(copper ores); and Rn (uranium bearing
ores). Astatine is produced in minute
quantities by irradiating bismuth.

Applications in common
For prevalent and speciality applications
of individual nonmetals see the main
article for each element.

A high-voltage circuit-breaker employing sulfur


hexafluoride SF6 as its inert (air replacement)
interrupting medium[73]

Nonmetals do not have any universal or


near-universal applications. This is not the
case with metals, most of which have
structural uses; nor the metalloids, the
typical uses of which extend to (for
example) oxide glasses, alloying
components, and semiconductors.

Shared applications of different subsets of


the nonmetals instead encompass their
presence in, or specific uses in the fields
of cryogenics and refrigerants: H, He, N, O,
F and Ne; fertilisers: H, N, P, S, Cl (as a
micronutrient) and Se; household
accoutrements: H (primary constituent of
water), He (party balloons), C (in pencils,
as graphite), N (beer widgets), O (as
peroxide, in detergents), F (as fluoride, in
toothpaste), Ne (lighting), P (matches), S
(garden treatments), Cl (bleach
constituent), Ar (insulated windows), Se
(glass; solar cells), Br (as bromide, for
purification of spa water), Kr (energy
saving fluorescent lamps), I (in antiseptic
solutions), Xe (in plasma TV display cells,
a technology subsequently made
redundant by low cost OLED displays),
while Rn also sometimes occurs, but then
as an unwanted, potentially hazardous
indoor pollutant;[74] industrial acids: C, N, F,
P, S and Cl; inert air replacements: N, Ne, S
(in sulfur hexafluoride SF6), Ar, Kr and Xe;
lasers and lighting: He, C (in carbon dioxide
lasers, CO2), N, O (in a chemical oxygen
iodine laser), F (in a hydrogen fluoride
laser, HF), Ne, S (in a sulfur lamp), Ar, Kr
and Xe; and medicine and pharmaceuticals:
He, O, F, Cl, Br, I, Xe and Rn.

The number of compounds formed by


nonmetals is vast.[75] The first nine places
in a "top 20" table of elements most
frequently encountered in 8,427,300
compounds, as listed in the Chemical
Abstracts Service register for July 1987,
were occupied by nonmetals. Hydrogen,
carbon, oxygen and nitrogen were found in
the majority (greater than 64 per cent) of
compounds. Silicon, a metalloid, was in
10th place. The highest rated metal, with
an occurrence frequency of 2.3 per cent,
was iron, in 11th place.[76]

Discovery
Antiquity: C, S, (Sb) …

Carbon, sulfur, and antimony were known


in antiquity. The earliest known use of
charcoal dates to around 3750 BCE. The
Egyptians and Sumerians employed it for
the reduction of copper, zinc, and tin ores
in the manufacture of bronze. Diamonds
were probably known from as early as
2500 BCE. The first true chemical analyses
were made in the 18th century; Lavoisier
recognized carbon as an element in 1789.
Sulfur usage dates from before 2500 BCE;
it was recognized as an element by
Antoine Lavoisier in 1777. Antimony usage
was concurrent with that of sulfur; the
Louvre holds a 5,000 year old vase made
of almost pure antimony.

13th century: (As) …

Albertus Magnus (Albert the Great, 1193–


1280) is believed to have been the first to
isolate the element from a compound in
1250, by heating soap together with
arsenic trisulfide. If so, it was the first
element to be chemically discovered.

17th century: P …

Phosphorus was prepared from urine, by


Hennig Brand, in 1669.

18th century: H, O, N, (Te), Cl …

Hydrogen: Cavendish, in 1766, was the


first to distinguish hydrogen from other
gases, although Paracelsus around 1500,
Robert Boyle (1670), and Joseph Priestley
(?) had observed its production by reacting
strong acids with metals. Lavoisier named
it in 1793. Oxygen: Carl Wilhelm Scheele
obtained oxygen by heating mercuric oxide
and nitrates in 1771, but did not publish
his findings until 1777. Priestley also
prepared this new "air" by 1774, but only
Lavoisier recognized it as a true element;
he named it in 1777. Nitrogen: Rutherford
discovered nitrogen while he was studying
at the University of Edinburgh. He showed
that the air in which animals breathed,
after removal of exhaled carbon dioxide,
was no longer able to burn a candle.
Scheele, Henry Cavendish, and Priestley
also studied this element at about the
same time; Lavoisier named it in 1775 or
1776. Tellurium: In 1783, Franz-Joseph
Müller von Reichenstein, who was then
serving as the Austrian chief inspector of
mines in Transylvania, concluded that a
new element was present in a gold ore
from the mines in Zlatna, near today's city
of Alba Iulia, Romania. In 1789, a
Hungarian scientist, Pál Kitaibel,
discovered the element independently in
an ore from Deutsch-Pilsen that had been
regarded as argentiferous molybdenite,
but later he gave the credit to Müller. In
1798, it was named by Martin Heinrich
Klaproth, who had earlier isolated it from
the mineral calaverite. Chlorine: In 1774,
Scheele obtained chlorine from
hydrochloric acid but thought it was an
oxide. Only in 1808 did Humphry Davy
recognize it as an element.

Early 19th century: (B) I, Se, (Si), Br …


Boron was identified by Sir Humphry Davy
in 1808 but not isolated in a pure form
until 1909, by the American chemist
Ezekiel Weintraub. Iodine was discovered
in 1811 by Courtois from the ashes of
seaweed. Selenium: In 1817, when
Berzelius and Johan Gottlieb Gahn were
working with lead they discovered a
substance that was similar to tellurium.
After more investigation Berzelius
concluded that it was a new element,
related to sulfur and tellurium. Because
tellurium had been named for the Earth,
Berzelius named the new element
"selenium", after the moon. Silicon: In
1823, Berzelius prepared amorphous
silicon by reducing potassium
fluorosilicate with molten potassium
metal. Bromine: Balard and Gmelin both
discovered bromine in the autumn of 1825
and published their results in the following
year.

Late 19th century: He, F, (Ge), Ar, Kr,


Ne, Xe

Helium: In 1868, Janssen and Lockyer
independently observed a yellow line in the
solar spectrum that did not match that of
any other element. In 1895, in each case at
around the same time, Ramsay, Cleve, and
Langlet independently observed helium
trapped in cleveite. Fluorine: André-Marie
Ampère predicted an element analogous
to chlorine obtainable from hydrofluoric
acid, and between 1812 and 1886 many
researchers tried to obtain it. Fluorine was
eventually isolated in 1886 by Moissan.
Germanium: In mid-1885, at a mine near
Freiberg, Saxony, a new mineral was
discovered and named argyrodite because
of its silver content. The chemist Clemens
Winkler analyzed this new mineral, which
proved to be a combination of silver, sulfur,
and a new element, germanium, which he
was able to isolate in 1886. Argon: Lord
Rayleigh and Ramsay discovered argon in
1894 by comparing the molecular weights
of nitrogen prepared by liquefaction from
air, and nitrogen prepared by chemical
means. It was the first noble gas to be
isolated. Krypton, neon, and xenon: In
1898, within a period of three weeks,
Ramsay and Travers successively
separated krypton, neon and xenon from
liquid argon by exploiting differences in
their boiling points.

20th century: Rn, (At) …

In 1898, Friedrich Ernst Dorn discovered a


radioactive gas resulting from the
radioactive decay of radium; Ramsay and
Robert Whytlaw-Gray subsequently
isolated radon in 1910. Astatine was
synthesised in 1940 by Dale R. Corson,
Kenneth Ross MacKenzie, and Emilio
Segrè. They bombarded bismuth-209 with
alpha particles in a cyclotron to produce,
after emission of two neutrons, astatine-
211.

Notes
1. An ionisation energy of less than
750 kJ/mol is taken to be low, 750–
1000 is moderate, and > 1000 is high
(> 2000 is very high); an electron
affinity of less than 70 kJ/mol is taken
to be low, 70–140 is moderate, and
> 140 is high; an electronegativity of
less than 1.8 is taken to be low; 1.8–
2.2 is moderate; and > than 2.2 is high
(> 4.0 is very high).
2. Revised Pauling values are used for
the metalloids, and reactive
nonmetals; Allred-Rochow values for
the noble gases
3. The nonmetallic halogens (F, Cl, Br, I)
readily form anions including in
aqueous solution; the oxide ion O2− is
unstable in aqueous solution—its
affinity for H+ is so great that it
abstracts a proton from a solvent H2O
molecule (O2− + H2O → 2 OH−)—but is
found in an extensive series of metal
oxides
4. The common oxide is the most stable
oxide for that element

References

Data sources …
Unless otherwise stated, melting points,
boiling points, densities, crystalline
structures, ionisation energies, electron
affinities, and electronegativity values are
from the CRC Handbook of Physics and
Chemistry;[77] standard electrode
potentials are from the 1989 compilation
by Steven Bratsch.[78]

Citations …

1. Sukys 1999, p. 60.


2. Bettelheim et al. 2016, p. 33.
3. Schulze-Makuch & Irwin 2008, p. 89.
4. Steurer 2007, p. 7.
5. Cox 2004, p. 26
. Meyer et al. 2005, p. 284; Manahan
2001, p. 911; Szpunar et al. 2004, p. 17
7. Brown & Rogers 1987, p. 40
. Kneen, Rogers & Simpson 1972, p. 262
9. Greenwood & Earnshaw 2002, p. 434
10. Bratsch 1989; Bard, Parsons & Jordan
1985, p. 133
11. Yoder, Suydam & Snavely 1975, p. 58
12. Kneen, Rogers & Simpson 1972, p. 360
13. Lee 1996, p. 240
14. Greenwood & Earnshaw 2002, p. 43
15. Cressey 2010
1 . Siekierski & Burgess 2002, p. 24–25
17. Siekierski & Burgess 2002, p. 23
1 . Cox 2004, p. 146
19. Kneen, Rogers & Simpson 1972, p. 362
20. Bailar et al. 1989, p. 742
21. Stein 1983, p. 165
22. Jolly 1966, p. 20
23. Clugston & Flemming 2000, pp. 100–1,
104–5, 302
24. Seaborg 1969, p. 626
25. Nash 2005
2 . Scerri 2013, pp. 204–8
27. Challoner 2014, p. 5; Government of
Canada 2015; Gargaud et al. 2006,
p. 447
2 . Ivanenko et al. 2011, p. 784
29. Catling 2013, p. 12
30. Crawford 1968, p. 540
31. Berkowitz 2012, p. 293
32. Jørgensen & Mitsch 1983, p. 59
33. Wulfsberg 1987, p. 159–160
34. Bettelheim et al. 2016, p. 33—34
35. Field & Gray 2011, p. 12; see also
Myers, Oldham & Tocci 2004, pp. 120–
121 who categorize nonmetals as
hydrogen; semiconductors "(also
known as metalloids)"; less active
nonmetals (C, N, O, P, S, Se); halogens;
or noble gases
3 . Vernon 2020
37. Dingle 2017, pp. 101, 179
3 . Stein 1969; Pitzer 1975; Schrobilgen
2011
39. Brasted 1974, p. 814
40. Sidorov 1960
41. Rochow 1966, p. 4
42. Atkins 2006 et al., pp. 8, 122–23
43. Ritter 2011, p. 10
44. Wiberg 2001, p. 680
45. Wiberg 2001, p. 403
4 . Greenwood & Earnshaw 2002, p. 612
47. Moeller 1952, p. 208
4 . Cotton 2003, p. 205
49. Wulfsberg 1987, p. 159
50. Cronyn 2003
51. Stoyanov et al.
52. Rayner-Canham 2011, p. 126
53. Miller et al.
54. Rayner-Canham 2011, p. 126
55. Dillon, Mathey & Nixon 1998
5 . Martin-Louis et al. 2020
57. Wiberg 2001, p. 686
5 . Roy et al. 1994
59. Monteil & Vincent 1976
0. Moss 1952
1. Kent 2007, p. 104
2. Cacace, de Petris & Troiani 2002
3. Koziel 2002, p. 18
4. Piro et al. 2006
5. Steudel & Eckert 2003, p. 1
. Greenwood & Earnshaw 2002,
pp. 659–660
7. Moss 1952, p. 192; Greenwood &
Earnshaw 2002, p. 751
. Shanabrook, Lannin & Hisatsune 1981
9. Yousuf 1998, p. 425
70. Ostriker & Steinhardt 2001
71. Nelson 1987, p. 732
72. Emsley 2001, p. 428
73. Bolin 2012, p. 2-1
74. Maroni 1995
75. King & Caldwell 1954, p. 17; Brady &
Senese 2009, p. 69
7 . Nelson 1987, p. 735
77. Lide 2003
7 . Bratsch 1989

Bibliography …

Addison WE 1964, The allotropy of the


elements, Oldbourne Press, London
Arunan E, Desiraju GR, Klein RA, Sadlej J,
Scheiner S, Alkorta I, Clary DC, Crabtree RH,
Dannenberg JJ, Hobza P, Kjaergaard HG,
Legon AC, Mennucci B & Nesbitt DJ 2011,
"Defining the hydrogen bond: An account
(IUPAC Technical Report)", Pure and Applied
Chemistry, vol. 83, no. 8, pp. 1619–36,
doi:10.1351/PAC-REP-10-01-01
Ashford TA 1967, The physical sciences:
From atoms to stars, 2nd ed., Holt, Rinehart
and Winston, New York
Atkins P & de Paula J 2011, Physical
chemistry for the life sciences, 2nd ed.,
Oxford University Press, Oxford, ISBN 978-
1429231145
Aylward G & Findlay T 2008, SI chemical data,
6th ed., John Wiley & Sons Australia, Milton,
Queensland
Bailar JC, Moeller T, Kleinberg J, Guss CO,
Castellion ME & Metz C 1989, Chemistry, 3rd
ed., Harcourt Brace Jovanovich, San Diego,
ISBN 0-15-506456-8
Ball P 2013, "The name's bond", Chemistry
World, vol. 10, no. 6, p. 41
Bard AJ, Parsons R & Jordan J 1985,
Standard potentials in aqueous solution,
Marcel Dekker, New York, ISBN 978-0-8247-
7291-8
Berkowitz J 2012, The stardust revolution:
The new story of our origin in the stars,
Prometheus Books, Amherst, New York,
ISBN 978-1-61614-549-1
Bettelheim FA, Brown WH, Campbell MK,
Farrell SO 2010, Introduction to general,
organic, and biochemistry, 9th ed.,
Brooks/Cole, Belmont California,
ISBN 9780495391128
Bettelheim FA, Brown WH, Campbell MK,
Farrell SO & Torres OJ 2016, Introduction to
general, organic, and biochemistry, 11th ed.,
Cengage Learning, Boston, ISBN 978-1-285-
86975-9
Bogoroditskii NP & Pasynkov VV 1967, Radio
and electronic materials, Iliffe Books, London
Bolin P 2000, "Gas-insulated substations, in
JD McDonald (ed.), Electric power
substations engineering, 3rd, ed., CRC Press,
Boca Raton, FL, pp. 2–1–2-19,
ISBN 9781439856383
Borg RJ & Dienes GJ 1992, The physical
chemistry of solids, Academic Press, San
Diego, California, ISBN 9780121184209
Brady JE & Senese F 2009, Chemistry: The
study of matter and its changes, 5th ed., John
Wiley & Sons, New York,
ISBN 9780470576427
Bratsch SG 1989, "Standard electrode
potentials and temperature coefficients in
water at 298.15 K," Journal of Physical
Chemical Reference Data, vol. 18, no. 1,
pp. 1–21, doi:10.1063/1.555839
Brown WH & Rogers EP 1987, General,
organic and biochemistry, 3rd ed.,
Brooks/Cole, Monterey, California,
ISBN 0534068707
Bryson PD 1989, Comprehensive review in
toxicology, Aspen Publishers, Rockville,
Maryland, ISBN 0871897776
Bunge AV & Bunge CF 1979, "Electron affinity
of helium (1s2s)3S", Physical Review A, vol.
19, no. 2, pp. 452–456,
doi:10.1103/PhysRevA.19.452
Cacace F, de Petris G & Troiani A 2002,
"Experimental detection of tetranitrogen",
Science, vol. 295, no. 5554, pp. 480–81,
doi:10.1126/science.1067681
Cairns D 2012, Essentials of pharmaceutical
chemistry, 4th ed., Pharmaceutical Press,
London, ISBN 9780853699798
Cambridge Enterprise 2013, "Carbon 'candy
floss' could help prevent energy blackouts" ,
Cambridge University, viewed 28 August
2013
Catling DC 2013, Astrobiology: A very short
introduction, Oxford University Press, Oxford,
ISBN 978-0-19-958645-5
Challoner J 2014, The elements: The new
guide to the building blocks of our universe,
Carlton Publishing Group, ISBN 978-0-233-
00436-5
Chapman B & Jarvis A 2003, Organic
chemistry, kinetics and equilibrium, rev. ed.,
Nelson Thornes, Cheltenham, ISBN 978-0-
7487-7656-6
Chung DD 1987, "Review of exfoliated
graphite", Journal of Materials Science, vol.
22, pp. 4190–98, doi:10.1007/BF01132008
Clugston MJ & Flemming R 2000, Advanced
chemistry, Oxford University Press, Oxford,
ISBN 9780199146338
Conroy EH 1968, "Sulfur", in CA Hampel (ed.),
The encyclopedia of the chemical elements,
Reinhold, New York, pp. 665–680
Cotton FA, Darlington C & Lynch LD 1976,
Chemistry: An investigative approach,
Houghton Mifflin, Boston ISBN 978-0-395-
21671-2
Cotton S 2006, Lanthanide and actinide
chemistry, 2nd ed., John Wiley & Sons, New
York, ISBN 9780470010068
Cox T 2004, Inorganic chemistry, 2nd ed.,
BIOS Scientific Publishers, London, ISBN 1-
85996-289-0
Cracolice MS & Peters EI 2011, Basics of
introductory chemistry: An active learning
approach, 2nd ed., Brooks/Cole, Belmont
California, ISBN 9780495558507
Crawford FH 1968, Introduction to the science
of physics, Harcourt, Brace & World, New
York
Cressey 2010, "Chemists re-define hydrogen
bond ", Nature newsblog, accessed 23
August 2017
Cronyn MW 2003, "The proper place for
hydrogen in the periodic table", Journal of
Chemical Education, vol. 80, no. 8, pp. 947–
951, doi:10.1021/ed080p947
Daniel PL & Rapp RA 1976, "Halogen
corrosion of metals", in MG Fontana & RW
Staehle (eds), Advances in corrosion science
and technology, Springer, Boston, pp. 55–172,
doi:10.1007/978-1-4615-9062-0_2
DeKock RL & Gray HB 1989, Chemical
structure and bonding, 2nd ed., University
Science Books, Mill Valley, California,
ISBN 093570261X
Desch CH 1914, Intermetallic Compounds,
Longmans, Green and Co., New York
Dias RP, Yoo C, Kim M & Tse JS 2011,
"Insulator-metal transition of highly
compressed carbon disulfide," Physical
Review B, vol. 84, pp. 144104–1–6,
doi:10.1103/PhysRevB.84.144104
Dillon KB, Mathey F & Nixon JF 1998,
Phosphorus: The carbon copy: From
organophosphorus to phospha-organic
chemistry, John Wiley & Sons, Chichester
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 0-89874-230-7
Eagleson M 1994, Concise encyclopedia
chemistry, Walter de Gruyter, Berlin,
ISBN 3110114518
Eastman ED, Brewer L, Bromley LA, Gilles PW,
Lofgren NL 1950, "Preparation and properties
of refractory cerium sulfides", Journal of the
American Chemical Society, vol. 72, no. 5,
pp. 2248–50, doi:10.1021/ja01161a102
Emsley J 1971, The inorganic chemistry of the
non-metals, Methuen Educational, London,
ISBN 0423861204
Emsley J 2001, Nature's building blocks: An
A–Z guide to the elements , Oxford University
Press, Oxford, ISBN 0198503415
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
Field SQ & Gray T 2011, Theodore Gray's
elements vault, Black Dog & Leventhal
Publishers, New York, ISBN 978-1-57912-880-
7
Finney J 2015, Water: A Very Short
Introduction, Oxford University Press, Oxford,
ISBN 978-0198708728,
Fujimori T, Morelos-Gómez A, Zhu Z,
Muramatsu H, Futamura R, Urita K, Terrones
M, Hayashi T, Endo M, Hong SY, Choi YC,
Tománek D & Kaneko K 2013, "Conducting
linear chains of sulphur inside carbon
nanotubes" , Nature Communications, vol. 4,
article no. 2162, doi:10.1038/ncomms3162
Gargaud M, Barbier B, Martin H & Reisse J
(eds) 2006, Lectures in astrobiology, vol. 1,
part 1: The early Earth and other cosmic
habitats for life, Springer, Berlin, ISBN 3-540-
29005-2
Government of Canada 2015, Periodic table
of the elements , accessed 30 August 2015
Godfrin H & Lauter HJ 1995, "Experimental
properties of 3He adsorbed on graphite", in
WP Halperin (ed.), Progress in low
temperature physics, volume 14, pp. 213–320
(216–8), Elsevier Science B.V., Amsterdam,
ISBN 9780080539935
Greenwood NN & Earnshaw A 2002,
Chemistry of the elements, 2nd ed.,
Butterworth-Heinemann, ISBN 0750633654
Henderson W 2000, Main group chemistry,
Royal Society of Chemistry, Cambridge,
ISBN 9780854046171
Holderness A & Berry M 1979, Advanced level
inorganic chemistry, 3rd ed., Heinemann
Educational Books, London,
ISBN 9780435654351
Irving KE 2005, "Using chime simulations to
visualize molecules", in RL Bell & J Garofalo
(eds), Science units for Grades 9–12,
International Society for Technology in
Education, Eugene, Oregon,
ISBN 9781564842176
Ivanenko NB, Ganeev AA, Solovyev ND &
Moskvin LN 2011, "Determination of trace
elements in biological fluids", Journal of
Analytical Chemistry, vol. 66, no. 9, pp. 784–
799 (784), doi:10.1134/S1061934811090036
Jenkins GM & Kawamura K 1976, Polymeric
carbons—carbon fibre, glass and char,
Cambridge University Press, Cambridge,
ISBN 0521206936
Jolly WL 1966, The chemistry of the non-
metals, Prentice-Hall, Englewood Cliffs, New
Jersey
Jones WN 1969, Textbook of general
chemistry, C. V. Mosby Company, St Louis,
ISBN 978-0-8016-2584-8
Jorgensen CK 2012, Oxidation numbers and
oxidation states, Springer-Verlag, Berlin,
ISBN 978-3-642-87760-5
Jørgensen SE & Mitsch WJ (eds) 1983,
Application of ecological modelling in
environmental management, part A, Elsevier
Science Publishing, Amsterdam, ISBN 0-444-
42155-6
Keith JA & Jacob T 2010, "Computational
simulations on the oxygen reduction reaction
in electrochemical systems", in PB Balbuena
& VR Subramanian, Theory and experiment in
electrocatalysis, Modern aspects of
electrochemistry, vol. 50, Springer, New York,
pp. 89–132, ISBN 978-1-4419-5593-7
Kent JA 2007, Kent and Riegel's Handbook of
industrial chemistry and biotechnology, 11th
ed., vol. 1, Spring Science + Business Media,
New York, ISBN 978-0-387-27842-1
King RB 2004, "The metallurgist's periodic
table and the Zintl-Klemm concept", in DH
Rouvray & BR King (eds), The periodic table:
into the 21st century, Research Studies Press,
Philadelphia, pp. 189–206, ISBN 0863802923
King GB & Caldwell WE 1954, The
fundamentals of college chemistry, American
Book Company, New York
Kneen WR, Rogers MJW & Simpson P 1972,
Chemistry: Facts, patterns, and principles,
Addison-Wesley, London, ISBN 0201037793
Koziel JA 2002, "Sampling and sample
preparation for indoor air analysis", in J
Pawliszyn (ed.), Comprehensive analytical
chemistry, vol. 37, Elsevier Science B.V.,
Amsterdam, pp. 1–32, ISBN 0444505105
Krikorian OH & Curtis PG 1988, "Synthesis of
CeS and interactions with molten metals",
High Temperatures – High Pressures, vol. 20,
pp. 9–17, ISSN 0018-1544
Labes MM, Love P & Nichols LF 1979,
"Polysulfur nitride—a metallic,
superconducting polymer", Chemical Review,
vol. 79, no. 1, pp. 1–15,
doi:10.1021/cr60317a002
Lee JD 1996, Concise inorganic chemistry, 5th
ed., Blackwell Science, Oxford, ISBN 978-0-
6320-5293-6
Lide DR (ed.) 2003, CRC handbook of
chemistry and physics, 84th ed., CRC Press,
Boca Raton, Florida, Section 6, Fluid
properties; Vapor pressure,
ISBN 0849304849
Manahan SE 2001, Fundamentals of
environmental chemistry, 2nd ed., CRC Press,
Boca Raton, Florida, ISBN 156670491X
Maroni M, Seifert B & Lindvall T (eds) 1995,
"Physical pollutants", in Indoor air quality: A
comprehensive reference book, Elsevier,
Amsterdam, pp. 108–123, ISBN 0444816429
Martin RM & Lander GD 1946, Systematic
inorganic chemistry: From the standpoint of
the periodic law, 6th ed., Blackie & Son,
London
Martin-Louis Y, Riu RL, Jones WJ, Transue
PM & Cummins CC 2020, "Isolation of an
elusive phosphatetrahedrane", Science
Advances, vol. 6, no. 13,
doi:10.1126/sciadv.aaz3168
McCall BJ & Oka T 2003, "Enigma of H3+ in
diffuse interstellar clouds", in SL Guberman
(ed.), Dissociative recombination of molecular
ions with electrons, Springer
Science+Business Media, New York,
ISBN 978-1-4613-4915-0
McMillan PF 2006, "Solid-state chemistry: A
glass of carbon dioxide", Nature, vol. 441,
p. 823, doi:10.1038/441823a
Merchant SS & Helmann JD 2012, "Elemental
economy: Microbial strategies for optimizing
growth in the face of nutrient limitation", in
Poole RK (ed), Advances in Microbial
Physiology, vol. 60, pp. 91–210,
doi:10.1016/B978-0-12-398264-3.00002-4
Meyer JS, Adams WJ, Brix KV, Luoma SM,
Mount DR, Stubblefield WA & Wood CM (eds)
2005, Toxicity of dietborne metals to aquatic
organisms, Proceedings from the Pellston
Workshop on Toxicity of Dietborne Metals to
Aquatic Organisms, 27 July–1 August 2002,
Fairmont Hot Springs, British Columbia,
Canada, Society of Environmental Toxicology
and Chemistry, Pensacola, Florida,
ISBN 1880611708
Miller T 1987, Chemistry: a basic introduction,
4th ed., Wadsworth, Belmont, California,
ISBN 0534069126
Miller, T.S., Belen, A., Suter, T.M., Sella, A.,
Corà, A., McMillan, P.F.: Carbon nitrides:
synthesis and characterization of a new
class of functional materials. Phys. Chem.
Chem. Phys. 24, 15613–15638 (2017)
Mitchell JBA & McGowan JW 1983,
"Experimental studies of electron-ion
combination", Physics of ion-ion and electron-
ion collisions, F Brouillard F & JW McGowan
(eds), Plenum Press, ISBN 978-1-4613-3547-
4
Mitchell SC 2006, "Biology of sulfur", in SC
Mitchell (ed.), Biological interactions of sulfur
compounds, Taylor & Francis, London,
pp. 20–41, ISBN 0203375122
Moeller T 1952, Inorganic chemistry: An
advanced textbook, John Wiley & Sons, New
York
Monteil, Y., Vincent, H.: Phosphorous
compounds with the VI B group elements. Z.
Naturforsch. B. J. Chem. Sci. 31b(5), 668–
672 (1976)
Moss, T.S.: Photoconductivity in the
elements. Butterworths Scientific, London,
pp. 180, 202 (1952). For amorphous
selenium, the increase in conductivity is a
thousand-fold; for “metallic” selenium the
increase is from three to as much as two-
hundred fold. Mikla, V.I.; Mikla, V.V.:
Amorphous Chalcogenides: The Past,
Present and Future. Elsevier, Boston, p. 63
(2012); Yost, D.M., Russell, H. Systematic
Inorganic Chemistry of the Fifth-and-Sixth-
Group Nonmetallic Elements. Prentice-Hall,
New York, p. 282 (1946)
Murray PRS & Dawson PR 1976, Structural
and comparative inorganic chemistry: A
modern approach for schools and colleges,
Heinemann Educational Book, London,
ISBN 9780435656447
Myers RT, Oldham KB & Tocci S 2004, Holt
Chemistry, teacher ed., Holt, Rinehart &
Winston, Orlando, ISBN 0-03-066463-2
Nash CS 2005, "Atomic and molecular
properties of elements 112, 114, and 118",
Journal of Physical Chemistry A, vol. 109,
pp. 3493–500, doi:10.1021/jp050736o
Nelson PG 1987, "Important elements",
Journal of Chemical Education, vol. 68, no. 9,
pp. 732–737, doi:10.1021/ed068p732
Nelson PG 1998, "Classifying substances by
electrical character: An alternative to
classifying by bond type", Journal of Chemical
Education, vol. 71, no. 1, pp. 24–6,
doi:10.1021/ed071p24
Novak A 1979, "Vibrational spectroscopy of
hydrogen bonded systems", in TM
Theophanides (ed.), Infrared and Raman
spectroscopy of biological molecules,
proceedings of the NATO Advanced Study
Institute held at Athens, Greece, August 22–
31, 1978, D. Reidel Publishing Company,
Dordrecht, Holland, pp. 279–304,
ISBN 9027709661
+
Oka T 2006, "Interstellar H3", PNAS, vol. 103,
no. 33, doi:10.1073_pnas.0601242103
Ostriker JP & Steinhardt PJ 2001, "The
quintessential universe", Scientific American,
January, pp. 46–53
Oxtoby DW, Gillis HP & Campion A 2008,
Principles of modern chemistry , 6th ed.,
Thomson Brooks/Cole, Belmont, California,
ISBN 0534493661
Partington JR 1944, A text-book of inorganic
chemistry, 5th ed., Macmillan & Co., London
Patil UN, Dhumal NR & Gejji SP 2004,
"Theoretical studies on the molecular
electron densities and electrostatic
potentials in azacubanes", Theoretica
Chimica Acta, vol. 112, no. 1, pp 27–32,
doi:10.1007/s00214-004-0551-2
Patten MN 1989, Other metals and some
related materials, in MN Patten (ed.),
Information sources in metallic materials,
Bowker-Saur, London, ISBN 0408014911
Patterson CS, Kuper HS & Nanney TR 1967,
Principles of chemistry, Appleton Century
Crofts, New York
Pearson WB 1972, The crystal chemistry and
physics of metals and alloys, Wiley-
Interscience, New York, ISBN 0-471-67540-7
Pearson RG & Mawby RJ 1967, "The nature
of metal–halogen bonds", in V Gutmann
(ed.), Halogen chemistry, Academic Press,
pp. 55–84
Phifer C 2000, "Ceramics, glass structure and
properties", in Kirk-Othmer Encyclopedia of
Chemical Technology,
doi:10.1002/0471238961.071201191608090
6.a01
Phillips CSG & Williams RJP 1965, Inorganic
chemistry, I: Principles and non-metals,
Clarendon Press, Oxford
Piro NA, Figueroa JS, McKellar JT & Troiani
CC 2006, "Triple-bond reactivity of
diphosphorus molecules", Science, vol. 313,
no. 5791, pp. 1276–9,
doi:10.1126/science.1129630
Pitzer K 1975, "Fluorides of radon and
elements 118", Journal of the Chemical
Society, Chemical Communications, no. 18,
pp. 760–1, doi:10.1039/C3975000760B
Raju GG 2005, Gaseous Electronics: Theory
and Practice, CRC Press, Boca Raton, Florida,
ISBN 978-0-203-02526-0
Rao KY 2002, Structural chemistry of
glasses , Elsevier, Oxford, ISBN 0080439586
Rayner-Canham G 2011, "Isodiagonality in
the periodic table", Foundations of Chemistry,
vol. 13, no. 2, pp. 121–129,
doi:10.1007/s10698-011-9108-y
Rayner-Canham G & Overton T 2006,
Descriptive inorganic chemistry, 4th ed., WH
Freeman, New York, ISBN 0716789639
Regnault MV 1853, Elements of chemistry,
vol. 1, 2nd ed., Clark & Hesser, Philadelphia
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
Rodgers GE 2012, Descriptive inorganic,
coordination, & solid-state chemistry, 3rd ed.,
Brooks/Cole, Belmont, California,
ISBN 9780840068460
Roy, A.K., Burns, G.T., Grigora, S., Lie, G.C.:
Poly(alkyl/aryloxothiazenes), [N = S(O)R]n:
New direction in inorganic polymers. In:
Wisian-Neilson, P., Alcock, H.R., Wynne, K.J.
(eds.) Inorganic and Organometallic
Polymers II: Advanced Materials and
Intermediates, pp. 344–357. American
Chemical Society, Washington DC (1994)
Russell AM & Lee KL 2005, Structure-property
relations in nonferrous metals , Wiley-
Interscience, New York, ISBN 047164952X
Scerri E 2013, A tale of seven elements,
Oxford University Press, Oxford,
ISBN 9780195391312
Schaefer JC 1968, "Boron" in CA Hampel
(ed.), The encyclopedia of the chemical
elements, Reinhold, New York, pp. 73–81
Scharfe ME & Schmidlin FW 1975, "Charged
pigment xerography", in L Marton (ed.),
Advances in Electronics and Electron Physics,
vol. 38, Academic Press, New York, ISBN 0-
12-014538-3, pp. 93–147
Schrobilgen GJ 2011, "radon (Rn)", in
Encyclopædia Britannica , accessed 7 Aug
2011
Schulze-Makuch D & Irwin LN 2008, Life in
the Universe: Expectations and constraints,
2nd ed., Springer-Verlag, Berlin,
ISBN 9783540768166
Seaborg GT 1969, "Prospects for further
considerable extension of the periodic table",
Journal of Chemical Education, vol. 46, no. 10,
pp. 626–34, doi:10.1021/ed046p626
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, pp. 130–133
Sherwin E & Weston GJ 1966, Chemistry of
the non-metallic elements, Pergamon Press,
Oxford
Shipman JT, Wilson JD & Todd AW 2009, An
introduction to physical science, 12th ed.,
Houghton Mifflin Company, Boston,
ISBN 9780618935963
Siebring BR & Schaff ME 1980, General
chemistry, Wadsworth Publishing, Belmont,
California
Siekierski S & Burgess J 2002, Concise
chemistry of the elements, Horwood Press,
Chichester, ISBN 1-898563-71-3
Silvera I & Walraven JTM 1981, "Monatomic
hydrogen – a new stable gas", New Scientist,
22 January
Smith MB 2011, Organic Chemistry: An Acid—
Base Approach, CRC Press, Boca Raton,
Florida, ISBN 978-1-4200-7921-0
Stein L 1969, "Oxidized radon in halogen
fluoride solutions", Journal of the American
Chemical Society, vol. 19, no. 19, pp. 5396–7,
doi:10.1021/ja01047a042
Stein L 1983, "The chemistry of radon",
Radiochimica Acta, vol. 32, pp. 163–71
Steudel R 1977, Chemistry of the non-metals:
With an introduction to atomic structure and
chemical bonding, Walter de Gruyter, Berlin,
ISBN 3110048825
Steudel R 2003, "Liquid sulfur", in R Steudel
(ed.), Elemental sulfur and sulfur-rich
compounds I, Springer-Verlag, Berlin, pp. 81–
116, ISBN 9783540401919
Steudel R & Eckert B 2003, "Solid sulfur
allotropes", in R Steudel (ed.), Elemental
sulfur and sulfur-rich compounds I, Springer-
Verlag, Berlin, pp. 1–80,
ISBN 9783540401919
Steudel R & Strauss E 1984, "Homcyclic
selenium molecules and related cations", in
HJ Emeleus (ed.), Advances in inorganic
chemistry and radiochemistry, vol. 28,
Academic Press, Orlando, Florida, pp. 135–
167, ISBN 9780080578774
Steurer W 2007, "Crystal structures of the
elements" in JW Marin (ed.), Concise
encyclopedia of the structure of materials,
Elsevier, Oxford, pp. 127–45,
ISBN 0080451276
Stoyanov, E.S., Stoyanova, I.V., Reed, C.A.:
The structure of the hydrogen ion (H+aq). J.
Am. Chem. Soc. 132(5), 1484–1485 (2010)
Stwertka A 2012, A guide to the elements, 3rd
ed., Oxford University Press, Oxford,
ISBN 9780199832521
Sukys P 1999, Lifting the scientific veil:
Science appreciation for the nonscientist,
Rowman & Littlefield, Oxford,
ISBN 0847696006
Szpunar J, Bouyssiere B & Lobinski R 2004,
"Advances in analytical methods for
speciation of trace elements in the
environment", in AV Hirner & H Emons (eds),
Organic metal and metalloid species in the
environment: Analysis, distribution processes
and toxicological evaluation, Springer-Verlag,
Berlin, pp. 17–40, ISBN 3540208291
Taylor MD 1960, First principles of chemistry,
Van Nostrand, Princeton, New Jersey
Townes CH & Dailey BP 1952, "Nuclear
quadrupole effects and electronic structure
of molecules in the solid state", Journal of
Chemical Physics, vol. 20, pp.  35–40,
doi:10.1063/1.1700192
Van Setten MJ, Uijttewaal MA, de Wijs GA &
Groot RA 2007, 'Thermodynamic Stability of
Boron: The Role of Defects and Zero Point
Motion' , Journal of the American Chemical
Society, vol. 129, no. 9, pp. 2458–65,
doi:10.1021/ja0631246
Vernon RE 2020, "Organising the metals and
nonmetals," Foundations of Chemistry, pp.
1−17, doi:10.1007/s10698-020-09356-6
(open access)
Wells AF 1984, Structural inorganic chemistry,
5th ed., Clarendon Press, Oxfordshire,
ISBN 0198553706
Wiberg N 2001, Inorganic chemistry ,
Academic Press, San Diego,
ISBN 0123526515
Winkler MT 2009, "Non-equilibrium
chalcogen concentrations in silicon: Physical
structure, electronic transport, and
photovoltaic potential", PhD thesis, Harvard
University, Cambridge, Massachusetts
Winkler MT, Recht D, Sher M, Said AJ, Mazur
E & Aziz MJ 2011, "Insulator-to-metal
transition in sulfur-doped silicon", Physical
Review Letters, vol. 106, pp. 178701–4
Wulfsberg G 1987, Principles of descriptive
Inorganic chemistry, Brooks/Cole Publishing
Company, Monterey, California ISBN 0-534-
07494-4
Yoder CH, Suydam FH & Snavely FA 1975,
Chemistry, 2nd ed, Harcourt Brace
Jovanovich, New York, ISBN 978-0-15-
506470-6
Yousuf M 1998, "Diamond anvil cells in high-
pressure studies of semiconductors", in T
Suski & W Paul (eds), High pressure in
semiconductor physics II, Semiconductors
and semimetals, vol. 55, Academic Press,
San Diego, pp. 382–436,
ISBN 9780080864532
Yu PY & Cardona M 2010, Fundamentals of
semiconductors: Physics and materials
properties, 4th ed., Springer, Heidelberg,
ISBN 9783642007101
Zumdahl SS & DeCoste DJ 2013, Chemical
principles, 7th ed., Brooks/Cole, Belmont,
California, ISBN 9781111580650

Monographs
Emsley J 1971, The inorganic chemistry
of the non-metals, Methuen Educational,
London, ISBN 0423861204
Johnson RC 1966, Introductory
descriptive chemistry: selected
nonmetals, their properties, and behavior,
WA Benjamin, New York
Jolly WL 1966, The chemistry of the non-
metals, Prentice-Hall, Englewood Cliffs,
New Jersey
Powell P & Timms PL 1974, The
chemistry of the non-metals, Chapman &
Hall, London, ISBN 0470695706
Sherwin E & Weston GJ 1966, Chemistry
of the non-metallic elements, Pergamon
Press, Oxford
Steudel R 1977, Chemistry of the non-
metals: with an introduction to atomic
structure and chemical bonding, English
edition by FC Nachod & JJ Zuckerman,
Berlin, Walter de Gruyter,
ISBN 3110048825

External links
Media related to Nonmetals at
Wikimedia Commons

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Nonmetal&oldid=956390697"

Last edited 8 days ago by SharabSalam


Content is available under CC BY-SA 3.0 unless
otherwise noted.

You might also like