You are on page 1of 24

Oxidation state

The oxidation state, sometimes referred to as oxidation number, describes the degree of oxidation (loss of electrons) of an atom in a
chemical compound. Conceptually, the oxidation state, which may be positive, negative or zero, is the hypothetical charge that an
atom would have if all bonds to atoms of different elements were 100% ionic, with no covalent component. This is never exactly true
for real bonds.

The term oxidation was first used by Antoine Lavoisier to signify reaction of a substance with oxygen. Much later, it was realized that
the substance, upon being oxidized, loses electrons, and the meaning was extended to include other reactions in which electrons are
lost, regardless of whether oxygen was involved.

Oxidation states are typically represented by integers which may be positive, zero, or negative. In some cases, the average oxidation
state of an element is a fraction, such as 83 for iron in magnetite (Fe3 O4 ). The highest known oxidation state is reported to be +9 in the
tetroxoiridium(IX) cation (IrO+4).[1] It is predicted that even a +10 oxidation state may be achievable by platinum in the
tetroxoplatinum(X) cation (PtO2+ [2]
4 ). The lowest oxidation state is −5, as for boron in Al3 BC.
[3]

The increase in oxidation state of an atom, through a chemical reaction, is known as an oxidation; a decrease in oxidation state is
known as a reduction. Such reactions involve the formal transfer of electrons: a net gain in electrons being a reduction, and a net loss
of electrons being an oxidation. For pure elements, the oxidation state is zero.

The oxidation state of an atom does not represent the "real" charge on that atom, or any other actual atomic property. This is
particularly true of high oxidation states, where the ionization energy required to produce a multiply positive ion is far greater than the
energies available in chemical reactions. Additionally, oxidation states of atoms in a given compound may vary depending on the
choice of electronegativity scale used in their calculation. Thus, the oxidation state of an atom in a compound is purely a formalism. It
is nevertheless important in understanding the nomenclature conventions of inorganic compounds. Also, a number of observations
pertaining to chemical reactions may be explained at a basic level in terms of oxidation states.

In inorganic nomenclature, the oxidation state is represented by a Roman numeral placed after the element name inside a parenthesis
or as a superscript after the element symbol.

Contents
IUPAC definition
Determination
Simple approach without bonding considerations
Algorithm of assigning bonds
The algorithm's caveat
Algorithm of summing bond orders
Applied to a Lewis structure
Applied to bond graph
Balancing redox
Appearances
Nominal oxidation states
Ambiguous oxidation states
Oxidation-state determination from resonance formulas is not straightforward
A physical measurement is needed to decide the oxidation state
Truly ambiguous oxidation states occur
Fractional oxidation states
Elements with multiple oxidation states
Oxidation state in metals
List of oxidation states of the elements
Early forms (octet rule)
Use in nomenclature
History of the oxidation state concept
Early days
Use in nomenclature
Development towards the current concept
See also
References

IUPAC definition
IUPAC has published a "Comprehensive definition of the term oxidation state (IUPAC Recommendations 2016)".[4] It is a distillation
of an IUPAC technical report "Toward a comprehensive definition of oxidation state" from 2014.[5] The current IUPAC Gold Book
definition of oxidation state is:

Oxidation state of an atom is the charge of this atom after ionic approximation of its heteronuclear bonds...

— IUPAC[6]

and the term oxidation number is nearly synonymous.[7]

The underlying principle is that the ionic charge is "the oxidation state of an atom, after ionic approximation of its bonds"[8], where
ionic approximation means, hypothesizing that all bonds are ionic. Several criteria were considered for the ionic approximation:

1) Extrapolation of the bond’s polarity; a) from the electronegativity difference b) from the dipole moment c) from quantum‐chemical
calculations of charges

2) Assignment of electrons according to the atom’s contribution to the bonding MO[9][10]/ the electron's allegiance in a LCAO–MO
model[11].

In a bond between two different elements, the bond's electrons are assigned to its main atomic contributor/higher electronegativity; in
a bond between two atoms of the same element, the electrons are divided equally. This is because most electronegativity scales
depend on the atom’s bonding state, which makes the assignment of the oxidation state a somewhat circular argument. For example,
some scales may turn out unusual oxidation states, such as -6 for platinum in PtH4 -2 , for Pauling and Mulliken scales.[12]. The dipole
moments would, sometimes, also turn out abnormal oxidation numbers, such as in CO and NO, which are oriented with their positive
end towards oxygen. Therefore, this leaves the atom's contribution to the bonding MO, the atomic-orbital energy, and from quantum-
chemical calculations of charges, as the only viable criteria with cogent, values for ionic approximation. However, for a simple
estimate for the ionic approximation, we can use Allen electronegativities[13], as only that electronegativity scale is truly independent
of the oxidation state, as it relates to the average valence‐electron energy of the free atom:

Electronegativity using the Allen scale


Group → 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
↓ Period
H He
1
2.300 4.160
Li Be B C N O F Ne
2
0.912 1.576 2.051 2.544 3.066 3.610 4.193 4.787
Na Mg Al Si P S Cl Ar
3
0.869 1.293 1.613 1.916 2.253 2.589 2.869 3.242
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
4
0.734 1.034 1.19 1.38 1.53 1.65 1.75 1.80 1.84 1.88 1.85 1.588 1.756 1.994 2.211 2.424 2.685 2.966
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
5
0.706 0.963 1.12 1.32 1.41 1.47 1.51 1.54 1.56 1.58 1.87 1.521 1.656 1.824 1.984 2.158 2.359 2.582
Cs Ba Lu Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
6
0.659 0.881 1.09 1.16 1.34 1.47 1.60 1.65 1.68 1.72 1.92 1.765 1.789 1.854 2.01 2.19 2.39 2.60
Fr Ra
7
0.67 0.89
See also: Electronegativities of the elements (data page)

Determination
While introductory levels of chemistry teaching use postulated oxidation states, the IUPAC recommendation[4] and the Gold Book
entry[6] list two entirely general algorithms for the calculation of the oxidation states of elements in chemical compounds.
Simple approach without bonding considerations

Introductory chemistry uses postulates: the oxidation state for an element in a chemical formula is calculated from the overall charge
and postulated oxidation states for all the other atoms.

A simple example is based on two postulates,

1. OS = +1 for hydrogen
2. OS = −2 for oxygen

where OS stands for oxidation state. This approach yields correct oxidation states in oxides and hydroxides of any single element, and
in acids such as H2 SO4 or H2 Cr2 O7 . Its coverage can be extended either by a list of exceptions or by assigning priority to the
postulates. The latter works for H2 O2 where the priority of rule 1 leaves both oxygens with oxidation state −1.

Additional postulates and their ranking may expand the range of compounds to fit a textbook's scope. As an example, one postulatory
algorithm from many possible; in a sequence of decreasing priority:

1. An element in a free form has OS = 0.


2. In a compound or ion, the sum of the oxidation states equals the total charge of the compound or ion.
3. Fluorine in compounds has OS = −1; this extends to chlorine and bromine only when not bonded to a lighter
halogen, oxygen or nitrogen.
4. Group 1 and group 2 metals in compounds have OS = +1 and +2, respectively.
5. Hydrogen has OS = +1, but adopts −1 when bonded as a hydride to metals or metalloids.
6. Oxygen in compounds has OS = −2.

This set of postulates covers oxidation states of fluorides, chlorides, bromides, oxides, hydroxides and hydrides of any single element.
It covers all oxoacids of any central atom (and all their fluoro-, chloro- and bromo-relatives), as well as salts of such acids with group
1 and 2 metals. It also covers iodides, sulfides and similar simple salts of these metals.

Algorithm of assigning bonds

This algorithm is performed on a Lewis structure (a diagram that shows all valence electrons). Oxidation state equals the charge of an
atom after each of its heteronuclear bonds has been assigned to the more-electronegative partner of the bond (except when that partner
is a reversibly bonded Lewis-acid ligand) and homonuclear bonds have been divided equally:

where each "—" represents an electron pair (either shared between two atoms or solely on one atom), and "OS" is the oxidation state
as a numerical variable.

After the electrons have been assigned according to the vertical red lines on the formula, the total number of valence electrons that
now "belong" to each atom are subtracted from the number N of valence electrons of the neutral atom (such as 5 for nitrogen in group
15) to yield that atom's oxidation state.

This example shows the importance of describing the bonding. Its summary formula, HNO3 , corresponds to two structural isomers;
the peroxynitrous acid in the above figure and the more stable nitric acid. With the formula HNO3 , the simple approach without
bonding considerations yields −2 for all three oxygens and +5 for nitrogen, which is correct for nitric acid. For the peroxynitrous acid,
however, the two oxygens in the O–O bond each have OS = −1 and the nitrogen has OS = +3, which requires a structure to
understand.

Organic compounds are treated in a similar manner; exemplified here on functional groups occurring in between CH4 and CO2 :

Analogously for transition-metal compounds; CrO(O2 )2 on the left has a total of 36 valence electrons (18 pairs to be distributed), and
Cr(CO)6 on the right has 66 valence electrons (33 pairs):
A key step is drawing the Lewis structure of the molecule (neutral, cationic, anionic): atom symbols are arranged so that pairs of
atoms can be joined by single two-electron bonds as in the molecule (a sort of "skeletal" structure), and the remaining valence
electrons are distributed such that sp atoms obtain an octet (duet for hydrogen) with priority that increases with electronegativity. In
some cases, this leads to alternative formulae that differ in bond orders (the full set of which is called the resonance formulas).
Consider the sulfate anion (SO2− 4 with 32 valence electrons; 24 from oxygens, 6 from sulfur, 2 of the anion charge obtained from the
implied cation). The bond orders to the terminal oxygens have no effect on the oxidation state so long as the oxygens have octets.
Already the skeletal structure, top left, yields the correct oxidation states, as does the Lewis structure, top right (one of the resonance
formulas):

The bond-order formula at bottom is closest to the reality of four equivalent oxygens each having a total bond order of 2. That total
includes the bond of order 12 to the implied cation and follows the 8 − N rule[5] requiring that the main-group atom's bond order
equals 8 minus N valence electrons of the neutral atom, enforced with priority that increases with electronegativity.

This algorithm works equally for molecular cations composed of several atoms. An example is the ammonium cation of 8 valence
electrons (5 from nitrogen, 4 from hydrogens, minus 1 electron for the cation's positive charge):

Drawing Lewis structures with electron pairs as dashes emphasizes the essential equivalence of bond pairs and lone pairs when
counting electrons and moving bonds onto atoms. Structures drawn with electron dot pairs are of course identical in every way:

The algorithm's caveat

The algorithm contains a caveat, which concerns rare cases of transition-metal complexes with a type of ligand that is reversibly
bonded as a Lewis acid (as an acceptor of the electron pair from the transition metal); termed a "Z-type" ligand in Green's covalent
bond classification method. The caveat originates from the simplifying use of electronegativity instead of the MO-based electron
allegiance to decide the ionic sign.[4] One early example is the O2 S−RhCl(CO)(PPh3 )2 complex[14] with SO2 as the reversibly-
bonded acceptor ligand (released upon heating). The Rh−S bond is therefore extrapolated ionic against Allen electronegativities of
rhodium and sulfur, yielding oxidation state +1 for rhodium:
Algorithm of summing bond orders

This algorithm works on Lewis structures and on bond graphs of extended (non-molecular) solids:

Oxidation state is obtained by summing the heteronuclear-bond orders at the atom as positive if that atom is the
electropositive partner in a particular bond and as negative if not, and the atom’s formal charge (if any) is added to that
sum.

Applied to a Lewis structure

An example of a Lewis structure with no formal charge,

illustrates that, in this algorithm, homonuclear bonds are simply ignored (the bond orders are in blue).

Carbon monoxide exemplifies a Lewis structure with formal charges:

To obtain the oxidation states, the formal charges are summed with the bond-order value taken positively at the carbon and negatively
at the oxygen.

Applied to molecular ions, this algorithm considers the actual location of the formal (ionic) charge, as drawn in the Lewis structure.
As an example, summing bond orders in the ammonium cation yields −4 at the nitrogen of formal charge +1, with the two numbers
adding to the oxidation state of −3:

The sum of oxidation states in the ion equals its charge (as it equals zero for a neutral molecule).

Also in anions, the formal (ionic) charges have to be considered when nonzero. For sulfate this is exemplified with the skeletal or
Lewis structures (top), compared with the bond-order formula of all oxygens equivalent and fulfilling the octet and 8 − N rules
(bottom):
Applied to bond graph

A bond graph in solid-state chemistry is a chemical formula of an extended structure, in which direct bonding connectivities are
shown. An example is the AuORb3 perovskite, the unit cell of which is drawn on the left and the bond graph (with added numerical
values) on the right:

We see that the oxygen atom bonds to the six nearest rubidium cations, each of which has 4 bonds to the auride anion. The bond
graph summarizes these connectivities. The bond orders (also called bond valences) sum up to oxidation states according to the
attached sign of the bond's ionic approximation (there are no formal charges in bond graphs).

Determination of oxidation states from a bond graph can be illustrated on ilmenite, FeTiO3 . We may ask whether the mineral contains
Fe2+ and Ti4+, or Fe3+ and Ti3+. Its crystal structure has each metal atom bonded to six oxygens and each of the equivalent oxygens
to two irons and two titaniums, as in the bond graph below. Experimental data show that three metal–oxygen bonds in the octahedron
are short and three are long (the metals are off-center). The bond orders (valences), obtained from the bond lengths by the bond
valence method, sum up to 2.01 at Fe and 3.99 at Ti; which can be rounded off to oxidation states +2 and +4, respectively:

Balancing redox

Oxidation states can be useful for balancing chemical equations for oxidation–reduction (or redox) reactions, because the changes in
the oxidized atoms have to be balanced by the changes in the reduced atoms. For example, in the reaction of acetaldehyde with
Tollens' reagent to form acetic acid (shown below), the carbonyl carbon atom changes its oxidation state from +1 to +3 (loses two
electrons). This oxidation is balanced by reducing two Ag+ cations to Ag0 (gaining two electrons in total).

An inorganic example is the Bettendorf reaction using SnCl2 to prove the presence of arsenite ions in a concentrated HCl extract.
When arsenic(III) is present, a brown coloration appears forming a dark precipitate of arsenic, according to the following simplified
reaction:

2 As3+ + 3 Sn2+ → 2 As0 + 3 Sn4+


Here three tin atoms are oxidized from oxidation state +2 to +4, yielding six electrons that reduce two arsenic atoms from oxidation
state +3 to 0. The simple one-line balancing goes as follows: the two redox couples are written down as they react;

As3+ + Sn2+ ⇌ As0 + Sn4+ .

One tin is oxidized from oxidation state +2 to +4, a two-electron step, hence 2 is written in front of the two arsenic partners. One
arsenic is reduced from +3 to 0, a three-electron step, hence 3 goes in front of the two tin partners. An alternative three-line procedure
is to write separately the half-reactions for oxidation and for reduction, each balanced with electrons, and then to sum them up such
that the electrons cross out. In general, these redox balances (the one-line balance or each half-reaction) need to be checked for the
ionic and electron charge sums on both sides of the equation being indeed equal. If they are not equal, suitable ions are added to
balance the charges and the non-redox elemental balance.

Appearances

Nominal oxidation states

A nominal oxidation state is a general term for two specific purpose-oriented values:

Electrochemical oxidation state; it represents a molecule or ion in the Latimer diagram or Frost diagram for its redox-
active element. An example is the Latimer diagram for sulfur at pH 0 where the electrochemical oxidation state +2
for sulfur puts HS2O−3 between S and H2SO3:

Systematic oxidation state; it is chosen from close alternatives for pedagogical reasons of descriptive chemistry. An
example is the oxidation state of phosphorus in H3PO3 (which is in fact the diprotic HPO(OH)2) taken nominally as
+3, while Allen electronegativities of phosphorus and hydrogen suggest +5 by a narrow margin that makes the two
alternatives almost equivalent:

Both alternative oxidation states of phosphorus make chemical sense, depending on the chemical property or reaction we wish to
emphasize. In contrast, their average (+4) does not.

Ambiguous oxidation states

Lewis formulae are fine rule-based approximations of chemical reality, as indeed are Allen electronegativities. Still, oxidation states
may seem ambiguous when their determination is not straightforward. Rule-based oxidation states feel ambiguous when only
experiment can decide. There are also truly dichotomous values to be decided by mere convenience.

Oxidation-state determination from resonance formulas is not straightforward

Seemingly ambiguous oxidation states are obtained on a set of resonance formulas of equal weights for a molecule of heteronuclear
bonds where the atom connectivity does not correspond to the number of two-electron bonds dictated by the 8 − N rule. An example
is S2 N2 where four resonance formulas featuring one S=N double bond have oxidation states +2 and +4 on the two sulfur atoms, to
be averaged to +3 because the two sulfur atoms are equivalent in this square-shaped molecule.

A physical measurement is needed to decide the oxidation state


This happens when a non-innocent ligand is present, of hidden or unexpected redox properties that could otherwise
be assigned to the central atom. An example is the nickel dithiolate complex, Ni(S2C2H2)2−
2 .
[5]:1056–1057
When the redox ambiguity of a central atom and ligand yields dichotomous oxidation states of close stability,
thermally induced tautomerism may result, as exemplified by manganese catecholate, Mn(C6H4O2)3.[5]:1057–1058
Assignment of such oxidation states in general requires spectroscopic,[15] magnetic or structural data.
When the bond order has to be ascertained along an isolated tandem of a heteronuclear and a homonuclear bond.
An example is S2O2−3 with two oxidation-state alternatives (bond orders are in blue and formal charges in green):

The S–S distance in thiosulfate is needed to reveal that this bond order is very close to 1, as in the formula on the
left.

Truly ambiguous oxidation states occur


When the electronegativity difference between two bonded atoms is very small (as in H3PO3 above). Two almost
equivalent pairs of oxidation states, open for a choice, are obtained for these atoms.
When an electronegative p-block atom forms solely homonuclear bonds, the number of which differs from the
number of two-electron bonds suggested by rules. Examples are homonuclear finite chains like N−3 (the central
nitrogen connects two atoms while three two-electron bonds are required by 8 − N rule) or I−3 (the central iodine
connects two atoms while one two-electron bond fulfills the 8 − N rule). A sensible approach is to distribute the ionic
charge over the two outer atoms.[5] Such a placement of charges in a polysulfide S2− n (where all inner sulfurs form
two bonds, fulfilling the 8 − N rule) follows already from its Lewis structure.[5]
When the isolated tandem of a heteronuclear and a homonuclear bond leads to a bonding compromise in between
two Lewis structures of limiting bond orders. An example here is N2O:

The typically-used oxidation state of nitrogen in N2O is +1, which also obtains for both nitrogens by a molecular
orbital approach.[16] The formal charges on the right comply with electronegativities, and this implies an added
ionic bonding contribution. Indeed, the estimated N−N and N−O bond orders are 2.76 and 1.9, respectively,[5]
approaching the formula of integer bond orders that would include the ionic contribution explicitly as a bond (in
green):

Conversely, formal charges against electronegativities in a Lewis structure decrease the bond order of the
corresponding bond. An example is carbon monoxide with a bond-order estimate of 2.6.[17]

Fractional oxidation states

Fractional oxidation states are often used to represent the average oxidation state of several atoms of the same element in a structure.
For example, the formula of magnetite is Fe3 O4 , implying an average oxidation state for iron of + 83 .[18]:81–82 However, this average
value may not be representative if the atoms are not equivalent. In a Fe3 O4 crystal below 120 K (−153 °C), two-thirds of the cations
3+ 2+
are Fe and one-third are Fe , and the formula may be more specifically represented as FeO·Fe2 O3 .[19]

Likewise, propane, C3 H8 , has been described as having a carbon oxidation state of − 83 .[20] Again, this is an average value since the
structure of the molecule is H3 C−CH2 −CH3 , with the first and third carbon atoms each having an oxidation state of −3 and the
central one −2.

An example with true fractional oxidation states for equivalent atoms is potassium superoxide, KO2 . The diatomic superoxide ion O−2
has an overall charge of −1, so each of its two equivalent oxygen atoms is assigned an oxidation state of − 12 . This ion can be
described as a resonance hybrid of two Lewis structures, where each oxygen has oxidation state 0 in one structure and −1 in the other.
For the cyclopentadienyl anion C5 H−5, the oxidation state of C is −1 + − 15 = − 65 . The −1 occurs because each carbon is bonded to one
hydrogen atom (a less electronegative element), and the − 15 because the total ionic charge of −1 is divided among five equivalent
carbons. Again this can be described as a resonance hybrid of five equivalent structures, each having four carbons with oxidation state
−1 and one with −2.

Examples of fractional oxidation


states for carbon
Oxidation state Example species

− 65 C5H−5

− 67 C7H+7

+ 32 C4O2−
4

Elements with multiple oxidation states

Most elements have more than one possible oxidation state. For example, carbon has nine possible integer oxidation states from −4 to
+4:

Integer oxidation states of carbon


Oxidation state Example compound

−4 CH4

−3 C2H6

−2 C2H4, CH3Cl

−1 C2H2, C6H6, (CH2OH)2

0 HCHO, CH2Cl2

+1 OCHCHO, CHCl2CHCl2

+2 HCOOH, CHCl3

+3 HOOCCOOH, C2Cl6

+4 CCl4, CO2

Oxidation state in metals

Many compounds with luster and electrical conductivity maintain a simple stoichiometric formula; such as the golden TiO, blue-black
RuO2 or coppery ReO3 , all of obvious oxidation state. Ultimately, however, the assignment of the free metallic electrons to one of the
bonded atoms has its limits and leads to unusual oxidation states. Simple examples are the LiPb and Cu3 Au ordered alloys, the
composition and structure of which are largely determined by atomic size and packing factors. Should oxidation state be needed for
redox balancing, it is best set to 0 for all atoms of such an alloy.

List of oxidation states of the elements


This is a list of known oxidation states of the chemical elements, excluding nonintegral values. The most common states appear in
bold. The table is based on that of Greenwood and Earnshaw,[21] with additions noted. Every element exists in oxidation state 0 when
it is the pure non-ionized element in any phase, whether monatomic or polyatomic allotrope. The column for oxidation state 0 only
shows elements known to exist in oxidation state 0 in compounds.

Noble gas
+1 Bold values are main oxidation states
Oxidation states of the elements
Negative states Positive states
Element Group Notes
−5 −4 −3 −2 −1 0 +1 +2 +3 +4 +5 +6 +7 +8 +9
Z
1 hydrogen H −1 +1 1
2 helium He 18

3 lithium Li +1 1 [22]

4 beryllium Be 0 +1 +2 2 [23][24]

5 boron B −5 −1 0 +1 +2 +3 13 [25][26][27]

6 carbon C −4 −3 −2 −1 0 +1 +2 +3 +4 14
7 nitrogen N −3 −2 −1 +1 +2 +3 +4 +5 15
8 oxygen O −2 −1 0 +1 +2 16
9 fluorine F −1 17
10 neon Ne 18

11 sodium Na −1 +1 1 [22]

12 magnesium Mg +1 +2 2 [28]

13 aluminium Al −2 −1 +1 +2 +3 13 [29][30][31]

14 silicon Si −4 −3 −2 −1 0 +1 +2 +3 +4 14 [32]

15 phosphorus P −3 −2 −1 0 +1 +2 +3 +4 +5 15 [33]

16 sulfur S −2 −1 0 +1 +2 +3 +4 +5 +6 16

17 chlorine Cl −1 +1 +2 +3 +4 +5 +6 +7 17 [34]

18 argon Ar 0 18 [35]

19 potassium K −1 +1 1 [22]

20 calcium Ca +1 +2 2 [36]

21 scandium Sc 0 +1 +2 +3 3 [37][38][39]

22 titanium Ti −2 −1 0 +1 +2 +3 +4 4 [40][41][42][43]

23 vanadium V −3 −1 0 +1 +2 +3 +4 +5 5 [41]

24 chromium Cr −4 −2 −1 0 +1 +2 +3 +4 +5 +6 6 [41]

25 manganese Mn −3 −2 −1 0 +1 +2 +3 +4 +5 +6 +7 7

26 iron Fe −4 −2 −1 0 +1 +2 +3 +4 +5 +6 +7 8 [44][45][46]

27 cobalt Co −3 −1 0 +1 +2 +3 +4 +5 9 [41]

28 nickel Ni −2 −1 0 +1 +2 +3 +4 10 [47]

29 copper Cu −2 0 +1 +2 +3 +4 11 [46][48]

30 zinc Zn −2 +1 +2 12 [46][49]

31 gallium Ga −5 −4 −3 −2 −1 +1 +2 +3 13 [30][50][51]

32 germanium Ge −4 −3 −2 −1 0 +1 +2 +3 +4 14 [52][32]

33 arsenic As −3 −2 −1 0 +1 +2 +3 +4 +5 15 [30][53][54][55]

34 selenium Se −2 −1 +1 +2 +3 +4 +5 +6 16 [56][57][58][59]

35 bromine Br −1 +1 +3 +4 +5 +7 17
36 krypton Kr 0 +1 +2 18

37 rubidium Rb −1 +1 1 [22]

38 strontium Sr +1 +2 2 [60]

39 yttrium Y 0 +1 +2 +3 3 [61][62][63]

40 zirconium Zr −2 +1 +2 +3 +4 4 [41][64]
41 niobium Nb −3 −1 +1 +2 +3 +4 +5 5 [41][65]

42 molybdenum Mo −4 −2 −1 0 +1 +2 +3 +4 +5 +6 6 [41]

43 technetium Tc −3 −1 0 +1 +2 +3 +4 +5 +6 +7 7

44 ruthenium Ru −4 −2 0 +1 +2 +3 +4 +5 +6 +7 +8 8 [41][46]

45 rhodium Rh −3 −1 0 +1 +2 +3 +4 +5 +6 9 [41][66]

46 palladium Pd 0 +1 +2 +3 +4 10 [67][68]

47 silver Ag −2 −1 +1 +2 +3 11 [46][69]

48 cadmium Cd −2 +1 +2 12 [46][70]

49 indium In −5 −2 −1 +1 +2 +3 13 [30][71][72]

50 tin Sn −4 −3 −2 −1 0 +1 +2 +3 +4 14 [30][73][74][32]

51 antimony Sb −3 −2 −1 0 +1 +2 +3 +4 +5 15 [30][75][76][77][78]

52 tellurium Te −2 −1 +1 +2 +3 +4 +5 +6 16 [30][79][80][81]

53 iodine I −1 +1 +3 +4 +5 +6 +7 17 [82][83]

54 xenon Xe 0 +1 +2 +4 +6 +8 18 [84][85][86]

55 caesium Cs −1 +1 1 [22]

56 barium Ba +1 +2 2 [87]

57 lanthanum La 0 +1 +2 +3 3 [61][88]

58 cerium Ce +2 +3 +4 n/a

59 praseodymium Pr 0 +1 +2 +3 +4 +5 n/a [61][89][90][91]

60 neodymium Nd 0 +2 +3 +4 n/a [61][92]

61 promethium Pm +2 +3 n/a [93]

62 samarium Sm 0 +2 +3 n/a [61]

63 europium Eu +2 +3 n/a

64 gadolinium Gd 0 +1 +2 +3 n/a [61]

65 terbium Tb 0 +1 +2 +3 +4 n/a [61][93]

66 dysprosium Dy 0 +2 +3 +4 n/a [61][94]

67 holmium Ho 0 +2 +3 n/a [61][93]

68 erbium Er 0 +2 +3 n/a [61][93]

69 thulium Tm +2 +3 n/a
70 ytterbium Yb +2 +3 n/a

71 lutetium Lu 0 +2 +3 n/a [61][93]

72 hafnium Hf −2 +1 +2 +3 +4 4 [41][95]

73 tantalum Ta −3 −1 +1 +2 +3 +4 +5 5 [41][65]

74 tungsten W −4 −2 −1 0 +1 +2 +3 +4 +5 +6 6 [41]

75 rhenium Re −3 −1 0 +1 +2 +3 +4 +5 +6 +7 7

76 osmium Os −4 −2 −1 0 +1 +2 +3 +4 +5 +6 +7 +8 8 [46][96]

77 iridium Ir −3 −1 0 +1 +2 +3 +4 +5 +6 +7 +8 +9 9 [97][98][99][100]

78 platinum Pt −3 −2 −1 0 +1 +2 +3 +4 +5 +6 10 [46][101][102]

79 gold Au −3 −2 −1 0 +1 +2 +3 +5 11 [46][103]

80 mercury Hg −2 +1 +2 12 [46][104]

81 thallium Tl −5 −2 −1 +1 +2 +3 13 [30][105][106][107]

82 lead Pb −4 −2 −1 +1 +2 +3 +4 14 [30][108][109]

83 bismuth Bi −3 −2 −1 +1 +2 +3 +4 +5 15 [110][111][112][113]

84 polonium Po −2 +2 +4 +5 +6 16 [114]
85 astatine At −1 +1 +3 +5 +7 17

86 radon Rn +2 +6 18 [115][116][117]

87 francium Fr +1 1
88 radium Ra +2 2
89 actinium Ac +3 3

90 thorium Th +1 +2 +3 +4 n/a [118][119]

91 protactinium Pa +3 +4 +5 n/a

92 uranium U +1 +2 +3 +4 +5 +6 n/a [120][121]

93 neptunium Np +2 +3 +4 +5 +6 +7 n/a [122]

94 plutonium Pu +2 +3 +4 +5 +6 +7 n/a [123]

95 americium Am +2 +3 +4 +5 +6 +7 n/a [124]

96 curium Cm +3 +4 +5 +6 n/a [125][126][127][128]

97 berkelium Bk +2 +3 +4 +5 n/a [125][126][129][130][131]

98 californium Cf +2 +3 +4 +5 n/a [125][126]

99 einsteinium Es +2 +3 +4 n/a [132]

100 fermium Fm +2 +3 n/a


101 mendelevium Md +2 +3 n/a
102 nobelium No +2 +3 n/a
103 lawrencium Lr +3 n/a
104 rutherfordium Rf +4 4

105 dubnium Db +5 5 [133]

106 seaborgium Sg 0 +6 6 [134][135]

107 bohrium Bh +7 7 [136]

108 hassium Hs +8 8 [137]

109 meitnerium Mt 9
110 darmstadtium Ds 10
111 roentgenium Rg 11

112 copernicium Cn +2 12 [138]

113 nihonium Nh 13
114 flerovium Fl 14
115 moscovium Mc 15
116 livermorium Lv 16
117 tennessine Ts 17
118 oganesson Og 18

Early forms (octet rule)

A figure with a similar format was used by Irving Langmuir in 1919 in one of the early papers about the octet rule.[139] The
periodicity of the oxidation states was one of the pieces of evidence that led Langmuir to adopt the rule.
Use in nomenclature
The oxidation state in compound naming is placed either as a right superscript to the element symbol in a chemical formula, such as
FeIII, or in parentheses after the name of the element in chemical names, such as iron(III). For example, Fe2 (SO4 )3 is named iron(III)
III
sulfate and its formula can be shown as Fe2 (SO4 )3 . This is because a sulfate ion has a charge of −2, so each iron atom takes a charge
of +3. Fractional oxidation numbers should not be used in naming.[140]:66 Red lead, Pb3 O4 , is represented as lead(II,IV) oxide,
showing the actual two oxidation states of the nonequivalent lead atoms.

History of the oxidation state concept

Early days

Oxidation itself was first studied by Antoine Lavoisier, who defined it as the result of reactions with oxygen (hence the
name).[141][142] The term has since been generalized to imply a formal loss of electrons. Oxidation states, called oxidation grades by
Friedrich Wöhler in 1835,[143] were one of the intellectual stepping stones that Dmitri Mendeleev used to derive the periodic table.
Jensen[144] gives an overview of the history up to 1938.

Use in nomenclature

When it was realized that some metals form two different binary compounds with the same nonmetal, the two compounds were often
distinguished by using the ending -ic for the higher metal oxidation state and the ending -ous for the lower. For example, FeCl3 is
ferric chloride and FeCl2 is ferrous chloride. This system is not very satisfactory (although sometimes still used) because different
metals have different oxidation states which have to be learned: ferric and ferrous are +3 and +2 respectively, but cupric and cuprous
are +2 and +1, and stannic and stannous are +4 and +2. Also there was no allowance for metals with more than two oxidation states,
such as vanadium with oxidation states +2, +3, +4 and +5.[18]:84

This system has been largely replaced by one suggested by Alfred Stock in 1919[145] and adopted[146] by IUPAC in 1940. Thus,
FeCl2 was written as iron(II) chloride rather than ferrous chloride. The Roman numeral II at the central atom came to be called the
"Stock number" (now an obsolete term), and its value was obtained as a charge at the central atom after removing its ligands along
with the electron pairs they shared with it.[140]:147

Development towards the current concept

The term "oxidation state" in English chemical literature was popularized by Wendell Mitchell Latimer in his 1938 book about
electrochemical potentials.[147] He used it for the value (synonymous with the German term Wertigkeit) previously termed "valence",
"polar valence" or "polar number"[148] in English, or "oxidation stage" or indeed[149][150] the "state of oxidation". Since 1938, the
term "oxidation state" has been connected with electrochemical potentials and electrons exchanged in redox couples participating in
redox reactions. By 1948, IUPAC used the 1940 nomenclature rules with the term "oxidation state",[151][152] instead of the
original[146] valency. In 1948 Linus Pauling proposed that oxidation number could be determined by extrapolating bonds to being
completely ionic in the direction of electronegativity.[153] A full acceptance of this suggestion was complicated by the fact that the
Pauling electronegativities as such depend on the oxidation state and that they may lead to unusual values of oxidation states for some
transition metals. In 1990 IUPAC resorted to a postulatory (rule-based) method to determine the oxidation state.[154] This was
complemented by the synonymous term oxidation number as a descendant of the Stock number introduced in 1940 into the
nomenclature. However, the terminology using "ligands"[140]:147 gave the impression that oxidation number might be something
specific to coordination complexes. This situation and the lack of a real single definition generated numerous debates about the
meaning of oxidation state, suggestions about methods to obtain it and definitions of it. To resolve the issue, an IUPAC project (2008-
040-1-200) was started in 2008 on the "Comprehensive Definition of Oxidation State", and was concluded by two reports[5][4] and
by the revised entries "Oxidation State"[6] and "Oxidation Number"[7] in the IUPAC Gold Book. The outcomes were a single
definition of oxidation state and two algorithms to calculate it in molecular and extended-solid compounds, guided by Allen
electronegativities that are independent of oxidation state.

See also
Electrochemistry

References
1. Wang, G.; Zhou, M.; Goettel, G. T.; Schrobilgen, G. J.; Su, J.; Li, J.; Schlöder, T.; Riedel, S. (2014). "Identification of
an iridium-containing compound with a formal oxidation state of IX". Nature. 514 (7523): 475–477.
doi:10.1038/nature13795 (https://doi.org/10.1038%2Fnature13795). PMID 25341786 (https://pubmed.ncbi.nlm.nih.g
ov/25341786).
2. Yu, H.-S.; Truhlar, D. G. (2016). "Oxidation state 10 exists" (https://doi.org/10.1002/anie.201604670). Angew. Chem.
Int. Ed. 55 (31): 9004–9006. doi:10.1002/anie.201604670 (https://doi.org/10.1002%2Fanie.201604670).
3. Schroeder, Melanie, Eigenschaften von borreichen Boriden und Scandium-Aluminium-Oxid-Carbiden (https://d-nb.i
nfo/995006210/34) (in German), p. 139
4. Karen, P.; McArdle, P.; Takats, J. (2016). "Comprehensive definition of oxidation state (IUPAC Recommendations
2016)". Pure Appl. Chem. 88 (8): 831–839. doi:10.1515/pac-2015-1204 (https://doi.org/10.1515%2Fpac-2015-1204).
hdl:10852/59520 (https://hdl.handle.net/10852%2F59520).
5. Karen, P.; McArdle, P.; Takats, J. (2014). "Toward a comprehensive definition of oxidation state (IUPAC Technical
Report)" (https://doi.org/10.1515/pac-2013-0505). Pure Appl. Chem. 86 (6): 1017–1081. doi:10.1515/pac-2013-0505
(https://doi.org/10.1515%2Fpac-2013-0505).
6. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "Oxidation state (https://goldbook.iupac.org/O04365.html)". doi:10.1351/goldbook.O04365 (https://doi.org/1
0.1351%2Fgoldbook.O04365)
7. IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "Oxidation number (https://goldbook.iupac.org/O04363.html)". doi:10.1351/goldbook.O04363 (https://doi.or
g/10.1351%2Fgoldbook.O04363)
8. https://doi.org/10.1002/anie.201407561
9. https://doi.org/10.1002/anie.201407561
10. Hooydonk, G. (1974). O n an Ionic Approximation to Chemical Bonding, Zeitschrift für Naturforschung A, 29(5), 763-
767. doi: https://doi.org/10.1515/zna-1974-0517
11. https://doi.org/10.1351/goldbook.O04365
12. Pure and Applied Chemistry (2014), 86(6), 1017-1081 CODEN: PACHAS; ISSN: 0033-4545. English.
13. https://doi.org/10.1002/anie.201407561
14. Muir, K. W.; Ibers, J. A. (1969). "The structure of chlorocarbonyl(sulfur dioxide)bis(triphenylphosphine)rhodium,
RhCl(CO)(SO2)(P(C6H5)3)2". Inorg. Chem. 8: 1921–1928. doi:10.1021/ic50079a024 (https://doi.org/10.1021%2Fic
50079a024).
15. Jørgensen, C. K. (1966). "Electric Polarizability, Innocent Ligands and Spectroscopic Oxidation States". Structure
and Bonding. 1. Berlin: Springer-Verlag. pp. 234–248.
16. Karen, P. (2015). "Oxidation state, a long-standing issue!" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4506524).
Angew. Chem. Int. Ed. 54 (16): 4716–4726. doi:10.1002/anie.201407561 (https://doi.org/10.1002%2Fanie.2014075
61). PMC 4506524 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4506524). PMID 25757151 (https://pubmed.ncbi.
nlm.nih.gov/25757151).
17. Martinie, R. J.; Bultema, J. J.; Wal, M. N. V.; Burkhart, B. J.; Griend, D. A. V.; DeCock, R. L. (2011). "Bond order and
chemical properties of BF, CO, and N2". J. Chem. Educ. 88: 1094–1097. doi:10.1021/ed100758t (https://doi.org/10.1
021%2Fed100758t).
18. Petrucci, R. H.; Harwood, W. S.; Herring, F. G. (2002). General Chemistry (https://archive.org/details/generalchemistr
y00hill) (8th ed.). Prentice-Hall.
19. Senn, M. S.; Wright, J. P.; Attfield, J. P. (2012). "Charge order and three-site distortions in the Verwey structure of
magnetite" (https://www.pure.ed.ac.uk/ws/files/10796489/Charge_order_and_three_site_distortions_in_the_Verwey
_structure_of_magnetite.pdf) (PDF). Nature. 481 (7380): 173–6. doi:10.1038/nature10704 (https://doi.org/10.1038%
2Fnature10704). PMID 22190035 (https://pubmed.ncbi.nlm.nih.gov/22190035).
20. Whitten, K. W.; Galley, K. D.; Davis, R. E. (1992). General Chemistry (https://archive.org/details/generalchemistry00
whit_0) (4th ed.). Saunders. p. 147 (https://archive.org/details/generalchemistry00whit_0/page/147).
21. Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann.
pp. 27–28. ISBN 978-0-08-037941-8.
22. Na(−1), K(−1), Rb(−1), and Cs(−1) are known in alkalides; the table by Greenwood and Earnshaw shows −1 only for
Na and also erroneously for Li; no lithides are described.
23. Be(0) has been observed; see "Beryllium(0) Complex Found" (https://www.chemistryviews.org/details/news/942600
1/Beryllium0_Complex_Found.html). ChemistryViews. 13 June 2016.
24. Be(I) has been observed in beryllium monohydride (BeH); see Shayesteh, A.; Tereszchuk, K.; Bernath, P. F.; Colin,
R. (2003). "Infrared Emission Spectra of BeH and BeD" (https://web.archive.org/web/20071202130342/http://bernat
h.uwaterloo.ca/media/252.pdf) (PDF). J. Chem. Phys. 118 (3): 1158. Bibcode:2003JChPh.118.1158S (https://ui.ads
abs.harvard.edu/abs/2003JChPh.118.1158S). doi:10.1063/1.1528606 (https://doi.org/10.1063%2F1.1528606).
Archived from the original (http://bernath.uwaterloo.ca/media/252.pdf) (PDF) on 2007-12-02. Retrieved 2007-12-10.
25. B(−5) has been observed in Al3BC, see Schroeder, Melanie. "Eigenschaften von borreichen Boriden und
Scandium-Aluminium-Oxid-Carbiden" (https://d-nb.info/995006210/34) (in German). p. 139.
26. B(−1) has been observed in magnesium diboride (MgB2), see Keeler, James; Wothers, Peter (2014). Chemical
Structure and Reactivity: An Integrated Approach (https://books.google.com/books?id=2RgbAgAAQBAJ&pg=PA315
&lpg=PA315#v=onepage&q&f=false). Oxford University Press. ISBN 9780199604135.
27. B(0) has been observed in diborynes, see Braunschweig, H.; Dewhurst, R. D.; Hammond, K.; Mies, J.; Radacki, K.;
Vargas, A. (2012). "Ambient-Temperature Isolation of a Compound with a Boron-Boron Triple Bond". Science. 336
(6087): 1420–2. Bibcode:2012Sci...336.1420B (https://ui.adsabs.harvard.edu/abs/2012Sci...336.1420B).
doi:10.1126/science.1221138 (https://doi.org/10.1126%2Fscience.1221138). PMID 22700924 (https://pubmed.ncbi.
nlm.nih.gov/22700924).
28. Low valent magnesium compounds with Mg(I) have been obtained using bulky ligands; see Green, S. P.; Jones C.;
Stasch A. (December 2007). "Stable Magnesium(I) Compounds with Mg-Mg Bonds". Science. 318 (5857): 1754–
1757. Bibcode:2007Sci...318.1754G (https://ui.adsabs.harvard.edu/abs/2007Sci...318.1754G).
doi:10.1126/science.1150856 (https://doi.org/10.1126%2Fscience.1150856). PMID 17991827 (https://pubmed.ncbi.
nlm.nih.gov/17991827).
29. Al(II) has been observed in aluminium(II) oxide (AlO); see Tyte, D.C. (1964). "Red (B2Π–A2σ) Band System of
Aluminium Monoxide". Nature. 202 (4930): 383–384. Bibcode:1964Natur.202..383T (https://ui.adsabs.harvard.edu/a
bs/1964Natur.202..383T). doi:10.1038/202383a0 (https://doi.org/10.1038%2F202383a0), and in dialanes (R2Al—
AlR2); see Uhl, Werner (2004). "Organoelement Compounds Possessing Al—Al, Ga—Ga, In—In, and Tl—Tl Single
Bonds". Advances in Organometallic Chemistry. Volume 51: 53–108. doi:10.1016/S0065-3055(03)51002-4 (https://d
oi.org/10.1016%2FS0065-3055%2803%2951002-4).
30. Negative oxidation states of p-block metals (Al, Ga, In, Sn, Tl, Pb, Bi, Po) and metalloids (Si, Ge, As, Sb, Te, At) may
occur in Zintl phases, see: Riedel, Erwin, ed. (2007). Moderne Anorganische Chemie (in German). p. 259, and
"Vorlesung Intermetallische Phasen § 6.2 Binäre Zintl-Phasen" (http://ruby.chemie.uni-freiburg.de/Vorlesung/interme
tallische_6_2.html) (in German).
31. Al(−2) has been observed in Sr14[Al4]2[Ge]3, see Wemdorff, Marco; Röhr, Caroline (2007). "Sr14[Al4]2[Ge]3: Eine
Zintl-Phase mit isolierten [Ge]4–- und [Al4]8–-Anionen / Sr14[Al4]2[Ge]3: A Zintl Phase with Isolated [Ge]4–- and
[Al4]8– Anions". Zeitschrift für Naturforschung B (in German). 62 (10): 1227. doi:10.1515/znb-2007-1001 (https://doi.
org/10.1515%2Fznb-2007-1001).
32. "New Type of Zero-Valent Tin Compound" (https://www.chemistryviews.org/details/news/9745121/New_Type_of_Z
ero-Valent_Tin_Compound.html). ChemistryViews. 27 August 2016.
33. P(0) has been observed, see Wang, Yuzhong; Xie, Yaoming; Wei, Pingrong; King, R. Bruce; Schaefer, Iii; Schleyer,
Paul v. R.; Robinson, Gregory H. (2008). "Carbene-Stabilized Diphosphorus". Journal of the American Chemical
Society. 130 (45): 14970–1. doi:10.1021/ja807828t (https://doi.org/10.1021%2Fja807828t). PMID 18937460 (https://
pubmed.ncbi.nlm.nih.gov/18937460).
34. The equilibrium Cl2O6⇌2ClO3 is mentioned by Greenwood and Earnshaw, but it has been refuted, see Lopez,
Maria; Juan E. Sicre (1990). "Physicochemical properties of chlorine oxides. 1. Composition, ultraviolet spectrum,
and kinetics of the thermolysis of gaseous dichlorine hexoxide". J. Phys. Chem. 94 (9): 3860–3863.
doi:10.1021/j100372a094 (https://doi.org/10.1021%2Fj100372a094)., and Cl2O6 is actually chlorine(V,VII) oxide.
However, ClO3 has been observed, see Grothe, Hinrich; Willner, Helge (1994). "Chlorine Trioxide: Spectroscopic
Properties, Molecular Structure, and Photochemical Behavior". Angew. Chem. Int. Ed. 33 (14): 1482–1484.
doi:10.1002/anie.199414821 (https://doi.org/10.1002%2Fanie.199414821).
35. Ar(0) has been observed in argon fluorohydride (HArF) and ArCF22+ , see Lockyear, J.F.; Douglas, K.; Price, S.D.;
Karwowska, M.; et al. (2010). "Generation of the ArCF22+ Dication". Journal of Physical Chemistry Letters. 1: 358.
doi:10.1021/jz900274p (https://doi.org/10.1021%2Fjz900274p).
36. Ca(I) has been observed; see Krieck, Sven; Görls, Helmar; Westerhausen, Matthias (2010). "Mechanistic
Elucidation of the Formation of the Inverse Ca(I) Sandwich Complex [(thf)3Ca(μ-C6H3-1,3,5-Ph3)Ca(thf)3] and
Stability of Aryl-Substituted Phenylcalcium Complexes". Journal of the American Chemical Society. 132 (35):
12492–501. doi:10.1021/ja105534w (https://doi.org/10.1021%2Fja105534w). PMID 20718434 (https://pubmed.ncbi.
nlm.nih.gov/20718434).
37. Sc(0) has been observed; see F. Geoffrey N. Cloke; Karl Khan & Robin N. Perutz (1991). "η-Arene complexes of
scandium(0) and scandium(II)". J. Chem. Soc., Chem. Commun. (19): 1372–1373. doi:10.1039/C39910001372 (http
s://doi.org/10.1039%2FC39910001372).
38. Sc(I) has been observed; see Polly L. Arnold; F. Geoffrey; N. Cloke; Peter B. Hitchcock & John F. Nixon (1996). "The
First Example of a Formal Scandium(I) Complex: Synthesis and Molecular Structure of a 22-Electron Scandium
Triple Decker Incorporating the Novel 1,3,5-Triphosphabenzene Ring". J. Am. Chem. Soc. 118 (32): 7630–7631.
doi:10.1021/ja961253o (https://doi.org/10.1021%2Fja961253o).
39. Sc(II) has been observed; see Woen, David H.; Chen, Guo P.; Ziller, Joseph W.; Boyle, Timothy J.; Furche, Filipp;
Evans, William J. (January 2017). "Solution Synthesis, Structure, and CO Reduction Reactivity of a Scandium(II)
Complex". Angewandte Chemie International Edition. 56 (8): 2050–2053. doi:10.1002/anie.201611758 (https://doi.o
rg/10.1002%2Fanie.201611758). PMID 28097771 (https://pubmed.ncbi.nlm.nih.gov/28097771).
40. Ti(I) has been observed in [Ti(η6-1,3,5-C6H3iPr3)2][BAr4] (Ar = C6H5, p-C6H4F, 3,5-C6H3(CF3)2); see Calderazzo,
Fausto; Ferri, Isabella; Pampaloni, Guido; Englert, Ulli; Green, Malcolm L. H. (1997). "Synthesis of [Ti(η6-1,3,5-
C6H3iPr3)2][BAr4] (Ar = C6H5, p-C6H4F, 3,5-C6H3(CF3)2), the First Titanium(I) Derivatives". Organometallics. 16
(14): 3100–3101. doi:10.1021/om970155o (https://doi.org/10.1021%2Fom970155o).
41. Ti(−2), V(−3), Cr(−4), Co(−3), Zr(−2), Nb(−3), Mo(−4), Ru(−2), Rh(−3), Hf(−2), Ta(−3), and W(−4) occur in anionic
binary metal carbonyls; see [1] (http://d-nb.info/956310893/34), p. 4 (in German); [2] (https://books.google.com/book
s?id=dP4LTfaPzAMC&pg=PA97&dq=%22%5BCr%28CO%294%5D4-%22&hl=en&sa=X&ei=ncroVMihLYO8ygO22
YLYAQ&ved=0CBsQ6AEwAA#v=onepage&q&f=false), pp. 97–100; [3] (https://books.google.com/books?id=n5r-NF
T46TkC&printsec=frontcover&hl=ru#v=onepage&q&f=false), p. 239
42. Ti(−1) has been reported in [Ti(bipy)3]− , but was later shown to be Ti(+3); see Bowman, A. C.; England, J.; Sprouls,
S.; Weihemüller, T.; Wieghardt, K. (2013). "Electronic structures of homoleptic [tris(2,2'-bipyridine)M]n complexes of
the early transition metals (M = Sc, Y, Ti, Zr, Hf, V, Nb, Ta; n = 1+, 0, 1-, 2-, 3-): an experimental and density functional
theoretical study". Inorganic Chemistry. 52 (4): 2242–56. doi:10.1021/ic302799s (https://doi.org/10.1021%2Fic3027
99s). PMID 23387926 (https://pubmed.ncbi.nlm.nih.gov/23387926). However, Ti(−1) occurs in [Ti(η-C6H6]− and
[Ti(η-C6H5CH3)]− , see Bandy, J. A.; Berry, A.; Green, M. L. H.; Perutz, R. N.; Prout, K.; Verpeautz, J.-N. (1984).
"Synthesis of anionic sandwich compounds: [Ti(η-C6H5R)2]– and the crystal structure of [K(18-crown-6)(µ-H)Mo(η-
C5H5)2]". Inorganic Chemistry. 52 (4): 729–731. doi:10.1039/C39840000729 (https://doi.org/10.1039%2FC3984000
0729).
43. Jilek, Robert E.; Tripepi, Giovanna; Urnezius, Eugenijus; Brennessel, William W.; Young, Victor G., Jr.; Ellis, John E.
(2007). "Zerovalent titanium–sulfur complexes. Novel dithiocarbamato derivatives of Ti(CO)6: [Ti(CO)4(S2CNR2)]−".
Chem. Commun. (25): 2639–2641. doi:10.1039/B700808B (https://doi.org/10.1039%2FB700808B). PMID 17579764
(https://pubmed.ncbi.nlm.nih.gov/17579764).
44. Fe(VII) has been observed in [FeO4]− ; see Lu, Jun-Bo; Jian, Jiwen; Huang, Wei; Lin, Hailu; Zhou, Mingfei (2016).
"Experimental and theoretical identification of the Fe(VII) oxidation state in FeO4−". Physical Chemistry Chemical
Physics. 18 (45): 31125–31131. Bibcode:2016PCCP...1831125L (https://ui.adsabs.harvard.edu/abs/2016PCCP...18
31125L). doi:10.1039/C6CP06753K (https://doi.org/10.1039%2FC6CP06753K). PMID 27812577 (https://pubmed.nc
bi.nlm.nih.gov/27812577).
45. Fe(VIII) has been reported; see Yurii D. Perfiliev; Virender K. Sharma (2008). "Higher Oxidation States of Iron in
Solid State: Synthesis and Their Mössbauer Characterization – Ferrates – ACS Symposium Series (ACS
Publications)". Platinum Metals Review. 48 (4): 157–158. doi:10.1595/147106704X10801 (https://doi.org/10.1595%
2F147106704X10801). However, its existence has been disputed.
46. Fe(−4), Ru(−4), and Os(−4) have been observed in metal-rich compounds containing octahedral complexes
[MIn6−x Snx ]; Pt(−3) (as a dimeric anion [Pt–Pt]6− ), Cu(−2), Zn(−2), Ag(−2), Cd(−2), Au(−2), and Hg(−2) have been
observed (as dimeric and monomeric anions; dimeric ions were initially reported to be [T–T]2− for Zn, Cd, Hg, but
later shown to be [T–T]4− for all these elements) in La2Pt2In, La2Cu2In, Ca5Au3, Ca5Ag3, Ca5Hg3, Sr5Cd3,
Ca5Zn3(structure (AE2+ )5(T–T)4− T2− ⋅4e− ), Yb3Ag2, Ca5Au4, and Ca3Hg2; Au(–3) has been observed in ScAuSn
and in other 18-electron half-Heusler compounds. See Changhoon Lee; Myung-Hwan Whangbo (2008). "Late
transition metal anions acting as p-metal elements". Solid State Sciences. 10 (4): 444–449.
Bibcode:2008SSSci..10..444K (https://ui.adsabs.harvard.edu/abs/2008SSSci..10..444K).
doi:10.1016/j.solidstatesciences.2007.12.001 (https://doi.org/10.1016%2Fj.solidstatesciences.2007.12.001). and
Changhoon Lee; Myung-Hwan Whangbo; Jürgen Köhler (2010). "Analysis of Electronic Structures and Chemical
Bonding of Metal-rich Compounds. 2. Presence of Dimer (T–T)4– and Isolated T2– Anions in the Polar Intermetallic
Cr5B3-Type Compounds AE5T3 (AE = Ca, Sr; T = Au, Ag, Hg, Cd, Zn)". Zeitschrift für Anorganische und
Allgemeine Chemie. 636 (1): 36–40. doi:10.1002/zaac.200900421 (https://doi.org/10.1002%2Fzaac.200900421).
47. Ni(−2) has been observed in Li2[Ni(1,5-COD)2], see Jonas, Klaus (1975). "Dilithium-Nickel-Olefin Complexes.
Novel Bimetal Complexes Containing a Transition Metal and a Main Group Metal". Angew. Chem. Int. Ed. 14 (11):
752–753. doi:10.1002/anie.197507521 (https://doi.org/10.1002%2Fanie.197507521). and Ellis, John E. (2006).
"Adventures with Substances Containing Metals in Negative Oxidation States". Inorganic Chemistry. 45 (8): 3167–
86. doi:10.1021/ic052110i (https://doi.org/10.1021%2Fic052110i). PMID 16602773 (https://pubmed.ncbi.nlm.nih.go
v/16602773).
48. Cu(0) has been observed in Cu(tris[2-(diisopropylphosphino)- phenyl]borane), see Moret, Marc-Etienne; Zhang,
Limei; Peters, Jonas C. (2013). "A Polar Copper–Boron One-Electron σ-Bond". J. Am. Chem. Soc. 135 (10): 3792–
3795. doi:10.1021/ja4006578 (https://doi.org/10.1021%2Fja4006578). PMID 23418750 (https://pubmed.ncbi.nlm.ni
h.gov/23418750).
49. Zn(I) has been observed in decamethyldizincocene (Zn2(η5–C5Me5)2); see Resa, I.; Carmona, E.; Gutierrez-Puebla,
E.; Monge, A. (2004). "Decamethyldizincocene, a Stable Compound of Zn(I) with a Zn-Zn Bond". Science. 305
(5687): 1136–8. Bibcode:2004Sci...305.1136R (https://ui.adsabs.harvard.edu/abs/2004Sci...305.1136R).
doi:10.1126/science.1101356 (https://doi.org/10.1126%2Fscience.1101356). PMID 15326350 (https://pubmed.ncbi.
nlm.nih.gov/15326350).
50. Ga(−2), Ga(−4), and Ga(−5) have been observed in the magnesium gallides MgGa, Mg2Ga, and Mg5Ga2,
respectively; see Patrick Hofmann. "Colture. Ein Programm zur interaktiven Visualisierung von Festkörperstrukturen
sowie Synthese, Struktur und Eigenschaften von binären und ternären Alkali- und Erdalkalimetallgalliden" (http://w
ww.uni-kassel.de/upress/online/frei/978-3-7281-2597-2.volltext.frei.pdf) (PDF) (in German). p. 72.
51. Ga(−3) has been observed in LaGa, see Dürr, Ines; Bauer, Britta; Röhr, Caroline (2011). "Lanthan-Triel/Tetrel-ide
La(Al,Ga)x(Si,Ge)1-x. Experimentelle und theoretische Studien zur Stabilität intermetallischer 1:1-Phasen" (http://w
ww.znaturforsch.com/s66b/s66b1107.pdf) (PDF). Z. Naturforsch. (in German). 66b: 1107–1121.
52. Ge(−1), Ge(−2), and Ge(−3) have been observed in germanides; see Holleman, Arnold F.; Wiberg, Egon; Wiberg,
Nils (1995). "Germanium". Lehrbuch der Anorganischen Chemie (in German) (101 ed.). Walter de Gruyter. pp. 953–
959. ISBN 978-3-11-012641-9.
53. As(0) has been observed; see Abraham, Mariham Y.; Wang, Yuzhong; Xie, Yaoming; Wei, Pingrong; Shaefer III,
Henry F.; Schleyer, P. von R.; Robinson, Gregory H. (2010). "Carbene Stabilization of Diarsenic: From Hypervalency
to Allotropy". Chemistry: a European Journal. 16 (2): 432–5. doi:10.1002/chem.200902840 (https://doi.org/10.1002%
2Fchem.200902840).
54. As(I) has been observed in arsenic(I) iodide (AsI); see Ellis, Bobby D.; MacDonald, Charles L. B. (2004). "Stabilized
Arsenic(I) Iodide: A Ready Source of Arsenic Iodide Fragments and a Useful Reagent for the Generation of
Clusters". Inorganic Chemistry. 43 (19): 5981–6. doi:10.1021/ic049281s (https://doi.org/10.1021%2Fic049281s).
PMID 15360247 (https://pubmed.ncbi.nlm.nih.gov/15360247).
55. As(IV) has been observed in arsenic(IV) hydroxide (As(OH)4) and HAsO-; see Kläning, Ulrik K.; Bielski, Benon H. J.;
Sehested, K. (1989). "Arsenic(IV). A pulse-radiolysis study". Inorganic Chemistry. 28 (14): 2717–24.
doi:10.1021/ic00313a007 (https://doi.org/10.1021%2Fic00313a007).
56. Se(−1) has been observed in diselenides(2−) (Se22− ).
57. Se(I) has been observed in selenium(I) chloride (Se2Cl2); see "Selenium: Selenium(I) chloride compound data" (htt
p://www.webelements.com/webelements/compounds/text/Se/Cl2Se2-10025680.html). WebElements.com.
Retrieved 2007-12-10.
58. Se(III) has been observed in Se2NBr3; see Lau, Carsten; Neumüller, Bernhard; Vyboishchikov, Sergei F.; Frenking,
Gernot; Dehnicke, Kurt; Hiller, Wolfgang; Herker, Martin (1996). "Se2NBr3, Se2NCl5, Se2NCl−6: New Nitride
Halides of Selenium(III) and Selenium(IV)". Chemistry: A European Journal. 2 (11): 1393–1396.
doi:10.1002/chem.19960021108 (https://doi.org/10.1002%2Fchem.19960021108).
59. Se(V) has been observed in SeO2- and HSeO2-; see Kläning, Ulrik K.; Sehested, K. (1986). "Selenium(V). A pulse
radiolysis study" (http://orbit.dtu.dk/en/publications/seleniumv-a-pulse-radiolysis-study(01d1260d-45a3-4a09-8316-
e6f6b4993561).html). Inorganic Chemistry. 90 (21): 5460–4. doi:10.1021/j100412a112 (https://doi.org/10.1021%2Fj
100412a112).
60. Sr(I) has been observed in strontium monofluoride (SrF); see P. Colarusso; Guo, B.; Zhang, K.-Q.; Bernath, P.F.; et
al. (1996). "High-Resolution Infrared Emission Spectrum of Strontium Monofluoride" (https://web.archive.org/web/20
120308063843/http://bernath.uwaterloo.ca/media/149.pdf) (PDF). Journal of Molecular Spectroscopy. 175 (1): 158–
171. Bibcode:1996JMoSp.175..158C (https://ui.adsabs.harvard.edu/abs/1996JMoSp.175..158C).
doi:10.1006/jmsp.1996.0019 (https://doi.org/10.1006%2Fjmsp.1996.0019). Archived from the original (http://bernath.
uwaterloo.ca/media/149.pdf) (PDF) on 2012-03-08.
61. Yttrium and all lanthanides except Ce, Pm, Eu, Tm, Yb have been observed in the oxidation state 0 in bis(1,3,5-tri-t-
butylbenzene) complexes, see Cloke, F. Geoffrey N. (1993). "Zero Oxidation State Compounds of Scandium,
Yttrium, and the Lanthanides". Chem. Soc. Rev. 22: 17–24. doi:10.1039/CS9932200017 (https://doi.org/10.1039%2
FCS9932200017).
62. Y(I) has been observed in yttrium(I) bromide (YBr); see "Yttrium: yttrium(I) bromide compound data" (https://web.arch
ive.org/web/20110723233118/http://www.openmopac.net/data_normal/yttrium%28i%29%20bromide_jmol.html).
OpenMOPAC.net. Archived from the original (http://www.openmopac.net/data_normal/yttrium(i)%20bromide_jmol.ht
ml) on 2011-07-23. Retrieved 2007-12-10.
63. Y(II) has been observed in [(18-crown-6)K][(C5H4SiMe3)3Y]; see MacDonald, M. R.; Ziller, J. W.; Evans, W. J. (2011).
"Synthesis of a Crystalline Molecular Complex of Y2+, [(18-crown-6)K][(C5H4SiMe3)3Y]". J. Am. Chem. Soc. 133
(40): 15914–17. doi:10.1021/ja207151y (https://doi.org/10.1021%2Fja207151y). PMID 21919538 (https://pubmed.nc
bi.nlm.nih.gov/21919538).
64. Zr(−1) has been reported in [Zr(bipy)3]− (see Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the
Elements (2nd ed.). Butterworth-Heinemann. p. 960. ISBN 978-0-08-037941-8. and Holleman, Arnold F.; Wiberg,
Egon; Wiberg, Nils (1995). "Zirconium". Lehrbuch der Anorganischen Chemie (in German) (101 ed.). Walter de
Gruyter. p. 1413. ISBN 978-3-11-012641-9.), but was later shown to be Zr(+4); see Bowman, A. C.; England, J.;
Sprouls, S.; Weihemüller, T.; Wieghardt, K. (2013). "Electronic structures of homoleptic [tris(2,2'-bipyridine)M]n
complexes of the early transition metals (M = Sc, Y, Ti, Zr, Hf, V, Nb, Ta; n = 1+, 0, 1-, 2-, 3-): an experimental and
density functional theoretical study". Inorganic Chemistry. 52 (4): 2242–56. doi:10.1021/ic302799s (https://doi.org/1
0.1021%2Fic302799s). PMID 23387926 (https://pubmed.ncbi.nlm.nih.gov/23387926).
65. Nb(I) and Ta(I) occur in CpNb(CO)4 and CpTa(CO)4, see Holleman, Arnold F.; Wiberg, Egon; Wiberg, Nils (1995).
"Tantal". Lehrbuch der Anorganischen Chemie (in German) (101 ed.). Walter de Gruyter. p. 1430. ISBN 978-3-11-
012641-9. and King, R. Bruce (1969). Transition-Metal Organometallic Chemistry: An Introduction. Academic Press.
p. 11. ISBN 978-0-32-315996-8.
66. George, G.N.; Klein, S.I.; Nixon, J.F. (1984). "Electron paramagnetic resonance spectroscopic studies on the zero-
valent rhodium complex [Rh(P(OPri)3)4] at X-and Q-band frequencies". Chemical Physics Letters. 108 (6): 627–
630. Bibcode:1984CPL...108..627G (https://ui.adsabs.harvard.edu/abs/1984CPL...108..627G). doi:10.1016/0009-
2614(84)85069-1 (https://doi.org/10.1016%2F0009-2614%2884%2985069-1).
67. Pd(I) has been observed; see Crabtree, R. H. (2002). "CHEMISTRY: A New Oxidation State for Pd?". Science. 295
(5553): 288–289. doi:10.1126/science.1067921 (https://doi.org/10.1126%2Fscience.1067921). PMID 11786632 (htt
ps://pubmed.ncbi.nlm.nih.gov/11786632).
68. Pd(III) has been observed; see Powers, D. C.; Ritter, T. (2011). Palladium(III) in Synthesis and Catalysis (https://web.
archive.org/web/20130612065217/http://www.chem.harvard.edu/groups/ritter/pdf/2011-129t.pdf) (PDF). Top.
Organomet. Chem. Topics in Organometallic Chemistry. 35. pp. 129–156. Bibcode:2011hoso.book..129P (https://ui.
adsabs.harvard.edu/abs/2011hoso.book..129P). doi:10.1007/978-3-642-17429-2_6 (https://doi.org/10.1007%2F978
-3-642-17429-2_6). ISBN 978-3-642-17428-5. PMC 3066514 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC30665
14). PMID 21461129 (https://pubmed.ncbi.nlm.nih.gov/21461129). Archived from the original on June 12, 2013.
69. The Ag− ion has been observed in metal ammonia solutions: see Tran, N. E.; Lagowski, J. J. (2001). "Metal
Ammonia Solutions: Solutions Containing Argentide Ions". Inorganic Chemistry. 40 (5): 1067–68.
doi:10.1021/ic000333x (https://doi.org/10.1021%2Fic000333x).
70. Cd(I) has been observed in cadmium(I) tetrachloroaluminate (Cd2(AlCl4)2); see Holleman, Arnold F.; Wiberg, Egon;
Wiberg, Nils (1985). "Cadmium". Lehrbuch der Anorganischen Chemie (in German) (91–100 ed.). Walter de Gruyter.
pp. 1056–1057. ISBN 978-3-11-007511-3.
71. In(–5) has been observed in La3InGe, see Guloy, A. M.; Corbett, J. D. (1996). "Synthesis, Structure, and Bonding of
Two Lanthanum Indium Germanides with Novel Structures and Properties". Inorganic Chemistry. 35 (9): 2616–22.
doi:10.1021/ic951378e (https://doi.org/10.1021%2Fic951378e).
72. In(−2) has been observed in Na2In, see [4] (https://books.google.com/books?id=v-04Kn758yIC&pg=PA69&lpg=PA6
9&dq=zintl+anions+Na2In&source=bl&ots=aXLYIpkfYq&sig=Mqh8WdnvGOt2J2OPVLNqn79YVyk&hl=ru&sa=X&ei
=XNDkVNeSJeXOyQOb8oBA&ved=0CBsQ6AEwADgK#v=onepage&q&f=false), p. 69.
73. Sn(−3) has been observed in [Sn2]6− , e.g. in (Ba2)4+ (Mg4)8+ Sn4− (Sn2)6− Sn2− (with square (Sn2− )n sheets), see
Papoian, Garegin A.; Hoffmann, Roald (2000). "Hypervalent Bonding in One, Two, and Three Dimensions:
Extending the Zintl–Klemm Concept to Nonclassical Electron-Rich Networks" (https://www.researchgate.net/publica
tion/12379848). Angew. Chem. Int. Ed. 2000 (39): 2408–2448. doi:10.1002/1521-3773(20000717)39:14<2408::aid-
anie2408>3.0.co;2-u (https://doi.org/10.1002%2F1521-3773%2820000717%2939%3A14%3C2408%3A%3Aaid-ani
e2408%3E3.0.co%3B2-u). Retrieved 2015-02-23.
74. Sn(I) and Sn(III) have been observed in organotin compounds
75. Sb(−2) has been observed in [Sb2]4− , e.g. in RbBa4[Sb2][Sb][O], see Boss, Michael; Petri, Denis; Pickhard, Frank;
Zönnchen, Peter; Röhr, Caroline (2005). "Neue Barium-Antimonid-Oxide mit den Zintl-Ionen [Sb]3−, [Sb2]4− und
1 [Sb ]n− / New Barium Antimonide Oxides containing Zintl Ions [Sb]3−, [Sb ]4− and 1 [Sb ]n−". Zeitschrift für
∞ n 2 ∞ n
Anorganische und Allgemeine Chemie (in German). 631 (6–7): 1181–1190. doi:10.1002/zaac.200400546 (https://do
i.org/10.1002%2Fzaac.200400546).
76. Sb(0) has been observed, see Anastas Sidiropoulos. "Studies of N-heterocyclic Carbene (NHC) Complexes of the
Main Group Elements" (https://pdfs.semanticscholar.org/8817/39f9dfc007d7f77dd7baa63fe12e6079f8ef.pdf) (PDF).
p. 39.
77. Sb(I) and Sb(II) have been observed in organoantimony compounds; for Sb(I), see Šimon, Petr; de Proft, Frank;
Jambor, Roman; Růžička, Aleš; Dostál, Libor (2010). "Monomeric Organoantimony(I) and Organobismuth(I)
Compounds Stabilized by an NCN Chelating Ligand: Syntheses and Structures". Angewandte Chemie International
Edition. 49 (32): 5468–5471. doi:10.1002/anie.201002209 (https://doi.org/10.1002%2Fanie.201002209).
PMID 20602393 (https://pubmed.ncbi.nlm.nih.gov/20602393).
2−
78. Sb(IV) has been observed in [SbCl] , see Nobuyoshi Shinohara; Masaaki Ohsima (2000). "Production of Sb(IV)
Chloro Complex by Flash Photolysis of the Corresponding Sb(III) and Sb(V) Complexes in CH3CN and CHCl3".
Bulletin of the Chemical Society of Japan. 73 (7): 1599–1604. doi:10.1246/bcsj.73.1599 (https://doi.org/10.1246%2F
bcsj.73.1599).
79. Te(I) has been observed in tellurium iodide (TeI), see "Tellurium: tellurium iodide" (http://www.webelements.com/com
pounds/tellurium/tellurium_iodide.html). WebElements.com. Retrieved 2015-02-23.
80. Te(III) has been observed in [Te(N(SiMe3)2)2]+ , see Heinze, Thorsten; Roesky, Herbert W.; Pauer, Frank; Stalke,
Dietmar; Sheldrick, George M. (1991). "Synthesis and Structure of the First Tellurium(III) Radical Cation" (https://ww
w.researchgate.net/publication/237225046). Angewandte Chemie International Edition. 30 (12): 1678.
doi:10.1002/anie.199116771 (https://doi.org/10.1002%2Fanie.199116771). Retrieved 2015-02-23.
81. Te(V) is mentioned by Greenwood and Earnshaw, but they do not give any example of a Te(V) compound. What was
long thought to be ditellurium decafluoride (Te2F10) is actually bis(pentafluorotelluryl) oxide, F5TeOTeF5: see
Watkins, P. M. (1974). "Ditellurium decafluoride - A Continuing Myth". Journal of Chemical Education. 51 (9): 520–
521. Bibcode:1974JChEd..51..520W (https://ui.adsabs.harvard.edu/abs/1974JChEd..51..520W).
doi:10.1021/ed051p520 (https://doi.org/10.1021%2Fed051p520). However, Te(V) has been observed in HTeO-,
TeO-, HTeO2-, and TeO3-; see Kläning, Ulrik K.; Sehested, K. (2001). "Tellurium(V). A Pulse Radiolysis Study" (htt
p://orbit.dtu.dk/en/publications/tellurium-5-a-pulse-radiolysis-study(58c2417f-34c0-436d-8a46-211f3d752423).html).
The Journal of Physical Chemistry A. 105 (27): 6637–45. Bibcode:2001JPCA..105.6637K (https://ui.adsabs.harvard.
edu/abs/2001JPCA..105.6637K). doi:10.1021/jp010577i (https://doi.org/10.1021%2Fjp010577i).
82. I(IV) has been observed in iodine dioxide (IO2); see Pauling, Linus (1988). "Oxygen Compounds of Nonmetallic
Elements". General Chemistry (3rd ed.). Dover Publications, Inc. p. 259. ISBN 978-0-486-65622-9.
83. I(VI) has been observed in IO3, IO42− , H5IO6− , H2IO52− , H4IO62− , and HIO53− ; see Kläning, Ulrik K.; Sehested, Knud;
Wolff, Thomas (1981). "Laser flash photolysis and pulse radiolysis of iodate and periodate in aqueous solution.
Properties of iodine(VI)". J. Chem. Soc., Faraday Trans. 1. 77 (7): 1707–18. doi:10.1039/F19817701707 (https://doi.
org/10.1039%2FF19817701707).
84. Xe compounds: see Xenon
85. Xe(0) has been observed in tetraxenonogold(II) (AuXe42+ ).
86. Xe(I) has been reported in xenon hexafluoroplatinate and xenon hexafluororhodate (see Pauling, Linus (1988).
General Chemistry (3rd ed.). Dover Publications, Inc. p. 250. ISBN 978-0-486-65622-9.), however these compounds
were later found to contain Xe(II).
87. Ba(I) has been observed in barium monofluoride (BaF); see P. Colarusso; Guo, B.; Zhang, K.-Q.; Bernath, P.F.; et al.
(1995). "High-Resolution Fourier Transform Infrared Emission Spectrum of Barium Monofluoride" (https://web.archiv
e.org/web/20050310180822/http://bernath.uwaterloo.ca/media/126.pdf) (PDF). Journal of Molecular Spectroscopy.
170: 59. Bibcode:1996JMoSp.175..158C (https://ui.adsabs.harvard.edu/abs/1996JMoSp.175..158C).
doi:10.1006/jmsp.1996.0019 (https://doi.org/10.1006%2Fjmsp.1996.0019). Archived from the original (http://bernath.
uwaterloo.ca/media/126.pdf) (PDF) on 2005-03-10.
88. La(I) has been observed in lanthanum monohydride (LaH); see Ram, R. S.; Bernath, P. F. (1996). "Fourier Transform
Emission Spectroscopy of New Infrared Systems of LaH and LaD" (https://web.archive.org/web/20050310194449/ht
tp://bernath.uwaterloo.ca/media/143.pdf) (PDF). Journal of Molecular Spectroscopy. 104 (17): 6444.
Bibcode:1996JChPh.104.6444R (https://ui.adsabs.harvard.edu/abs/1996JChPh.104.6444R). doi:10.1063/1.471365
(https://doi.org/10.1063%2F1.471365). Archived from the original (http://bernath.uwaterloo.ca/media/143.pdf) (PDF)
on 2005-03-10.
89. Pr(I) has been observed in [PrB4]− ; see Chen, Xin; Chen, Teng-Teng; Li, Wang-Lu; Lu, Jun-Bo; Zhao, Li-Juan; Jian,
Tian; Hu, Han-Shi; Wang, Lai-Sheng; Li, Jun (2018-12-13). "Lanthanides with Unusually Low Oxidation States in
the PrB3– and PrB4– Boride Clusters". Inorganic Chemistry. 58 (1): 411–418. doi:10.1021/acs.inorgchem.8b02572
(https://doi.org/10.1021%2Facs.inorgchem.8b02572). PMID 30543295 (https://pubmed.ncbi.nlm.nih.gov/30543295).
90. Pr(V) has been observed in [PrO2]+ ; see Zhang, Qingnan; Hu, Shu-Xian; Qu, Hui; Su, Jing; Wang, Guanjun; Lu, Jun-
Bo; Chen, Mohua; Zhou, Mingfei; Li, Jun (2016-06-06). "Pentavalent Lanthanide Compounds: Formation and
Characterization of Praseodymium(V) Oxides". Angewandte Chemie International Edition. 55 (24): 6896–6900.
doi:10.1002/anie.201602196 (https://doi.org/10.1002%2Fanie.201602196). ISSN 1521-3773 (https://www.worldcat.
org/issn/1521-3773). PMID 27100273 (https://pubmed.ncbi.nlm.nih.gov/27100273).
91. Hu, Shu-Xian; Jian, Jiwen; Su, Jing; Wu, Xuan; Li, Jun; Zhou, Mingfei (2017). "Pentavalent lanthanide nitride-
oxides: NPrO and NPrO− complexes with N≡Pr triple bonds" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC54349
15). Chemical Science. 8 (5): 4035–4043. doi:10.1039/C7SC00710H (https://doi.org/10.1039%2FC7SC00710H).
ISSN 2041-6520 (https://www.worldcat.org/issn/2041-6520). PMC 5434915 (https://www.ncbi.nlm.nih.gov/pmc/articl
es/PMC5434915). PMID 28580119 (https://pubmed.ncbi.nlm.nih.gov/28580119).
92. Nd(IV) has been observed in unstable solid state compounds; see Holleman, Arnold Frederik; Wiberg, Egon (2001),
Wiberg, Nils (ed.), Inorganic Chemistry, translated by Eagleson, Mary; Brewer, William, San Diego/Berlin: Academic
Press/De Gruyter, ISBN 0-12-352651-5
93. All the lanthanides (La–Lu) in the +2 oxidation state have been observed (except La, Gd, Lu) in dilute, solid
solutions of dihalides of these elements in alkaline earth dihalides (see Holleman, Arnold Frederik; Wiberg, Egon
(2001), Wiberg, Nils (ed.), Inorganic Chemistry, translated by Eagleson, Mary; Brewer, William, San Diego/Berlin:
Academic Press/De Gruyter, ISBN 0-12-352651-5) and (except Pm) in organometallic molecular complexes, see
Lanthanides Topple Assumptions (http://cen.acs.org/articles/91/i24/Lanthanides-Topple-Assumptions.html) and
Meyer, G. (2014). "All the Lanthanides Do It and Even Uranium Does Oxidation State +2". Angewandte Chemie
International Edition. 53 (14): 3550–51. doi:10.1002/anie.201311325 (https://doi.org/10.1002%2Fanie.201311325).
PMID 24616202 (https://pubmed.ncbi.nlm.nih.gov/24616202).. Additionally, all the lanthanides (La–Lu) form
dihydrides (LnH2), dicarbides (LnC2), monosulfides (LnS), monoselenides (LnSe), and monotellurides (LnTe), but
for most elements these compounds have Ln3+ ions with electrons delocalized into conduction bands, e. g. Ln3+
(H− )2(e− ).
94. Dy(IV) has been observed in unstable solid state compounds; see Holleman, Arnold Frederik; Wiberg, Egon (2001),
Wiberg, Nils (ed.), Inorganic Chemistry, translated by Eagleson, Mary; Brewer, William, San Diego/Berlin: Academic
Press/De Gruyter, ISBN 0-12-352651-5
95. Hf(I) has been observed in hafnium monobromide (HfBr), see Marek, G.S.; Troyanov, S.I.; Tsirel'nikov, V.I. (1979).
"Кристаллическое строение и термодинамические характеристики монобромидов циркония и гафния /
Crystal structure and thermodynamic characteristics of monobromides of zirconium and hafnium" (https://inis.iaea.or
g/search/search.aspx?orig_q=RN:11520917). Журнал неорганической химии / Russian Journal of Inorganic
Chemistry (in Russian). 24 (4): 890–893.
96. Os(−1) has been observed in Na[Os(CO)13]; see Krause, J.; Siriwardane, Upali; Salupo, Terese A.; Wermer, Joseph
R.; Knoeppel, David W.; Shore, Sheldon G. (1993). "Preparation of [Os3(CO)11]2− and its reactions with
Os3(CO)12; structures of [Et4N] [HOs3(CO)11] and H2OsS4(CO)". Journal of Organometallic Chemistry. 454: 263–
271. doi:10.1016/0022-328X(93)83250-Y (https://doi.org/10.1016%2F0022-328X%2893%2983250-Y). and Carter,
Willie J.; Kelland, John W.; Okrasinski, Stanley J.; Warner, Keith E.; Norton, Jack R. (1982). "Mononuclear hydrido
alkyl carbonyl complexes of osmium and their polynuclear derivatives". Inorganic Chemistry. 21 (11): 3955–3960.
doi:10.1021/ic00141a019 (https://doi.org/10.1021%2Fic00141a019).
97. Ir(−3) has been observed in Ir(CO)33− ; see Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the
Elements (2nd ed.). Butterworth-Heinemann. p. 1117. ISBN 978-0-08-037941-8.
98. Ir(VII) has been observed in [(η2-O2)IrO2]+ ; see C&EN: Iridium dressed to the nines (http://2014.cenmag.org/iridium-d
ressed-to-the-nines/).
99. Ir(VIII) has been observed in iridium tetroxide (IrO4); see Gong, Yu; Zhou, Mingfei; Kaupp, Martin; Riedel, Sebastian
(2009). "Formation and Characterization of the Iridium Tetroxide Molecule with Iridium in the Oxidation State +VIII".
Angewandte Chemie International Edition. 48 (42): 7879–7883. doi:10.1002/anie.200902733 (https://doi.org/10.100
2%2Fanie.200902733). PMID 19593837 (https://pubmed.ncbi.nlm.nih.gov/19593837).
00. Ir(IX) has been observed in IrO+4 ; see Wang, Guanjun; Zhou, Mingfei; Goettel, James T.; Schrobilgen, Gary G.; Su,
Jing; Li, Jun; Schlöder, Tobias; Riedel, Sebastian (21 August 2014). "Identification of an iridium-containing
compound with a formal oxidation state of IX". Nature. 514 (7523): 475–477. Bibcode:2014Natur.514..475W (https://
ui.adsabs.harvard.edu/abs/2014Natur.514..475W). doi:10.1038/nature13795 (https://doi.org/10.1038%2Fnature137
95). PMID 25341786 (https://pubmed.ncbi.nlm.nih.gov/25341786).
01. Pt(−1) and Pt(−2) have been observed in the barium platinides Ba2Pt and BaPt, respectively: see Karpov, Andrey;
Konuma, Mitsuharu; Jansen, Martin (2006). "An experimental proof for negative oxidation states of platinum: ESCA-
measurements on barium platinides". Chemical Communications (8): 838–840. doi:10.1039/b514631c (https://doi.or
g/10.1039%2Fb514631c). PMID 16479284 (https://pubmed.ncbi.nlm.nih.gov/16479284).
02. Pt(I) and Pt(III) have been observed in bimetallic and polymetallic species; see Kauffman, George B.; Thurner,
Joseph J.; Zatko, David A. (1967). Ammonium Hexachloroplatinate(IV). Inorganic Syntheses. 9. pp. 182–185.
doi:10.1002/9780470132401.ch51 (https://doi.org/10.1002%2F9780470132401.ch51). ISBN 978-0-470-13240-1.
03. Au(0) has been observed, see Mézaille, Nicolas; Avarvari, Narcis; Maigrot, Nicole; Ricard, Louis; Mathey, François;
Le Floch, Pascal; Cataldo, Laurent; Berclaz, Théo; Geoffroy, Michel (1999). "Gold(I) and Gold(0) Complexes of
Phosphinine‐Based Macrocycles". Angewandte Chemie International Edition (21): 3194–3197.
doi:10.1002/(SICI)1521-3773(19991102)38:21<3194::AID-ANIE3194>3.0.CO;2-O (https://doi.org/10.1002%2F%28
SICI%291521-3773%2819991102%2938%3A21%3C3194%3A%3AAID-ANIE3194%3E3.0.CO%3B2-O).
04. Hg(IV) has been reported in mercury(IV) fluoride (HgF4); see Xuefang Wang; Lester Andrews; Sebastian Riedel;
Martin Kaupp (2007). "Mercury Is a Transition Metal: The First Experimental Evidence for HgF4". Angew. Chem. Int.
Ed. 46 (44): 8371–8375. doi:10.1002/anie.200703710 (https://doi.org/10.1002%2Fanie.200703710).
PMID 17899620 (https://pubmed.ncbi.nlm.nih.gov/17899620). However, it could not be confirmed by later
experiments; see Is mercury a transition metal? (http://www2.hull.ac.uk/science/chemistry/research/inorganicmateria
ls/mercurytransitionmaterial.aspx) Archived (https://web.archive.org/web/20161012232329/http://www2.hull.ac.uk/sc
ience/chemistry/research/inorganicmaterials/mercurytransitionmaterial.aspx) 2016-10-12 at the Wayback Machine
05. Tl(−5) has been observed in Na23K9Tl15.3, see Dong, Z.-C.; Corbett, J. D. (1996). "Na23K9Tl15.3: An Unusual Zintl
Compound Containing Apparent Tl57−, Tl48−, Tl37−, and Tl5− Anions". Inorganic Chemistry. 35 (11): 3107–12.
doi:10.1021/ic960014z (https://doi.org/10.1021%2Fic960014z).
06. Tl(−1) has been observed in caesium thallide (CsTl); see King, R. B.; Schleyer, R. (2004). "Theory and concepts in
main-group cluster chemistry". In Driess, M.; Nöth, H. (eds.). Molecular clusters of the main group elements. Wiley-
VCH, Chichester. p. 19. ISBN 978-3-527-61437-0.
07. Tl(+2) has been observed in tetrakis(hypersilyl)dithallium ([(Me3Si)Si]2Tl—Tl[Si(SiMe3)]2), see Sonja Henkel; Dr.
Karl Wilhelm Klinkhammer; Dr. Wolfgang Schwarz (1994). "Tetrakis(hypersilyl)dithallium(Tl—Tl): A Divalent
Thallium Compound". Angew. Chem. Int. Ed. 33 (6): 681–683. doi:10.1002/anie.199406811 (https://doi.org/10.100
2%2Fanie.199406811).
08. Pb(−2) has been observed in BaPb, see Ferro, Riccardo (2008). Nicholas C. Norman (ed.). Intermetallic Chemistry.
Elsevier. p. 505. ISBN 978-0-08-044099-6. and Todorov, Iliya; Sevov, Slavi C. (2004). "Heavy-Metal Aromatic Rings:
Cyclopentadienyl Anion Analogues Sn56− and Pb56− in the Zintl Phases Na8BaPb6, Na8BaSn6, and
Na8EuSn6". Inorganic Chemistry. 43 (20): 6490–94. doi:10.1021/ic000333x (https://doi.org/10.1021%2Fic000333x).
09. Pb(+1) and Pb(+3) have been observed in organolead compounds, e.g. hexamethyldiplumbane Pb2(CH3)6; for
Pb(I), see Siew-Peng Chia; Hong-Wei Xi; Yongxin Li; Kok Hwa Lim; Cheuk-Wai So (2013). "A Base-Stabilized
Lead(I) Dimer and an Aromatic Plumbylidenide Anion". Angew. Chem. Int. Ed. 52 (24): 6298–6301.
doi:10.1002/anie.201301954 (https://doi.org/10.1002%2Fanie.201301954). PMID 23629949 (https://pubmed.ncbi.nl
m.nih.gov/23629949).
10. Bi(−2) and Bi(−1) occur in Zintl phases, e.g. (Ca2+ )22[Bi4]4− ([Bi2]4− )4[Bi3− ]8; see Ponou, Siméon (2006).
"Germanides, Germanide-Tungstate Double Salts and Substitution Effects in Zintl Phases" (http://d-nb.info/9855276
76/34?origin=publication_detailSim). Technische Universität München. Lehrstuhl für Anorganische Chemie mit
Schwerpunkt Neue Materialien. p. 68.
11. Bi(I) has been observed in bismuth monobromide (BiBr) and bismuth monoiodide (BiI); see Godfrey, S. M.;
McAuliffe, C. A.; Mackie, A. G.; Pritchard, R. G. (1998). Nicholas C. Norman (ed.). Chemistry of arsenic, antimony,
and bismuth. Springer. pp. 67–84. ISBN 978-0-7514-0389-3.
12. Bi(+2) has been observed in dibismuthines (R2Bi—BiR2), see Arthur J. Ashe III (1990). Thermochromic Distibines
and Dibismuthines. Advances in Organometallic Chemistry. 30. pp. 77–97. doi:10.1016/S0065-3055(08)60499-2 (ht
tps://doi.org/10.1016%2FS0065-3055%2808%2960499-2). ISBN 9780120311309.
13. Bi(IV) has been observed; see A. I. Aleksandrov, I. E. Makarov (1987). "Formation of Bi(II) and Bi(IV) in aqueous
hydrochloric solutions of Bi(III)". Bulletin of the Academy of Sciences of the USSR, Division of Chemical Science.
36 (2): 217–220. doi:10.1007/BF00959349 (https://doi.org/10.1007%2FBF00959349).
14. Po(V) has been observed in dioxidopolonium(1+) (PoO+); see Thayer, John S. (2010). "Relativistic Effects and the
Chemistry of the Heavier Main Group Elements". Relativistic Methods for Chemists. p. 78. doi:10.1007/978-1-4020-
9975-5_2 (https://doi.org/10.1007%2F978-1-4020-9975-5_2). ISBN 978-1-4020-9974-8.
15. Rn(II) has been observed in radon difluoride (RnF2); see Stein, L. (1970). "Ionic Radon Solution". Science. 168
(3929): 362–4. Bibcode:1970Sci...168..362S (https://ui.adsabs.harvard.edu/abs/1970Sci...168..362S).
doi:10.1126/science.168.3929.362 (https://doi.org/10.1126%2Fscience.168.3929.362). PMID 17809133 (https://pub
med.ncbi.nlm.nih.gov/17809133). and Kenneth S. Pitzer (1975). "Fluorides of radon and element 118". J. Chem.
Soc., Chem. Commun. (18): 760b–761. doi:10.1039/C3975000760b (https://doi.org/10.1039%2FC3975000760b).
16. Rn(IV) is reported by Greenwood and Earnshaw, but is not known to exist; see Sykes, A. G. (1998). "Recent
Advances in Noble-Gas Chemistry" (https://books.google.com/books?id=6iqXRtz6p3QC). Advances in Inorganic
Chemistry. 46. Academic Press. pp. 91–93. ISBN 978-0-12-023646-6. Retrieved 22 November 2012.
17. Rn(VI) is known in radon trioxide (RnO3); see Sykes, A. G. (1998). "Recent Advances in Noble-Gas Chemistry" (http
s://books.google.com/books?id=6iqXRtz6p3QC). Advances in Inorganic Chemistry. 46. Academic Press. pp. 91–93.
ISBN 978-0-12-023646-6. Retrieved 22 November 2012.
18. Th(I) is known in thorium(I) bromide (ThBr); see Wickleder, Mathias S.; Fourest, Blandine; Dorhout, Peter K. (2006).
"Thorium". In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and
Transactinide Elements (https://web.archive.org/web/20160307160941/http://radchem.nevada.edu/classes/rdch710/
files/Thorium.pdf) (PDF). 3 (3rd ed.). Dordrecht, the Netherlands: Springer. pp. 52–160. doi:10.1007/1-4020-3598-
5_3 (https://doi.org/10.1007%2F1-4020-3598-5_3). ISBN 978-1-4020-3555-5. Archived from the original (http://radch
em.nevada.edu/classes/rdch710/files/thorium.pdf) (PDF) on 2016-03-07.
19. Th(II) and Th(III) are observed in [ThII{η5-C5H3(SiMe3)2}3]− and [ThIII{η5-C5H3(SiMe3)2}3], see Langeslay, Ryan R.;
Fieser, Megan E.; Ziller, Joseph W.; Furche, Philip; Evans, William J. (2015). "Synthesis, structure, and reactivity of
crystalline molecular complexes of the {[C5H3(SiMe3)2]3Th}1− anion containing thorium in the formal +2 oxidation
state" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5811171). Chem. Sci. 6 (1): 517–521.
doi:10.1039/C4SC03033H (https://doi.org/10.1039%2FC4SC03033H). PMC 5811171 (https://www.ncbi.nlm.nih.go
v/pmc/articles/PMC5811171). PMID 29560172 (https://pubmed.ncbi.nlm.nih.gov/29560172).
20. U(I) has been observed in uranium monofluoride (UF) and uranium monochloride (UCl), see Sykes, A. G. (1990).
"Compounds of Thorium and Uranium" (https://books.google.com/books?id=MZRm6E9LmgMC). Advances in
Inorganic Chemistry. 34. Academic Press. pp. 87–88. ISBN 978-0-12-023634-3. Retrieved 22 March 2015.
21. U(II) has been observed in [K(2.2.2-Cryptand)][(C5H4SiMe3)3U], see MacDonald, Matthew R.; Fieser, Megan E.;
Bates, Jefferson E.; Ziller, Joseph W.; Furche, Filipp; Evans, William J. (2013). "Identification of the +2 Oxidation
State for Uranium in a Crystalline Molecular Complex, [K(2.2.2-Cryptand)][(C5H4SiMe3)3U]". J. Am. Chem. Soc.
135 (36): 13310–13313. doi:10.1021/ja406791t (https://doi.org/10.1021%2Fja406791t). PMID 23984753 (https://pub
med.ncbi.nlm.nih.gov/23984753).
22. Np(II), (III) and (IV) have been observed, see Dutkiewicz, Michał S.; Apostolidis, Christos; Walter, Olaf; Arnold, Polly
L (2017). "Reduction chemistry of neptunium cyclopentadienide complexes: from structure to understanding" (https://
www.ncbi.nlm.nih.gov/pmc/articles/PMC5431675). Chem. Sci. 8 (4): 2553–2561. doi:10.1039/C7SC00034K (https://
doi.org/10.1039%2FC7SC00034K). PMC 5431675 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5431675).
PMID 28553487 (https://pubmed.ncbi.nlm.nih.gov/28553487).
23. Pu(II) has been observed in {Pu[C5H3(SiMe3)2]3}−; see Windorff, Cory J.; Chen, Guo P; Cross, Justin N; Evans,
William J.; Furche, Filipp; Gaunt, Andrew J.; Janicke, Michael T.; Kozimor, Stosh A.; Scott, Brian L. (2017).
"Identification of the Formal +2 Oxidation State of Plutonium: Synthesis and Characterization ofref name="curium5"
{PuII[C5H3(SiMe3)2]3}−". J. Am. Chem. Soc. 139 (11): 3970–3973. doi:10.1021/jacs.7b00706 (https://doi.org/10.102
1%2Fjacs.7b00706). PMID 28235179 (https://pubmed.ncbi.nlm.nih.gov/28235179).
24. Am(VII) has been observed in AmO5-; see Americium (http://www.chemie-master.de/FrameHandler.php?loc=http://w
ww.chemie-master.de/pse/pse.php?modul=Am), Das Periodensystem der Elemente für den Schulgebrauch (The
periodic table of elements for schools) chemie-master.de (in German), Retrieved 28 November 2010 and
Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann.
p. 1265. ISBN 978-0-08-037941-8.
25. Cm(V), Bk(V), and Cf(V) have been observed in BkO2+ , CfO2+ , CmO2(NO3)2− , BkO2(NO3)2− , and CfO2(NO3)2− ; see
Dau, Phuong Diem; Vasiliu, Monica; Peterson, Kirk A; Dixon, David A; Gibsoon, John K (October 2017).
"Remarkably High Stability of Late Actinide Dioxide Cations: Extending Chemistry to Pentavalent Berkelium and
Californium" (http://www.escholarship.org/uc/item/1mk7j6ps). Chemistry - A European Journal. 23 (68): 17369–
17378. doi:10.1002/chem.201704193 (https://doi.org/10.1002%2Fchem.201704193). PMID 29024093 (https://pubm
ed.ncbi.nlm.nih.gov/29024093).
26. Kovács, Attila; Dau, Phuong D.; Marçalo, Joaquim; Gibson, John K. (2018). "Pentavalent Curium, Berkelium, and
Californium in Nitrate Complexes: Extending Actinide Chemistry and Oxidation States". Inorg. Chem. American
Chemical Society. 57 (15): 9453–9467. doi:10.1021/acs.inorgchem.8b01450 (https://doi.org/10.1021%2Facs.inorgc
hem.8b01450). PMID 30040397 (https://pubmed.ncbi.nlm.nih.gov/30040397).
27. Cm(VI) has been observed in curium trioxide (CmO3) and dioxidocurium(2+) (CmO2+); see Domanov, V. P.;
Lobanov, Yu. V. (October 2011). "Formation of volatile curium(VI) trioxide CmO3". Radiochemistry. 53 (5): 453–6.
doi:10.1134/S1066362211050018 (https://doi.org/10.1134%2FS1066362211050018).
28. Cm(VIII) has been reported to possibly occur in curium tetroxide (CmO4); see Domanov, V. P. (January 2013).
"Possibility of generation of octavalent curium in the gas phase in the form of volatile tetraoxide CmO4".
Radiochemistry. 55 (1): 46–51. doi:10.1134/S1066362213010098 (https://doi.org/10.1134%2FS106636221301009
8). However, new experiments seem to indicate its nonexistence: Zaitsevskii, Andréi; Schwarz, W H Eugen (April
2014). "Structures and stability of AnO4 isomers, An = Pu, Am, and Cm: a relativistic density functional study".
Physical Chemistry Chemical Physics. 2014 (16): 8997–9001. Bibcode:2014PCCP...16.8997Z (https://ui.adsabs.har
vard.edu/abs/2014PCCP...16.8997Z). doi:10.1039/c4cp00235k (https://doi.org/10.1039%2Fc4cp00235k).
PMID 24695756 (https://pubmed.ncbi.nlm.nih.gov/24695756).
29. Peterson, J. R.; Hobart, D. E. (1984). "The Chemistry of Berkelium" (https://books.google.com/books?id=U-YOlLVuV
1YC&pg=PA29). In Emeléus, Harry Julius (ed.). Advances in inorganic chemistry and radiochemistry (https://archiv
e.org/details/advancesininorga0000unse_p1d0/page/29). 28. Academic Press. pp. 29–64 (https://archive.org/detail
s/advancesininorga0000unse_p1d0/page/29). doi:10.1016/S0898-8838(08)60204-4 (https://doi.org/10.1016%2FS0
898-8838%2808%2960204-4). ISBN 978-0-12-023628-2.
30. Peterson 1984, p. 55.
31. Sullivan, Jim C.; Schmidt, K. H.; Morss, L. R.; Pippin, C. G.; Williams, C. (1988). "Pulse radiolysis studies of
berkelium(III): preparation and identification of berkelium(II) in aqueous perchlorate media". Inorganic Chemistry. 27
(4): 597. doi:10.1021/ic00277a005 (https://doi.org/10.1021%2Fic00277a005).
32. Es(IV) is known in einsteinium(IV) fluoride (EsF4); see Kleinschmidt, P (1994). "Thermochemistry of the actinides".
Journal of Alloys and Compounds. 213–214: 169–172. doi:10.1016/0925-8388(94)90898-2 (https://doi.org/10.101
6%2F0925-8388%2894%2990898-2).
33. Db(V) has been observed in dubnium pentachloride (DbCl5); see H. W. Gäggeler (2007). "Gas Phase Chemistry of
Superheavy Elements" (https://web.archive.org/web/20120220090755/http://lch.web.psi.ch/files/lectures/TexasA%2
6M/TexasA%26M.pdf) (PDF). Paul Scherrer Institute. pp. 26–28. Archived from the original (http://lch.web.psi.ch/file
s/lectures/TexasA&M/TexasA&M.pdf) (PDF) on 2012-02-20.
34. Sg(VI) has been observed in seaborgium oxide hydroxide (SgO2(OH)2); see Huebener, S.; Taut, S.; Vahle, A.;
Dressler, R.; Eichler, B.; Gäggeler, H. W.; Jost, D.T.; Piguet, D.; et al. (2001). "Physico-chemical characterization of
seaborgium as oxide hydroxide" (https://web.archive.org/web/20141025201143/http://www-w2k.gsi.de/kernchemie/i
mages/pdf_Artikel/Radiochim_Acta_89_737_2001.pdf) (PDF). Radiochim. Acta. 89 (11–12_2001): 737–741.
doi:10.1524/ract.2001.89.11-12.737 (https://doi.org/10.1524%2Fract.2001.89.11-12.737). Archived from the original
(http://www-w2k.gsi.de/kernchemie/images/pdf_Artikel/Radiochim_Acta_89_737_2001.pdf) (PDF) on 2014-10-25.
35. Sg(0) has been observed in seaborgium hexacarbonyl (Sg(CO)6); see Even, J.; Yakushev, A.; Dullmann, C. E.;
Haba, H.; Asai, M.; Sato, T. K.; Brand, H.; Di Nitto, A.; Eichler, R.; Fan, F. L.; Hartmann, W.; Huang, M.; Jager, E.; Kaji,
D.; Kanaya, J.; Kaneya, Y.; Khuyagbaatar, J.; Kindler, B.; Kratz, J. V.; Krier, J.; Kudou, Y.; Kurz, N.; Lommel, B.;
Miyashita, S.; Morimoto, K.; Morita, K.; Murakami, M.; Nagame, Y.; Nitsche, H.; et al. (2014). "Synthesis and
detection of a seaborgium carbonyl complex". Science. 345 (6203): 1491–3. Bibcode:2014Sci...345.1491E (https://u
i.adsabs.harvard.edu/abs/2014Sci...345.1491E). doi:10.1126/science.1255720 (https://doi.org/10.1126%2Fscience.
1255720). PMID 25237098 (https://pubmed.ncbi.nlm.nih.gov/25237098). (subscription required)
36. Bh(VII) has been observed in bohrium oxychloride (BhO3Cl); see "Gas chemical investigation of bohrium (Bh,
element 107)" (http://www.gsi.de/informationen/wti/library/scientificreport2000/Chemistry/9/r_eichler_jb2000.pdf)
Archived (https://web.archive.org/web/20080228023225/http://www.gsi.de/informationen/wti/library/scientificreport20
00/Chemistry/9/r_eichler_jb2000.pdf) 2008-02-28 at the Wayback Machine, Eichler et al., GSI Annual Report 2000.
Retrieved on 2008-02-29
37. Hs(VIII) has been observed in hassium tetroxide (HsO4); see "Chemistry of Hassium" (http://www.gsi.de/documents/
DOC-2003-Jun-29-2.pdf) (PDF). Gesellschaft für Schwerionenforschung mbH. 2002. Retrieved 2007-01-31.
38. Cn(II) has been observed in copernicium selenide (CnSe); see "Annual Report 2015: Laboratory of Radiochemistry
and Environmental Chemistry" (https://www.psi.ch/luc/AnnualReportsEN/PSI_LCH_AnnualReport2015.pdf) (PDF).
Paul Scherrer Institute. 2015. p. 3.
39. Langmuir, Irving (1919). "The arrangement of electrons in atoms and molecules" (https://zenodo.org/record/142902
6). J. Am. Chem. Soc. 41 (6): 868–934. doi:10.1021/ja02227a002 (https://doi.org/10.1021%2Fja02227a002).
40. Connelly, N. G.; Damhus, T.; Hartshorn, R. M.; Hutton, A. T. Nomenclature of Inorganic Chemistry (IUPAC
Recommendations 2005) (http://www.old.iupac.org/publications/books/rbook/Red_Book_2005.pdf) (PDF). RSC
Publishing.
41. "Antoine Laurent Lavoisier The Chemical Revolution - Landmark - American Chemical Society" (https://www.acs.or
g/content/acs/en/education/whatischemistry/landmarks/lavoisier.html). American Chemical Society. Retrieved
14 July 2018.
42. "Lavoisier on Elements" (http://chem125-oyc.webspace.yale.edu/125/history99/2Pre1800/Lavoisier/Nomenclature/L
avoisier_on_Elements.html). Chem125-oyc.webspace.yale.edu. Retrieved 14 July 2018.
43. Wöhler, F. (1835). Grundriss der Chemie: Unorganische Chemie [Foundations of Chemistry: Inorganic Chemistry].
Berlin: Duncker und Humblot. p. 4.
44. Jensen, W. B. (2007). "the origin of the oxidation-state concept". J. Chem. Educ. 84 (9): 1418–1419.
doi:10.1021/ed084p1418 (https://doi.org/10.1021%2Fed084p1418).
45. Stock, A. (1919). "Einige Nomenklaturfragen der anorganischen Chemie" (https://zenodo.org/record/1424478)
[Some nomenclature issues of inorganic chemistry]. Angew. Chem. 32 (98): 373–374.
doi:10.1002/ange.19190329802 (https://doi.org/10.1002%2Fange.19190329802).
46. Jorissen, W. P.; Bassett, H.; Damiens, A.; Fichter, F.; Rémy, H. (1941). "Rules for naming inorganic compounds". J.
Am. Chem. Soc. 63: 889–897. doi:10.1021/ja01849a001 (https://doi.org/10.1021%2Fja01849a001).
47. Latimer, W. M. (1938). The Oxidation States of the Elements and their Potentials in Aqueous Solutions (1st ed.).
Prentice-Hall.
48. Bray, W. C.; Branch, G. E. K. (1913). "Valence and tautomerism" (https://zenodo.org/record/1428999/files/article.pdf)
(PDF). J. Am. Chem. Soc. 35 (10): 1440–1447. doi:10.1021/ja02199a003 (https://doi.org/10.1021%2Fja02199a003).
49. Noyes, A. A.; Pitzer, K. S.; Dunn, C. L. (1935). "Argentic salts in acid solution, I. The oxidation and reduction
reactions". J. Am. Chem. Soc. 57 (7): 1221–1229. doi:10.1021/ja01310a018 (https://doi.org/10.1021%2Fja01310a0
18).
50. Noyes, A. A.; Pitzer, K. S.; Dunn, C. L. (1935). "Argentic salts in acid solution, II. The oxidation state of argentic
salts". J. Am. Chem. Soc. 57 (7): 1229–1237. doi:10.1021/ja01310a019 (https://doi.org/10.1021%2Fja01310a019).
51. Fernelius, W. C. (1948). "Some problems of inorganic nomenclature". Chem. Eng. News. 26: 161–163.
doi:10.1021/cen-v026n003.p161 (https://doi.org/10.1021%2Fcen-v026n003.p161).
52. Fernelius, W. C.; Larsen, E. M.; Marchi, L. E.; Rollinson, C. L. (1948). "Nomenclature of coördination compounds".
Chem. Eng. News. 26 (8): 520–523. doi:10.1021/cen-v026n008.p520 (https://doi.org/10.1021%2Fcen-v026n008.p5
20).
53. Pauling, L. (1948). "The modern theory of valency". J. Chem. Soc. 1948: 1461–1467. doi:10.1039/JR9480001461 (h
ttps://doi.org/10.1039%2FJR9480001461).
54. Calvert, J. G. (1990). "IUPAC Recommendation 1990" (https://doi.org/10.1351/pac199062112167). Pure Appl.
Chem. 62: 2204. doi:10.1351/pac199062112167 (https://doi.org/10.1351%2Fpac199062112167).
Retrieved from "https://en.wikipedia.org/w/index.php?title=Oxidation_state&oldid=975839976"

This page was last edited on 30 August 2020, at 17:57 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this site, you agree to the
Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.

You might also like