Mineral: Difference between revisions

Source: Wikipedia, the free encyclopedia.
Content deleted Content added
→‎Hardness: fix formulas, replace topaz and corundum images
Line 165: Line 165:
::Many minerals range from transparent to translucent or translucent to opaque. Calcite, for instance, can be translucent or opaque. Some minerals that are naturally translucent become opaque with weathering.
::Many minerals range from transparent to translucent or translucent to opaque. Calcite, for instance, can be translucent or opaque. Some minerals that are naturally translucent become opaque with weathering.


===Colour===
===Colour and streak===
{{main article|Streak (mineralogy)}}
[[File:Tourmaline.xtal.750pix.jpg|Watermelon [[elbaite]]|left|thumb]]
[[File:Tourmaline.xtal.750pix.jpg|Watermelon [[elbaite]]|left|thumb]]
Colour indicates the appearance of the mineral in reflected light or transmitted light for translucent minerals (i.e. what it looks like to the naked eye).
Colour indicates the appearance of the mineral in reflected light or transmitted light for translucent minerals (i.e. what it looks like to the naked eye).
*''Iridescence'' – the ''play of colors'' due to surface or internal interference. [[Labradorite]] exhibits internal iridescence whereas hematite and sphalerite often show the surface effect.
*''Iridescence'' – the ''play of colors'' due to surface or internal interference. [[Labradorite]] exhibits internal iridescence whereas hematite and sphalerite often show the surface effect.


===Streak===
{{main article|Streak (mineralogy)}}
[[File:Streak plate with Pyrite and Rhodochrosite.jpg|thumb|Streak plates with [[pyrite]] (left) and [[rhodochrosite]] (right)]]
[[File:Streak plate with Pyrite and Rhodochrosite.jpg|thumb|Streak plates with [[pyrite]] (left) and [[rhodochrosite]] (right)]]
Streak refers to the color of the powder a mineral leaves after rubbing it on an unglazed porcelain ''streak plate''. Note that this is not always the same color as the original mineral.
Streak refers to the color of the powder a mineral leaves after rubbing it on an unglazed porcelain ''streak plate''. Note that this is not always the same color as the original mineral.

Revision as of 22:02, 1 August 2012

Amethyst, a variety of quartz

A mineral is a naturally occurring

atomic structure, and specific physical properties. By comparison, a rock is an aggregate of minerals and/or mineraloids and does not have a specific chemical composition. Minerals range in composition from pure elements and simple salts to very complex silicates with thousands of known forms.[1] The study of minerals is called mineralogy
.

There are currently more than 4,000 known minerals, according to the International Mineralogical Association (IMA), which is responsible for the approval of and naming of new mineral species found in nature. Of these, perhaps 100 can be called "common", 50 are "occasional", and the rest are "rare" to "extremely rare".

Mineral definition and classification

To be classified as a true mineral, a substance must be a

crystalline
structure. It must also be a naturally occurring, homogeneous substance with a defined chemical composition.

The International Mineralogical Association approved the following definition in 1995:

"A mineral is an element or chemical compound that is normally crystalline and that has been formed as a result of geological processes."[2]

According to this definition and classification scheme, biogenic materials were excluded from the mineral kingdom:

"Biogenic substances are chemical compounds produced entirely by biological processes without a geological component (e.g., urinary calculi, oxalate crystals in plant tissues, shells of marine molluscs, etc.) and are not regarded as minerals. However, if geological processes were involved in the genesis of the compound, then the product can be accepted as a mineral."[2]: 690 

However, other researchers do not adhere to this exclusion rule. Lowenstam (1981), for example, states the following:[3]

"Organisms are capable of forming a diverse array of minerals, some of which cannot be formed inorganically in the biosphere.": 1126 

The distinction is a matter of classification and less to do with the constituents of the minerals themselves. Skinner (2005) views all solids as potential minerals and includes biominerals in the mineral kingdom, which are those that are created by the metabolic activities of organisms. Inclusion of these biogenic minerals requires an expanded definition of a mineral as:

"An element or compound, amorphous or crystalline, formed through biogeochemical processes."[4]: 621 

Mineral classification schemes and their definitions are evolving to match recent advances in mineral science. More recent classifications, for example, include an organic class – in both the new Dana and the

hydrocarbons. The IMA Commission on New Minerals and Mineral Names recently adopted (in 2009) a hierarchical scheme for the naming and classification of mineral groups and group names[7] and established seven commissions and four working groups to review and classify minerals into an official listing of their published names.[8] According to these new rules, "mineral species can be grouped in a number of different ways, on the basis of chemistry, crystal structure, occurrence, association, genetic history, or resource, for example, depending on the purpose to be served by the classification."[7]
: 1073 

Recent advances in high-resolution genetic and x-ray absorption spectroscopy is opening new revelations on the biogeochemical relations between microrganisms and minerals that may make Nickel's (1995)[2] biogenic mineral exclusion obsolete and Skinner's (2005) biogenic mineral inclusion a necessity.[4] For example, the IMA commissioned 'Environmental Mineralogy and Geochemistry Working Group'[9] deals with minerals in the hydrosphere, atmosphere, and biosphere. Mineral forming microorganisms inhabit the areas that this working group deals with. These organisms exist on nearly every rock, soil, and particle surface spanning the globe reaching depths at 1600 meters below the sea floor (possibly further) and 70 kilometers into the stratosphere (possibly entering the mesosphere).[10][11][12] Biologists and geologists have recently started to research and appreciate the magnitude of mineral geoengineering that these creatures are capable of. Bacteria have contributed to the formation of minerals for billions of years and critically define the biogeochemical cycles on this planet. Microorganisms can precipitate metals from solution contributing to the formation of ore deposits in addition to their ability to catalyze mineral dissolution, to respire, precipitate, and form minerals.[13][14][15]

Prior to the International Mineralogical Association's listing, over 60 biominerals had been discovered, named, and published.[16] These minerals (a sub-set tabulated in Lowenstam (1981)[3]) are considered minerals proper according to the Skinner (2005) definition.[4] These biominerals are not listed in the International Mineral Association official list of mineral names,[17] however, many of these biomineral representatives are distributed amongst the 78 mineral classes listed in the `Dana' classification scheme.[4] Another rare class of minerals (primarily biological in origin) include the mineral liquid crystals that are crystalline and liquid at the same time. To date over 80,000 liquid crystalline compounds have been identified.[18][19]

Concerning the use of the term “mineral” to name this family of liquid crystals, one can argue that the term inorganic would be more appropriate. However, inorganic liquid crystals have long been used for organometallic liquid crystals. Therefore in order to avoid any confusion between these fairly chemically different families, and taking into account that a large number of these liquid crystals occur naturally in nature, we think that the use of the old fashioned but adequate “mineral” adjective taken sensus largo is more specific that an alternative such as “purely inorganic”, to name this subclass of the inorganic liquid crystals family.[19]

The Skinner (2005) definition[4] of a mineral takes this matter into account by stating that a mineral can be crystalline or amorphous. Liquid mineral crystals are amorphous. Biominerals and liquid mineral crystals, however, are not the primary form of minerals, most are geological in origin,[20] but these groups do help to identify at the margins of what constitutes a mineral proper.

Mineral chemistry and nomenclature

The

end members of a mineral system. For example the plagioclase feldspars comprise a continuous series from sodium and silicon-rich albite (NaAlSi3O8) to calcium and aluminium-rich anorthite (CaAl2Si2O8) with four recognized intermediate compositions between. Mineral-like substances that don't strictly meet the definition are sometimes classified as mineraloids
.

Minerals with the same structure and forming

classification of minerals (silicates)
].

Some ion groups with a similar radius can occupy the same structural site in the crystal cell:

More recent definitions:

Differences between minerals and rocks

A mineral is a naturally occurring solid with a definite chemical composition and a specific crystalline structure. A

kaolin, calcite, epidote, olivine, augite, hornblende, magnetite, hematite, limonite and a few other minerals.[24]
Over half of the mineral species known are so rare that they have only been found in a handful of samples, and many are known from only one or two small grains.

Commercially valuable minerals and rocks are referred to as

ores (the rocks and minerals that remain, after the desired mineral has been separated from the ore, are referred to as tailings
).

Mineral composition of rocks

A main determining factor in the formation of minerals in a rock mass is the chemical composition of the mass, for a certain mineral can be formed only when the necessary elements are present in the rock. Calcite is most common in

silica
.

Other factors are of equal importance in determining the natural association or paragenesis of rock-forming minerals, principally the mode of origin of the rock and the stages through which it has passed in attaining its present condition. Two rock masses may have very much the same bulk composition and yet consist of entirely different assemblages of minerals. The tendency is always for those compounds to be formed which are stable under the conditions under which the rock mass originated. A granite arises by the consolidation of a molten magma at high temperatures and great pressures and its component minerals are those stable under such conditions. Exposed to moisture, carbonic acid and other subaerial agents at the ordinary temperatures of the Earth's surface, some of these original minerals, such as quartz and white mica are relatively stable and remain unaffected; others weather or decay and are replaced by new combinations. The feldspar passes into kaolinite, muscovite and quartz, and any mafic minerals such as pyroxenes, amphiboles or biotite have been present they are often altered to chlorite, epidote, rutile and other substances. These changes are accompanied by disintegration, and the rock falls into a loose, incoherent, earthy mass which may be regarded as a sand or soil. The materials thus formed may be washed away and deposited as sandstone or siltstone. The structure of the original rock is now replaced by a new one; the mineralogical constitution is profoundly altered; but the bulk chemical composition may not be very different. The sedimentary rock may again undergo metamorphism. If penetrated by igneous rocks it may be recrystallized or, if subjected to enormous pressures with heat and movement during mountain building, it may be converted into a gneiss not very different in mineralogical composition though radically different in structure to the granite which was its original state.[24]

Physical properties of minerals

Classifying minerals can range from simple to very 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.

Commonly used physical properties include crystal structure and habit, hardness, lustre, diapheniety, colour, streak, cleavage and fracture, and specific gravity. Other less general tests include

radioactivity, tenacity (response to mechanical induced changes of shape or form), piezoelectricity and reactivity to dilute acids.[1]

Crystal structure and habit

Acicular natrolite

A mineral may show good crystal habit or form, or it may be massive, granular or compact with only microscopically visible crystals.

A

crystal lattice arrangements of atoms in three dimensions, and these are referred to as the 14 "Bravais lattices". Each of these lattices can be classified into one of the seven crystal systems, and all crystal structures currently recognized fit in one Bravais lattice and one crystal system. 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. Chemistry and crystal structure together define a mineral. In fact, two or more minerals may have the same chemical composition, but differ in crystal structure (these are known as polymorphs). For example, pyrite and marcasite are both iron sulfide, but their arrangement of atoms differs. Similarly, some minerals have different chemical compositions, but the same crystal structure: for example, halite (made from sodium and chlorine), galena (made from lead and sulfur) and periclase (made from magnesium and oxygen) all share the same cubic crystal structure.

Crystal structure greatly influences a mineral's physical properties. For example, though diamond and graphite have the same composition (both are pure carbon), graphite is very soft, while diamond is the hardest of all known minerals. This happens because the carbon atoms in graphite are arranged into sheets which can slide easily past each other, while the carbon atoms in diamond form a strong, interlocking three-dimensional network.

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 render some directions softer than others.[25] An example of this property exists inf kyanite, which has a Mohs hardness of 5½ parallel to [001] but 7 parallel to [100].[26]

The most common scale of measurement is the ordinal Mohs hardness scale. Defined by 10 indicators, a mineral with a higher index will scratch those below it. The scale ranges from talc, a

phyllosilicate, to diamond, a carbon polymorph which is the hardness natural material. The scale is provided below:[25]


Mohs hardness Mineral Chemical formula Image
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 Feldspar
KAlSi3O8
7 Quartz SiO2
8 Topaz Al2SiO4(OH,F)2
9 Corundum Al2O3
10 Diamond C

Lustre

Pyrite has a metallic lustre.

Lustre indicates how light reflects from the mineral's surface, with regards to its quality and intensity. There are numerous qualitative terms used to describe this property:[27]

  • Metallic – high reflectivity like metal: galena and pyrite
  • Sub-metallic – slightly less than metallic reflectivity: magnetite
  • Non-metallic lusters:
    • Adamantine – brilliant, the luster of diamond also cerussite and anglesite
    • Vitreous – the luster of a broken glass: quartz
    • Pearly – iridescent and pearl-like: talc and apophyllite
    • Resinous – the luster of resin: sphalerite and sulfur
    • Silky – a soft light shown by fibrous materials: gypsum and chrysotile
    • Dull/earthy – shown by finely crystallized minerals: the kidney ore variety of hematite

Diaphaneity

Diaphaneity
describes how well light passes through a mineral; there are three basic degrees of transparency:

  • Transparent objects can be seen through a transparent mineral, such as a clear quartz crystal
  • Translucent light passes through the mineral but no objects can be seen
  • Opaque no light passes through the mineral
Many minerals range from transparent to translucent or translucent to opaque. Calcite, for instance, can be translucent or opaque. Some minerals that are naturally translucent become opaque with weathering.

Colour and streak

Watermelon elbaite

Colour indicates the appearance of the mineral in reflected light or transmitted light for translucent minerals (i.e. what it looks like to the naked eye).

  • Iridescence – the play of colors due to surface or internal interference. Labradorite exhibits internal iridescence whereas hematite and sphalerite often show the surface effect.
Streak plates with pyrite (left) and rhodochrosite (right)

Streak refers to the color of the powder a mineral leaves after rubbing it on an unglazed porcelain streak plate. Note that this is not always the same color as the original mineral.

Cleavage and fracture

Cleavage describes the way a mineral may split apart along various planes. In thin sections, cleavage is visible as thin parallel lines across a mineral. Fracture describes how a mineral breaks when broken contrary to its natural cleavage planes.

  • Chonchoidal fracture is a smooth curved fracture with concentric ridges of the type shown by glass.
  • Hackley is jagged fracture with sharp edges.
  • Fibrous
  • Irregular

Specific gravity

Specific gravity relates the mineral mass to the mass of an equal volume of water, namely the density of the material. While most minerals, including all the common rock-forming minerals, have a specific gravity of 2.5–3.5, a few are noticeably more or less dense, e.g. several sulfide minerals have high specific gravity compared to the common rock-forming minerals.

Classification of minerals

There are two major systems of mineral classification: the

Strunz classification
. The following list provides to the classes of minerals according to the Strunz classification.

Elements

The elemental group includes

meteorites
).

Sulfides

Many

sulfosalts
(sulfur and a second anion such as arsenic).

Halides

Halite

The

sal ammoniac (ammonium chloride). Halides, like sulfates, are commonly found in evaporite settings such as salt lakes and landlocked seas such as the Dead Sea and Great Salt Lake. The halide class includes the fluoride, chloride, bromide and iodide
minerals.

Oxides and hydroxides

oxidation products of other minerals in the near surface weathering zone, and as accessory minerals in igneous rocks of the crust and mantle. Common oxides include hematite (iron oxide), magnetite (iron oxide), chromite (iron chromium oxide), spinel (magnesium aluminium oxide – a common component of the mantle), ilmenite (iron titanium oxide), rutile (titanium dioxide), and ice (hydrogen oxide). The oxide class includes the oxide and the hydroxide
minerals.

Carbonates and nitrates

The

carbonate minerals consist of those minerals containing the anion (CO3)2− and include calcite and aragonite (both calcium carbonate), dolomite (magnesium/calcium carbonate) and siderite (iron carbonate). Carbonates are commonly deposited in marine settings when the shells of dead planktonic life settle and accumulate on the sea floor. Carbonates are also found in evaporitic settings (e.g. the Great Salt Lake, Utah) and also in karst regions, where the dissolution and reprecipitation of carbonates leads to the formation of caves, stalactites and stalagmites. The carbonate class also includes the nitrate and borate minerals
.

Borates

Sulfates

Hanksite, Na22K(SO4)9(CO3)2Cl, one of the few minerals that is considered a carbonate and a sulfate

barite (barium sulfate), and gypsum (hydrated calcium sulfate). The sulfate class also includes the chromate, molybdate, selenate, sulfite, tellurate, and tungstate
minerals.

Phosphates

The phosphate mineral group actually includes any mineral with a tetrahedral unit AO4 where A can be phosphorus, antimony, arsenic or vanadium. By far the most common phosphate is apatite which is an important biological mineral found in teeth and bones of many animals. The phosphate class includes the phosphate, arsenate, vanadate, and antimonate minerals.

Silicates

The largest group of minerals by far are the

silicates (most rocks are ≥95% silicates), which are composed largely of silicon and oxygen, with the addition of ions such as aluminium, magnesium, iron, and calcium. Some important rock-forming silicates include the feldspars, quartz, olivines, pyroxenes, amphiboles, garnets, and micas
.

Organic compounds

The organic mineral class includes

hydrocarbons and other miscellaneous species.[5]
Examples include whewellite, moolooite, mellite, fichtelite, carpathite, evenkite and abelsonite.

See also

Bibliography

  • Busbey, A.B.; Coenraads, R.E.; Roots, D.; Willis, P. (2007). Rocks and Fossils. San Francisco: Fog City Press.
    ISBN 978-1-74089_632-0. {{cite book}}: Check |isbn= value: invalid character (help); Cite has empty unknown parameter: |1= (help
    )
  • Dyar, M.D.; Gunter, M.E. (2008). Mineralogy and Optical Mineralogy. Chantilly, Virginia: .

References

  1. ^
    ISBN 0-471-80580-7. free older version: 1912 edition
  2. ^ a b c d Nickel, Ernest H. (1995). "The definition of a mineral". The Canadian Mineralogist. 33 (3): 689–690. alt version
  3. ^
    PMID 7008198
    .
  4. ^ .
  5. ^ a b Dana Classification 8th edition – Organic Compounds. Mindat.org. Retrieved on 2011-10-20.
  6. ^ Strunz Classification – Organic Compounds. Mindat.org. Retrieved on 2011-10-20.
  7. ^ .
  8. ^ IMA divisions. Ima-mineralogy.org (2011-01-12). Retrieved on 2011-10-20.
  9. ^ Working Group On Environmental Mineralogy (Wgem). Ima-mineralogy.org. Retrieved on 2011-10-20.
  10. ^ Takai, K. (2010). "Limits of life and the biosphere: Lessons from the detection of microorganisms in the deep sea and deep subsurface of the Earth.". In Gargaud, M.; Lopez-Garcia, P.; Martin, H. (eds.). Origins and Evolution of Life: An Astrobiological Perspective. Cambridge, UK: Cambridge University Press. pp. 469–486. {{cite book}}: External link in |chapterurl= (help); Unknown parameter |chapterurl= ignored (|chapter-url= suggested) (help)
  11. .
  12. .
  13. .
  14. .
  15. doi:10.1144/​SP336.3. {{cite journal}}: zero width space character in |doi= at position 9 (help
    )
  16. .
  17. ^ Official IMA list of mineral names (updated from March 2009 list). uws.edu.au
  18. ^ Bouligand, Y. (2006). "Liquid crystals and morphogenesis.". In Bourgine, P.; Lesne, A. (eds.). Morphogenesis: Origins of Patterns and Shape. Cambridge, UK: Springer Verlag. pp. 49-. {{cite book}}: External link in |chapterurl= (help); Unknown parameter |chapterurl= ignored (|chapter-url= suggested) (help)
  19. ^
    doi:10.1007/b10827. Cite error: The named reference "Gabriel03" was defined multiple times with different content (see the help page
    ).
  20. .
  21. ^ Hr. Dr. Udo Neumann der Uni-Tuebingen (Systematik der Minerale)
  22. ^ IMA Database of Mineral Properties/ RRUFF Project. Rruff.info. Retrieved on 2011-10-21.
  23. doi:10.1127/0935-1221/2009/0021-1994.{{cite journal}}: CS1 maint: multiple names: authors list (link
    )
  24. ^ a b Public Domain This article incorporates text from a publication now in the public domainChisholm, Hugh, ed. (1911). "Petrology". Encyclopædia Britannica (11th ed.). Cambridge University Press.
  25. ^ a b Dyar and Gunter, pp. 28–29
  26. ^ "Kyanite". Mindat.org. Retrieved 2012-08-01.
  27. ^ Dyar and Darby, pp. 26–28

External links

Template:Link GA Template:Link FA