Mineral
Mineral
Mineral
Definition
Basic definition
One definition of a mineral encompasses the following criteria:[3]
Chemistry
The abundance and diversity of minerals is controlled directly by their chemistry, in turn
dependent on elemental abundances in the Earth. The majority of minerals observed are derived
from the Earth's crust. Eight elements account for most of the key components of minerals, due
to their abundance in the crust. These eight elements, summing to over 98% of the crust by
weight, are, in order of decreasing
abundance: oxygen, silicon, aluminium, iron, magnesium, calcium, sodium and potassium.
Oxygen and silicon are by far the two most important – oxygen composes 47% of the crust by
weight, and silicon accounts for 28%.[42]
The minerals that form are directly controlled by the bulk chemistry of the parent body. For
example, a magma rich in iron and magnesium will form mafic minerals, such as olivine and the
pyroxenes; in contrast, a more silica-rich magma will crystallize to form minerals that
incorporate more SiO2, such as the feldspars and quartz. In a limestone, calcite or aragonite (both
CaCO3) form because the rock is rich in calcium and carbonate. A corollary is that a mineral will
not be found in a rock whose bulk chemistry does not resemble the bulk chemistry of a given
mineral with the exception of trace minerals. For example, kyanite, Al2SiO5 forms from
the metamorphism of aluminium-rich shales; it would not likely occur in aluminium-poor rock,
such as quartzite.
The chemical composition may vary between end member species of a solid solution series. For
example, the plagioclase feldsparscomprise a continuous series from sodium-rich end
member albite (NaAlSi3O8) to calcium-rich anorthite (CaAl2Si2O8) with four recognized
intermediate varieties between them (given in order from sodium- to calcium-
rich): oligoclase, andesine, labradorite, and bytownite.[43] Other examples of series include the
olivine series of magnesium-rich forsterite and iron-rich fayalite, and the wolframiteseries
of manganese-rich hübnerite and iron-rich ferberite.
Chemical substitution and coordination polyhedra explain this common feature of minerals. In
nature, minerals are not pure substances, and are contaminated by whatever other elements are
present in the given chemical system. As a result, it is possible for one element to be substituted
for another.[44] Chemical substitution will occur between ions of a similar size and charge; for
example, K+ will not substitute for Si4+ because of chemical and structural incompatibilities
caused by a big difference in size and charge. A common example of chemical substitution is
that of Si4+ by Al3+, which are close in charge, size, and abundance in the crust. In the example of
plagioclase, there are three cases of substitution. Feldspars are all framework silicates, which
have a silicon-oxygen ratio of 2:1, and the space for other elements is given by the substitution of
Si4+ by Al3+ to give a base unit of [AlSi3O8]−; without the substitution, the formula would be
charge-balanced as SiO2, giving quartz.[45] The significance of this structural property will be
explained further by coordination polyhedra. The second substitution occurs between Na+ and
Ca2+; however, the difference in charge has to accounted for by making a second substitution of
Si4+ by Al3+.[46]
Coordination polyhedra are geometric representation of how a cation is surrounded by an anion.
In mineralogy, coordination polyhedra are usually considered in terms of oxygen, due its
abundance in the crust. The base unit of silicate minerals is the silica tetrahedron – one
Si4+ surrounded by four O2−. An alternate way of describing the coordination of the silicate is by
a number: in the case of the silica tetrahedron, the silicon is said to have a coordination number
of 4. Various cations have a specific range of possible coordination numbers; for silicon, it is
almost always 4, except for very high-pressure minerals where compound is compressed such
that silicon is in six-fold (octahedral) coordination by oxygen. Bigger cations have a bigger
coordination number because of the increase in relative size as compared to oxygen (the
last orbital subshell of heavier atoms is different too). Changes in coordination numbers between
leads to physical and mineralogical differences; for example, at high pressure such as in
the mantle, many minerals, especially silicates such as olivine and garnet will change to
a perovskite structure, where silicon is in octahedral coordination. Another example are the
aluminosilicates kyanite, andalusite, and sillimanite (polymorphs, as they share the formula
Al2SiO5), which differ by the coordination number of the Al3+; these minerals transition from
one another as a response to changes in pressure and temperature.[42] In the case of silicate
materials, the substitution of Si4+ by Al3+ allows for a variety of minerals because of the need to
balance charges.[47]
Changes in temperature and pressure, and composition alter the mineralogy of a rock sample.
Changes in composition can be caused by processes such
as weathering or metasomatism (hydrothermal alteration). Changes in temperature and pressure
occur when the host rock undergoes tectonic or magmatic movement into differing physical
regimes. Changes in thermodynamic conditions make it favourable for mineral assemblages to
react with each other to produce new minerals; as such, it is possible for two rocks to have an
identical or a very similar bulk rock chemistry without having a similar mineralogy. This process
of mineralogical alteration is related to the rock cycle. An example of a series of mineral
reactions is illustrated as follows.[48]
Orthoclase feldspar (KAlSi3O8) is a mineral commonly found in granite, a plutonic igneous rock.
When exposed to weathering, it reacts to form kaolinite (Al2Si2O5(OH)4, a sedimentary mineral,
and silicic acid):
2 KAlSi3O8 + 5 H2O + 2 H+ → Al2Si2O5(OH)4 + 4 H2SiO3 + 2 K+
Under low-grade metamorphic conditions, kaolinite reacts with quartz to
form pyrophyllite (Al2Si4O10(OH)2):
Al2Si2O5(OH)4 + SiO2 → Al2Si4O10(OH)2 + H2O
As metamorphic grade increases, the pyrophyllite reacts to form kyanite and quartz:
Al2Si4O10(OH)2 → Al2SiO5 + 3 SiO2 + H2O
Alternatively, a mineral may change its crystal structure as a consequence of changes in
temperature and pressure without reacting. For example, quartz will change into a variety of its
SiO2 polymorphs, such as tridymite and cristobalite at high temperatures, and coesite at high
pressures.[49]
Physical properties
Classifying minerals ranges from simple to difficult. A mineral can be identified by several
physical properties, some of them being sufficient for full identification without equivocation.
In other cases, minerals can only be classified by more complex optical, chemical or X-ray
diffraction analysis; these methods, however, can be costly and time-consuming. Physical
properties applied for classification include crystal structure and habit, hardness, lustre,
diaphaneity, colour, streak, cleavage and fracture, and specific gravity. Other less general tests
include fluorescence, phosphorescence, magnetism, radioactivity, tenacity (response to
[50]
mechanical induced changes of shape or form), piezoelectricity and reactivity to dilute acids.
Crystal structure and habit
Crystal structure results from the orderly geometric spatial arrangement of atoms in the internal
structure of a mineral. This crystal structure is based on regular internal atomic
or ionic arrangement that is often expressed in the geometric form that the crystal takes. Even
when the mineral grains are too small to see or are irregularly shaped, the underlying crystal
structure is always periodic and can be determined by X-ray diffraction.[3] Minerals are typically
described by their symmetry content. Crystals are restricted to 32 point groups, which differ by
their symmetry. These groups are classified in turn into more broad categories, the most
encompassing of these being the six crystal families.[51]
These families can be described by the relative lengths of the three crystallographic
axes, and the angles between them; these relationships correspond to the symmetry operations
that define the narrower point groups. They are summarized below; a, b, and c represent the axes,
and α, β, γ represent the angle opposite the respective crystallographic axis (e.g. α is the angle
opposite the a-axis, viz. the angle between the b and c axes):[51]
The hexagonal crystal family is also split into two crystal systems – the trigonal, which has a
three-fold axis of symmetry, and the hexagonal, which has a six-fold axis of symmetry.
Chemistry and crystal structure together define a mineral. With a restriction to 32 point groups,
minerals of different chemistry may have identical crystal structure. For
example, halite (NaCl), galena (PbS), and periclase (MgO) all belong to the hexaoctahedral point
group (isometric family), as they have a similar stoichiometry between their different constituent
elements. In contrast, polymorphs are groupings of minerals that share a chemical formula but
have a different structure. For example, pyriteand marcasite, both iron sulfides, have the formula
FeS2; however, the former is isometric while the latter is orthorhombic. This polymorphism
extends to other sulfides with the generic AX2 formula; these two groups are collectively known
as the pyrite and marcasite groups.[52]
Polymorphism can extend beyond pure symmetry content. The aluminosilicates are a group of
three minerals – kyanite, andalusite, and sillimanite – which share the chemical formula Al2SiO5.
Kyanite is triclinic, while andalusite and sillimanite are both orthorhombic and belong to the
dipyramidal point group. These difference arise correspond to how aluminium is coordinated
within the crystal structure. In all minerals, one aluminium ion is always in six-fold coordination
by oxygen; the silicon, as a general rule is in four-fold coordination in all minerals; an exception
is a case like stishovite (SiO2, an ultra-high pressure quartz polymorph with rutile
structure).[53] In kyanite, the second aluminium is in six-fold coordination; its chemical formula
can be expressed as Al[6]Al[6]SiO5, to reflect its crystal structure. Andalusite has the second
aluminium in five-fold coordination (Al[6]Al[5]SiO5) and sillimanite has it in four-fold
coordination (Al[6]Al[4]SiO5).[54]
Differences in crystal structure and chemistry greatly influence other physical properties of the
mineral. The carbon allotropes diamond and graphite have vastly different properties; diamond is
the hardest natural substance, has an adamantine lustre, and belongs to the isometric crystal
family, whereas as graphite is very soft, has a greasy lustre, and crystallises in the hexagonal
family. This difference is accounted by differences in bonding. In diamond, the carbons are in
sp3 hybrid orbitals, which means they form a framework where each carbon is covalently bonded
to four neighbours in a tetrahedral fashion; on the other hand, graphite is composed of sheets of
carbons in sp2 hybrid orbitals, where each carbon is bonded covalently to only three others.
These sheets are held together by much weaker van der Waals forces, and this discrepancy
translates to big macroscopic differences.[55]
Twinning is the intergrowth of two or more crystal of a single mineral species. The geometry of
the twinning is controlled by the mineral's symmetry. As a result, there are several types of
twins, including contact twins, reticulated twins, geniculated twins, penetration twins, cyclic
twins, and polysynthetic twins. Contact, or simple twins, consist of two crystals joined at a plane;
this type of twinning is common in spinel. Reticulated twins, common in rutile, are interlocking
crystals resembling netting. Geniculated twins have a bend in the middle that is caused by start
of the twin. Penetration twins consist of two single crystals that have grown into each other;
examples of this twinning include cross-shaped staurolite twins and Carlsbad twinning in
orthoclase. Cyclic twins are caused by repeated twinning around a rotation axis. It occurs around
three, four, five, six, or eight-fold axes, and the corresponding patterns are called threelings,
fourlings, fivelings, sixlings, and eightlings. Sixlings are common in aragonite. Polysynthetic
twins are similar to cyclic twinning by the presence of repetitive twinning; however, instead of
occurring around a rotational axis, it occurs along parallel planes, usually on a microscopic
scale.[56][57]
Crystal habit refers to the overall shape of crystal. Several terms are used to describe this
property. Common habits include acicular, which described needlelike crystals like in natrolite,
bladed, dendritic (tree-pattern, common in native copper), equant, which is typical of garnet,
prismatic (elongated in one direction), and tabular, which differs from bladed habit in that the
former is platy whereas the latter has a defined elongation. Related to crystal form, the quality of
crystal faces is diagnostic of some minerals, especially with a petrographic microscope. Euhedral
crystals have a defined external shape, while anhedral crystals do not; those intermediate forms
are termed subhedral.[58][59]
Hardness
The hardness of a mineral defines how much it can resist scratching. This physical property is
controlled by the chemical composition and crystalline structure of a mineral. A mineral's
hardness is not necessarily constant for all sides, which is a function of its structure;
crystallographic weakness renders some directions softer than others.[60] An example of this
property exists in kyanite, which has a Mohs hardness of 5½ parallel to [001] but 7 parallel to
[100].[61]
The most common scale of measurement is the ordinal Mohs hardness scale. Defined by ten
indicators, a mineral with a higher index scratches those below it. The scale ranges from talc,
a phyllosilicate, to diamond, a carbon polymorph that is the hardest natural material. The scale is
provided below:[60]
1 Talc Mg3Si4O10(OH)2
2 Gypsum CaSO4·2H2O
3 Calcite CaCO3
4 Fluorite CaF2
5 Apatite Ca5(PO4)3(OH,Cl,F)
6 Orthoclase KAlSi3O8
7 Quartz SiO2
8 Topaz Al2SiO4(OH,F)2
9 Corundum Al2O3
10 Diamond C