You are on page 1of 22

Non-stoichiometric

compound

Origin of title phenomenon in crystallographic


defects. Shown is a two-dimensional slice through a
primitive cubic crystal system showing the regular
square array of atoms on one face (open circles, o),
and with these, places where atoms are missing from
a regular site to create vacancies, displaced to an
adjacent acceptable space to create a Frenkel pair, or
substituted by a smaller or larger atom not usually
seen (closed circles, • ), in each case resulting in a
material that is moved toward being measurably non-
stoichiometric.

Non-stoichiometric compounds are


chemical compounds, almost always solid
inorganic compounds, having elemental
composition whose proportions cannot be
represented by a ratio of small natural
numbers; most often, in such materials,
some small percentage of atoms are
missing or too many atoms are packed
into an otherwise perfect lattice work.
Contrary to earlier definitions, modern
understanding of non-stoichiometric
compounds view them as homogenous,
and not mixtures of stoichiometric
chemical compounds. Since the solids are
overall electrically neutral, the defect is
compensated by a change in the charge of
other atoms in the solid, either by
changing their oxidation state, or by
replacing them with atoms of different
elements with a different charge. Many
metal oxides and sulfides have non-
stoichiometric examples; for example,
stoichiometric iron(II) oxide, which is rare,
has the formula FeO, whereas the more
common material is nonstoichiometric,
with the formula Fe0.95O. The type of
equilibrium defects in non-stoichiometric
compounds can vary with attendant
variation in bulk properties of the
material.[1] Non-stoichiometric compounds
also exhibit special electrical or chemical
properties because of the defects; for
example, when atoms are missing,
electrons can move through the solid more
rapidly. Non-stoichiometric compounds
have applications in ceramic and
superconductive material and in
electrochemical (i.e., battery) system
designs.
Occurrence
This section needs expansion with: more general
information, with sources, on the scopeLearn
of the
more

Iron oxides …

Nonstoichiometry is pervasive for metal


oxides, especially when the metal is not in
its highest oxidation state.[2]:642–644 For
example, although wüstite (ferrous oxide)
has an ideal (stoichiometric) formula FeO,
the actual stoichiometry is closer to
Fe0.95O. The non-stoichiometry reflect the
ease of oxidation of Fe2+ to Fe3+
effectively replacing a small portion of
Fe2+ with two thirds their number of Fe3+.
Thus for every three "missing" Fe2+ ions,
the crystal contains two Fe3+ ions to
balance the charge. The composition of a
non-stoichiometric compound usually
varies in a continuous manner over a
narrow range. Thus, the formula for
wüstite is written as Fe1−xO, where x is a
small number (0.05 in the previous
example) representing the deviation from
the "ideal" formula.[3] Nonstoichiometry is
especially important in solid, three-
dimensional polymers that can tolerate
mistakes. To some extent, entropy drives
all solids to be non-stoichiometric. But for
practical purposes, the term describes
materials where the non-stoichiometry is
measurable, usually at least 1% of the
ideal composition.

Iron sulfides …

Pyrrhotite, an example of a non-stoichiometric


inorganic compound, with formula Fe1−xS (x = 0 to
0.2).

The monosulfides of the transition metals


are often nonstoichiometric. Best known
perhaps is nominally iron(II) sulfide (the
mineral pyrrhotite) with a composition
Fe1−xS (x = 0 to 0.2). The rare
stoichiometric FeS endmember is known
as the mineral troilite. Pyrrhotite is
remarkable in that it has numerous
polytypes, i.e. crystalline forms differing in
symmetry (monoclinic or hexagonal) and
composition (Fe7S8, Fe9S10, Fe11S12 and
others). These materials are always iron-
deficient owing to the presence of lattice
defects, namely iron vacancies. Despite
those defects, the composition is usually
expressed as a ratio of large numbers and
the crystals symmetry is relatively high.
This means the iron vacancies are not
randomly scattered over the crystal, but
form certain regular configurations. Those
vacancies strongly affect the magnetic
properties of pyrrhotite: the magnetism
increases with the concentration of
vacancies and is absent for the
stoichiometric FeS.[4]

Palladium hydrides …

Palladium hydride is a nonstoichiometric


material of the approximate composition
PdHx (0.02 < x < 0.58). This solid conducts
hydrogen by virtue of the mobility of the
hydrogen atoms within the solid.
Tungsten oxides …

It is sometimes difficult to determine if a


material is non-stoichiometric or if the
formula is best represented by large
numbers. The oxides of tungsten illustrate
this situation. Starting from the idealized
material tungsten trioxide, one can
generate a series of related materials that
are slightly deficient in oxygen. These
oxygen-deficient species can be described
as WO3−x, but in fact they are
stoichiometric species with large unit cells
with the formulas WnO3n−2, where n = 20,
24, 25, 40. Thus, the last species can be
described with the stoichiometric formula
W40O118, whereas the non-stoichiometric
description WO2.95 implies a more random
distribution of oxide vacancies.

Other cases …

At high temperatures (1000 °C), titanium


sulfides present a series of non-
stoichiometric compounds.[2]:679

The coordination polymer Prussian blue,


nominally Fe7(CN)18 and their analogs are
well known to form in non-stoichiometric
proportions.[5]:114 The non-stoichiometric
phases exhibit useful properties vis-à-vis
their ability to bind caesium and thallium
ions.

Applications

Oxidation catalysis …

Many useful compounds are produced by


the reactions of hydrocarbons with
oxygen, a conversion that is catalyzed by
metal oxides. The process operates via the
transfer of "lattice" oxygen to the
hydrocarbon substrate, a step that
temporarily generates a vacancy (or
defect). In a subsequent step, the missing
oxygen is replenished by O2. Such
catalysts rely on the ability of the metal
oxide to form phases that are not
stoichiometric.[6] An analogous sequence
of events describes other kinds of atom-
transfer reactions including hydrogenation
and hydrodesulfurization catalysed by
solid catalysts. These considerations also
highlight the fact that stoichiometry is
determined by the interior of crystals: the
surfaces of crystals often do not follow
the stoichiometry of the bulk. The complex
structures on surfaces are described by
the term "surface reconstruction".

Ion conduction …
The migration of atoms within a solid is
strongly influenced by the defects
associated with non-stoichiometry. These
defect sites provide pathways for atoms
and ions to migrate through the otherwise
dense ensemble of atoms that form the
crystals. Oxygen sensors and solid state
batteries are two applications that rely on
oxide vacancies. One example is the CeO2-
based sensor in automotive exhaust
systems. At low partial pressures of O2,
the sensor allows the introduction of
increased air to effect more thorough
combustion.[6]
Superconductivity …

Many superconductors are non-


stoichiometric. For example, yttrium
barium copper oxide, arguably the most
notable high-temperature superconductor,
is a non-stoichiometric solid with the
formula YxBa2Cu3O7−x. The critical
temperature of the superconductor
depends on the exact value of x. The
stoichiometric species has x = 0, but this
value can be as great as 1.[6]

History
It was mainly through the work of Nikolai
Semenovich Kurnakov and his students
that Berthollet's opposition to Proust's law
was shown to have merit for many solid
compounds. Kurnakov divided non-
stoichiometric compounds into
berthollides and daltonides depending on
whether their properties showed
monotonic behavior with respect to
composition or not. The term berthollide
was accepted by IUPAC in 1960.[7] The
names come from Claude Louis Berthollet
and John Dalton, respectively, who in the
19th century advocated rival theories of
the composition of substances. Although
Dalton "won" for the most part, it was later
recognized that the law of definite
proportions had important exceptions.[8]
See also
F-Center
Vacancy defect

References
1. Geng, Hua Y.; et al. (2012). "Anomalies
in nonstoichiometric uranium dioxide
induced by a pseudo phase transition
of point defects". Phys. Rev. B. 85 (14):
144111. arXiv:1204.4607 .
doi:10.1103/PhysRevB.85.144111 .
2. N. N. Greenwood & A. Earnshaw, 2012,
"Chemistry of the Elements," 2nd Edn.,
Amsterdam, NH, NLD:Elsevier,
ISBN 0080501095, see [1] , accessed
8 July 2015. [Page numbers marked by
superscript, inline.]
3. Lesley E. Smart (2005). Solid State
Chemistry: An Introduction, 3rd
edition. CRC Press. p. 214. ISBN 978-
0-7487-7516-3.
4. Hubert Lloyd Barnes (1997).
Geochemistry of hydrothermal ore
deposits . John Wiley and Sons.
pp. 382–390. ISBN 978-0-471-57144-
5.
5. Metal-Organic and Organic Molecular
Magnets P D ,A EU
Royal Society of Chemistry, 2007,
ISBN 1847551394,
ISBN 9781847551399
. Atkins, P. W.; Overton, T. L.; Rourke, J.
P.; Weller, M. T.; Armstrong, F. A., 2010,
Shriver and Atkins' Inorganic
Chemistry 5th Edn., pp. 65, 75, 99f,
268, 271, 277, 287, 356, 409, Oxford,
OXF, GBR: Oxford University Press,
ISBN 0199236178, see [2] , accessed
8 July 2015.
7. The Rare Earth Trifluorides, Part 2
Arxius de les Seccions de Ciències
D N. K ,B P
S ,I V. A , Institut
d'Estudis Catalans, 2000, p75ff.
ISBN 847283610X,
ISBN 9788472836105
. Henry Marshall Leicester (1971). The
Historical Background of Chemistry.
Courier Dover Publications. p. 153.
ISBN 9780486610535.

Further reading
F. Albert Cotton, Geoffrey Wilkinson,
Carlos A. Murillo & Manfred Bochmann,
1999, Advanced Inorganic Chemistry, 6th
Edn., pp. 202, 271, 316, 777, 888. 897,
and 1145, New York, NY, USA:Wiley-
Interscience, ISBN 0471199575, see [3] ,
accessed 8 July 2015.
Roland Ward, 1963, Nonstoichiometric
Compounds, Advances in Chemistry
series, Vol. 39, Washington, DC, USA:
American Chemical Society,
ISBN 9780841222076, DOI 10.1021/ba-
1964-0039, see [4] , accessed 8 July
2015.
J. S. Anderson, 1963, "Current problems
in nonstoichiometry (Ch. 1)," in
Nonstoichiometric Compounds (Roland
Ward, Ed.), pp. 1–22, Advances in
Chemistry series, Vol. 39, Washington,
DC, USA: American Chemical Society,
ISBN 9780841222076, DOI 10.1021/ba-
1964-0039.ch001, see [5] , accessed 8
July 2015.
Retrieved from
"https://en.wikipedia.org/w/index.php?title=Non-
stoichiometric_compound&oldid=968176501"

Last edited 2 days ago by Vsmith

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like