Calcite

Last updated
Calcite
Calcite Variation.png
Clockwise from top left: scalenohedral, rhomboedral, stalactitic, and botryoidal calcite
General
Category Carbonate mineral
Formula
(repeating unit)
CaCO3
Strunz classification 5.AB.05
Crystal system Trigonal
Crystal class Hexagonal scalenohedral (3m)
H-M symbol: (3 2/m)
Space group R3c
Unit cell a = 4.9896(2)  Å,
c = 17.0610(11) Å; Z = 6
Identification
ColorTypically colorless or creamy white - may have shades of brownish colors
Crystal habit Botryoidal, concretionary, druse, globular, granular, massive, rhombohedral, scalenohedral, stalactitic
Twinning Common by four twin laws
Cleavage Perfect on {1011} three directions with angle of 74° 55' [1]
Fracture Conchoidal
Tenacity Brittle
Mohs scale hardness3 (defining mineral)
Luster Vitreous to pearly on cleavage surfaces
Streak White
Diaphaneity Transparent to translucent
Specific gravity 2.71
Optical propertiesUniaxial (−); low relief
Refractive index nω = 1.640–1.660
nε = 1.486
Birefringence δ = 0.154–0.174
Fusibility Infusible (decrepitates energetically) [2]
Solubility Soluble in dilute acids
Other characteristicsMay fluoresce red, blue, yellow, and other colors under either SW and LW UV; phosphorescent
References [3] [4] [5]

Calcite is a carbonate mineral and the most stable polymorph of calcium carbonate (CaCO3). It is a very common mineral, particularly as a component of limestone. Calcite defines hardness 3 on the Mohs scale of mineral hardness, based on scratch hardness comparison. Large calcite crystals are used in optical equipment, and limestone composed mostly of calcite has numerous uses.

Contents

Other polymorphs of calcium carbonate are the minerals aragonite and vaterite. Aragonite will change to calcite over timescales of days or less at temperatures exceeding 300 °C, [6] [7] and vaterite is even less stable.

Etymology

Calcite is derived from the German Calcit, a term from the 19th century that came from the Latin word for lime, calx (genitive calcis) with the suffix -ite used to name minerals. It is thus a doublet of the word chalk . [8]

When applied by archaeologists and stone trade professionals, the term alabaster is used not just as in geology and mineralogy, where it is reserved for a variety of gypsum; but also for a similar-looking, translucent variety of fine-grained banded deposit of calcite. [9]

Unit cell and Miller indices

Crystal structure of calcite Calcite.png
Crystal structure of calcite

In publications, two different sets of Miller indices are used to describe directions in hexagonal and rhombohedral crystals, including calcite crystals: three Miller indices h, k, l in the directions, or four Bravais-Miller indices h, k, i, l in the directions, where is redundant but useful in visualizing permutation symmetries.

To add to the complications, there are also two definitions of unit cell for calcite. One, an older "morphological" unit cell, was inferred by measuring angles between faces of crystals, typically with a goniometer, and looking for the smallest numbers that fit. Later, a "structural" unit cell was determined using X-ray crystallography. The morphological unit cell is rhombohedral, having approximate dimensions a = 10 Å and c = 8.5 Å , while the structural unit cell is hexagonal (i.e. a rhombic prism), having approximate dimensions a = 5 Å and c = 17 Å . For the same orientation, c must be multiplied by 4 to convert from morphological to structural units. As an example, calcite cleavage is given as "perfect on {1 0 1 1}" in morphological coordinates and "perfect on {1 0 1 4}" in structural units. In indices, these are {1 0 1} and {1 0 4}, respectively. Twinning, cleavage and crystal forms are often given in morphological units. [4] [10]

Properties

The diagnostic properties of calcite include a defining Mohs hardness of 3, a specific gravity of 2.71 and, in crystalline varieties, a vitreous luster. Color is white or none, though shades of gray, red, orange, yellow, green, blue, violet, brown, or even black can occur when the mineral is charged with impurities. [4]

Crystal habits

Calcite has numerous habits, representing combinations of over 1000 crystallographic forms. [3] Most common are scalenohedra, with faces in the hexagonal {2 1 1} directions (morphological unit cell) or {2 1 4} directions (structural unit cell); and rhombohedral, with faces in the {1 0 1} or {1 0 4} directions (the most common cleavage plane). [10] Habits include acute to obtuse rhombohedra, tabular habits, prisms, or various scalenohedra. Calcite exhibits several twinning types that add to the observed habits. It may occur as fibrous, granular, lamellar, or compact. A fibrous, efflorescent habit is known as lublinite. [11] Cleavage is usually in three directions parallel to the rhombohedron form. Its fracture is conchoidal, but difficult to obtain.

Scalenohedral faces are chiral and come in pairs with mirror-image symmetry; their growth can be influenced by interaction with chiral biomolecules such as L- and D-amino acids. Rhombohedral faces are not chiral. [10] [12]

Optical

Photograph of calcite displaying the characteristic birefringence optical behaviour Calcite-refraction-property.jpg
Photograph of calcite displaying the characteristic birefringence optical behaviour

Calcite is transparent to opaque and may occasionally show phosphorescence or fluorescence. A transparent variety called "Iceland spar" is used for optical purposes. [13] Acute scalenohedral crystals are sometimes referred to as "dogtooth spar" while the rhombohedral form is sometimes referred to as "nailhead spar". [2] The rhombohedral form may also have been the "sunstone" whose use by Viking navigators is mentioned in the Icelandic Sagas. [14]

Single calcite crystals display an optical property called birefringence (double refraction). This strong birefringence causes objects viewed through a clear piece of calcite to appear doubled. The birefringent effect (using calcite) was first described by the Danish scientist Rasmus Bartholin in 1669. At a wavelength of about 590 nm, calcite has ordinary and extraordinary refractive indices of 1.658 and 1.486, respectively. [15] Between 190 and 1700 nm, the ordinary refractive index varies roughly between 1.9 and 1.5, while the extraordinary refractive index varies between 1.6 and 1.4. [16]

Thermoluminescence

Demonstration of birefringence in calcite, using 445 nm laser Fluorescence in calcite.jpg
Demonstration of birefringence in calcite, using 445 nm laser

Calcite has thermoluminescent properties mainly due to manganese divalent (Mn2+). [17] An experiment was conducted by adding activators such as ions of Mn, Fe, Co, Ni, Cu, Zn, Ag, Pb, and Bi to the calcite samples to observe if they emitted heat or light. The results showed that adding ions (Cu+, Cu2+, Zn2+, Ag+, Bi3+, Fe2+, Fe3+, Co2+, Ni2+) did not react. [17] However, a reaction occurred when both manganese and lead ions were present in calcite. [17] By changing the temperature and observing the glow curve peaks, it was found that Pb2+and Mn2+acted as activators in the calcite lattice, but Pb2+ was much less efficient than Mn2+. [17]

Measuring mineral thermoluminescence experiments usually use x-rays or gamma-rays to activate the sample and record the changes in glowing curves at a temperature of 700-7500K. [17] Mineral thermoluminescence can form various glow curves of crystals under different conditions, such as temperature changes, because impurity ions or other crystal defects present in minerals supply luminescence centers and trapping levels. [17] Observing these curve changes also can help infer geological correlation and age determination. [17]

Chemical

Calcite, like most carbonates, dissolves in acids by the following reaction

CaCO3 + 2 H+ → Ca2+ + H2O + CO2

The carbon dioxide released by this reaction produces a characteristic effervescence when a calcite sample is treated with an acid.

Due to its acidity, carbon dioxide has a slight solubilizing effect on calcite. The overall reaction is

CaCO3(s) + H2O + CO2(aq) → Ca2+(aq) + 2HCO3(aq)

If the amount of dissolved carbon dioxide drops, the reaction reverses to precipitate calcite. As a result, calcite can be either dissolved by groundwater or precipitated by groundwater, depending on such factors as the water temperature, pH, and dissolved ion concentrations. When conditions are right for precipitation, calcite forms mineral coatings that cement rock grains together and can fill fractures. When conditions are right for dissolution, the removal of calcite can dramatically increase the porosity and permeability of the rock, and if it continues for a long period of time, may result in the formation of caves. Continued dissolution of calcium carbonate-rich formations can lead to the expansion and eventual collapse of cave systems, resulting in various forms of karst topography. [18]

Calcite exhibits an unusual characteristic called retrograde solubility: it is less soluble in water as the temperature increases. Calcite is also more soluble at higher pressures. [19]

Pure calcite has the composition CaCO3. However, the calcite in limestone often contains a few percent of magnesium. Calcite in limestone is divided into low-magnesium and high-magnesium calcite, with the dividing line placed at a composition of 4% magnesium. High-magnesium calcite retains the calcite mineral structure, which is distinct from that of dolomite, MgCa(CO3)2. [20] Calcite can also contain small quantities of iron and manganese. [21] Manganese may be responsible for the fluorescence of impure calcite, as may traces of organic compounds. [22]

Distribution

Calcite is found all over the world, and its leading global distribution is as follows:

The United States

Calcite Quarry, Michigan. Pd36-2-iss013e14843.jpg
Calcite Quarry, Michigan.

Calcite is found in many different areas in the United States. One of the best examples is the Calcite Quarry in Michigan. [23] The Calcite Quarry is the largest carbonate mine in the world and has been in use for more than 85 years. [23] Large quantities of calcite can be mined from these sizeable open pit mines.

Canada

Calcite can also be found throughout Canada, such as in Thorold Quarry and Madawaska Mine, Ontario, Canada. [24]

Mexico

Abundant calcite is mined in the Santa Eulalia mining district, Chihuahua, Mexico. [25]

Iceland

Large quantities of calcite in Iceland are concentrated in the Helgustadir mine. [26] The mine was once the primary mining location of "Iceland spar." [27] However, it currently serves as a nature reserve, and calcite mining will not be allowed. [27]

England

Calcite is found in parts of England, such as Alston Moor, Egremont, and Frizington, Cumbria. [26]

Germany

St. Andreasberg, Harz Mountains, and Freiberg, Saxony can find calcite. [26]

Use and applications

One of several calcite or alabaster perfume jars from the tomb of Tutankhamun, d. 1323 BC Tutankhamun's Alabaster Jar.jpg
One of several calcite or alabaster perfume jars from the tomb of Tutankhamun, d. 1323 BC

Ancient Egyptians carved many items out of calcite, relating it to their goddess Bast, whose name contributed to the term alabaster because of the close association. Many other cultures have used the material for similar carved objects and applications. [28]

A transparent variety of calcite known as Iceland spar may have been used by Vikings for navigating on cloudy days. A very pure crystal of calcite can split a beam of sunlight into dual images, as the polarized light deviates slightly from the main beam. By observing the sky through the crystal and then rotating it so that the two images are of equal brightness, the rings of polarized light that surround the sun can be seen even under overcast skies. Identifying the sun's location would give seafarers a reference point for navigating on their lengthy sea voyages. [29]

In World War II, high-grade optical calcite was used for gun sights, specifically in bomb sights and anti-aircraft weaponry. [30] It was used as a polarizer (in Nicol prisms) before the invention of Polaroid plates and still finds use in optical instruments. [31] Also, experiments have been conducted to use calcite for a cloak of invisibility. [32]

Microbiologically precipitated calcite has a wide range of applications, such as soil remediation, soil stabilization and concrete repair. [33] [34] It also can be used for tailings management and is designed to promote sustainable development in the mining industry. [35]

Calcite can help synthesize precipitated calcium carbonate (PCC) (mainly used in the paper industry) and increase carbonation. [36] Furthermore, due to its particular crystal habit, such as rhombohedron, hexagonal prism, etc., it promotes the production of PCC with specific shapes and particle sizes. [36]

Calcite, obtained from an 80 kg sample of Carrara marble, [37] is used as the IAEA-603 isotopic standard in mass spectrometry for the calibration of δ18O and δ13C. [38]

Calcite can be formed naturally or synthesized. However, artificial calcite is the preferred material to be used as a scaffold in bone tissue engineering due to its controllable and repeatable properties. [39]

Calcite can be used to alleviate water pollution caused by the excessive growth of cyanobacteria. Lakes and rivers can lead to cyanobacteria blooms due to eutrophication, which pollutes water resources. [40] Phosphorus (P) is the leading cause of excessive growth of cyanobacteria. [40] As an active capping material, calcite can help reduce P release from sediments into the water, thus inhibiting cyanobacteria overgrowth. [40]

Natural occurrence

Calcite is a common constituent of sedimentary rocks, limestone in particular, much of which is formed from the shells of dead marine organisms. Approximately 10% of sedimentary rock is limestone. It is the primary mineral in metamorphic marble. It also occurs in deposits from hot springs as a vein mineral; in caverns as stalactites and stalagmites; and in volcanic or mantle-derived rocks such as carbonatites, kimberlites, or rarely in peridotites.

Calcite is often the primary constituent of the shells of marine organisms, such as plankton (such as coccoliths and planktic foraminifera), the hard parts of red algae, some sponges, brachiopods, echinoderms, some serpulids, most bryozoa, and parts of the shells of some bivalves (such as oysters and rudists). Calcite is found in spectacular form in the Snowy River Cave of New Mexico as mentioned above, where microorganisms are credited with natural formations. Trilobites, which became extinct a quarter billion years ago, had unique compound eyes that used clear calcite crystals to form the lenses. [41] It also forms a substantial part of birds' eggshells, and the δ13C of the diet is reflected in the δ13C of the calcite of the shell. [42]

The largest documented single crystal of calcite originated from Iceland, measured 7 m × 7 m × 2 m (23 ft × 23 ft × 6.6 ft) and 6 m × 6 m × 3 m (20 ft × 20 ft × 9.8 ft) and weighed about 250 tons. [43] Classic samples have been produced at Madawaska Mine, near Bancroft, Ontario. [44]

Bedding parallel veins of fibrous calcite, often referred to in quarrying parlance as beef, occur in dark organic rich mudstones and shales, these veins are formed by increasing fluid pressure during diagenesis. [45]

Formation processes

Calcite formation can proceed by several pathways, from the classical terrace ledge kink model [46] to the crystallization of poorly ordered precursor phases like amorphous calcium carbonate (ACC) via an Ostwald ripening process, or via the agglomeration of nanocrystals. [47]

The crystallization of ACC can occur in two stages. First, the ACC nanoparticles rapidly dehydrate and crystallize to form individual particles of vaterite. Second, the vaterite transforms to calcite via a dissolution and reprecipitation mechanism, with the reaction rate controlled by the surface area of a calcite crystal. [48] The second stage of the reaction is approximately 10 times slower.

However, crystallization of calcite has been observed to be dependent on the starting pH and concentration of magnesium in solution. A neutral starting pH during mixing promotes the direct transformation of ACC into calcite without a vaterite intermediate. But when ACC forms in a solution with a basic initial pH, the transformation to calcite occurs via metastable vaterite, following the pathway outlined above. [48] Magnesium has a noteworthy effect on both the stability of ACC and its transformation to crystalline CaCO3, resulting in the formation of calcite directly from ACC, as this ion destabilizes the structure of vaterite.

Epitaxial overgrowths of calcite precipitated on weathered cleavage surfaces have morphologies that vary with the type of weathering the substrate experienced: growth on physically weathered surfaces has a shingled morphology due to Volmer-Weber growth, growth on chemically weathered surfaces has characteristics of Stranski-Krastanov growth, and growth on pristine cleavage surfaces has characteristics of Frank - van der Merwe growth. [49] These differences are apparently due to the influence of surface roughness on layer coalescence dynamics.

Calcite may form in the subsurface in response to microorganism activity, such as sulfate-dependent anaerobic oxidation of methane, where methane is oxidized and sulfate is reduced, leading to precipitation of calcite and pyrite from the produced bicarbonate and sulfide. These processes can be traced by the specific carbon isotope composition of the calcites, which are extremely depleted in the 13C isotope, by as much as −125 per mil PDB13C). [50]


In Earth history

Calcite seas existed in Earth's history when the primary inorganic precipitate of calcium carbonate in marine waters was low-magnesium calcite (lmc), as opposed to the aragonite and high-magnesium calcite (hmc) precipitated today. Calcite seas alternated with aragonite seas over the Phanerozoic, being most prominent in the Ordovician and Jurassic periods. Lineages evolved to use whichever morph of calcium carbonate was favourable in the ocean at the time they became mineralised, and retained this mineralogy for the remainder of their evolutionary history. [51] Petrographic evidence for these calcite sea conditions consists of calcitic ooids, lmc cements, hardgrounds, and rapid early seafloor aragonite dissolution. [52] The evolution of marine organisms with calcium carbonate shells may have been affected by the calcite and aragonite sea cycle. [53]

Calcite is one of the minerals that has been shown to catalyze an important biological reaction, the formose reaction, and may have had a role in the origin of life. [10] Interaction of its chiral surfaces (see Form) with aspartic acid molecules results in a slight bias in chirality; this is one possible mechanism for the origin of homochirality in living cells. [54]

Climate change

Ocean acidification reduces pH, which affects calcification in shelled organisms. Effect of Ocean Acidification on Calcification.png
Ocean acidification reduces pH, which affects calcification in shelled organisms.

Climate change is exacerbating ocean acidification, which may lead to lower natural calcite production. The oceans absorb large amounts of CO2 from fossil fuels emission into the air. [55] The total amount of artificial CO2 absorbed by the oceans is calculated to be 118 ± 19 Gt C. [56] If a large amount of CO2 dissolves in the sea, it will cause the acidity of the seawater to increase, thereby affecting the pH value of the ocean. [55] Calcifying organisms in the sea, such as molluscs foraminifera, crustaceans, echinoderms and corals, are susceptible to pH changes. [55] Meanwhile, these calcifying organisms are also an essential source of calcite. As ocean acidification causes pH to drop, carbonate ion concentrations will decline, potentially reducing natural calcite production. [55]

See also

Related Research Articles

<span class="mw-page-title-main">Limestone</span> Type of sedimentary rock

Limestone is a type of carbonate sedimentary rock which is the main source of the material lime. It is composed mostly of the minerals calcite and aragonite, which are different crystal forms of CaCO3. Limestone forms when these minerals precipitate out of water containing dissolved calcium. This can take place through both biological and nonbiological processes, though biological processes, such as the accumulation of corals and shells in the sea, have likely been more important for the last 540 million years. Limestone often contains fossils which provide scientists with information on ancient environments and on the evolution of life.

<span class="mw-page-title-main">Calcium carbonate</span> Chemical compound

Calcium carbonate is a chemical compound with the chemical formula CaCO3. It is a common substance found in rocks as the minerals calcite and aragonite, most notably in chalk and limestone, eggshells, gastropod shells, shellfish skeletons and pearls. Materials containing much calcium carbonate or resembling it are described as calcareous. Calcium carbonate is the active ingredient in agricultural lime and is produced when calcium ions in hard water react with carbonate ions to form limescale. It has medical use as a calcium supplement or as an antacid, but excessive consumption can be hazardous and cause hypercalcemia and digestive issues.

<span class="mw-page-title-main">Dolomite (mineral)</span> Carbonate mineral - CaMg(CO₃)₂

Dolomite is an anhydrous carbonate mineral composed of calcium magnesium carbonate, ideally CaMg(CO3)2. The term is also used for a sedimentary carbonate rock composed mostly of the mineral dolomite (see Dolomite (rock)). An alternative name sometimes used for the dolomitic rock type is dolostone.

<span class="mw-page-title-main">Exoskeleton</span> External skeleton of an organism

An exoskeleton is a skeleton that is on the exterior of an animal to both support the body shape and protect the internal organs, in contrast to an internal endoskeleton which is enclosed underneath other soft tissues. Some large, hard and non-flexible protective exoskeletons are known as shell or armour.

<span class="mw-page-title-main">Aragonite</span> Calcium carbonate mineral

Aragonite is a carbonate mineral and one of the three most common naturally occurring crystal forms of calcium carbonate, the others being calcite and vaterite. It is formed by biological and physical processes, including precipitation from marine and freshwater environments.

<span class="mw-page-title-main">Speleothem</span> Structure formed in a cave by the deposition of minerals from water

A speleothem is a geological formation by mineral deposits that accumulate over time in natural caves. Speleothems most commonly form in calcareous caves due to carbonate dissolution reactions. They can take a variety of forms, depending on their depositional history and environment. Their chemical composition, gradual growth, and preservation in caves make them useful paleoclimatic proxies.

<span class="mw-page-title-main">Dolomite (rock)</span> Sedimentary carbonate rock that contains a high percentage of the mineral dolomite

Dolomite (also known as dolomite rock, dolostone or dolomitic rock) is a sedimentary carbonate rock that contains a high percentage of the mineral dolomite, CaMg(CO3)2. It occurs widely, often in association with limestone and evaporites, though it is less abundant than limestone and rare in Cenozoic rock beds (beds less than about 66 million years in age). The first geologist to distinguish dolomite from limestone was Déodat Gratet de Dolomieu; a French mineralogist and geologist whom it is named after. He recognized and described the distinct characteristics of dolomite in the late 18th century, differentiating it from limestone.

<span class="mw-page-title-main">Vaterite</span> Calcium carbonate mineral

Vaterite is a mineral, a polymorph of calcium carbonate (CaCO3). It was named after the German mineralogist Heinrich Vater. It is also known as mu-calcium carbonate (μ-CaCO3). Vaterite belongs to the hexagonal crystal system, whereas calcite is trigonal and aragonite is orthorhombic.

<span class="mw-page-title-main">Flowstone</span> Archaeological phenomenon

Flowstones are sheetlike deposits of calcite or other carbonate minerals, formed where water flows down the walls or along the floors of a cave. They are typically found in "solution caves", in limestone, where they are the most common speleothem. However, they may form in any type of cave where water enters that has picked up dissolved minerals. Flowstones are formed via the degassing of vadose percolation waters.

<span class="mw-page-title-main">Carbonate rock</span> Class of sedimentary rock

Carbonate rocks are a class of sedimentary rocks composed primarily of carbonate minerals. The two major types are limestone, which is composed of calcite or aragonite (different crystal forms of CaCO3), and dolomite rock (also known as dolostone), which is composed of mineral dolomite (CaMg(CO3)2). They are usually classified based on texture and grain size. Importantly, carbonate rocks can exist as metamorphic and igneous rocks, too. When recrystallized carbonate rocks are metamorphosed, marble is created. Rare igneous carbonate rocks even exist as intrusive carbonatites and, even rarer, there exists volcanic carbonate lava.

<span class="mw-page-title-main">Kutnohorite</span> Mineral of calcium manganese carbonate

Kutnohorite is a rare calcium manganese carbonate mineral with magnesium and iron that is a member of the dolomite group. It forms a series with dolomite, and with ankerite. The end member formula is CaMn2+(CO3)2, but Mg2+ and Fe2+ commonly substitute for Mn2+, with the manganese content varying from 38% to 84%, so the formula Ca(Mn2+,Mg,Fe2+)(CO3)2 better represents the species. It was named by Professor Bukowsky in 1901 after the type locality of Kutná Hora, Bohemia, in the Czech Republic. It was originally spelt "kutnahorite" but "kutnohorite" is the current IMA-approved spelling.

<span class="mw-page-title-main">Ikaite</span> Hexahydrated calcium carbonate mineral

Ikaite is the mineral name for the hexahydrate of calcium carbonate, CaCO3·6H2O. Ikaite tends to form very steep or spiky pyramidal crystals, often radially arranged, of varied sizes from thumbnail size aggregates to gigantic salient spurs. It is only found in a metastable state and decomposes rapidly by losing most of its water content once removed from near-freezing water. This "melting mineral" is more commonly known through its pseudomorphs.

<span class="mw-page-title-main">Calcite sea</span> Sea chemistry favouring low-magnesium calcite as the inorganic calcium carbonate precipitate

A calcite sea is a sea in which low-magnesium calcite is the primary inorganic marine calcium carbonate precipitate. An aragonite sea is the alternate seawater chemistry in which aragonite and high-magnesium calcite are the primary inorganic carbonate precipitates. The Early Paleozoic and the Middle to Late Mesozoic oceans were predominantly calcite seas, whereas the Middle Paleozoic through the Early Mesozoic and the Cenozoic are characterized by aragonite seas.

<span class="mw-page-title-main">Aragonite sea</span> Chemical conditions of the sea favouring aragonite deposition

An aragonite sea contains aragonite and high-magnesium calcite as the primary inorganic calcium carbonate precipitates. The chemical conditions of the seawater must be notably high in magnesium content relative to calcium for an aragonite sea to form. This is in contrast to a calcite sea in which seawater low in magnesium content relative to calcium favors the formation of low-magnesium calcite as the primary inorganic marine calcium carbonate precipitate.

<span class="mw-page-title-main">Shortite</span>

Shortite is a sodium-calcium carbonate mineral, with the chemical formula Na2Ca2(CO3)3. It was discovered by J. J. Fahey in well cuttings from the Green River Formation, Sweetwater County, Wyoming, US, and was named to honor Maxwell N. Short (1889–1952), Professor of Mineralogy, University of Arizona.

<span class="mw-page-title-main">Amorphous calcium carbonate</span>

Amorphous calcium carbonate (ACC) is the amorphous and least stable polymorph of calcium carbonate. ACC is extremely unstable under normal conditions and is found naturally in taxa as wide-ranging as sea urchins, corals, mollusks, and foraminifera. It is usually found as a monohydrate, holding the chemical formula CaCO3·H2O; however, it can also exist in a dehydrated state, CaCO3. ACC has been known to science for over 100 years when a non-diffraction pattern of calcium carbonate was discovered by Sturcke Herman, exhibiting its poorly-ordered nature.

<span class="mw-page-title-main">Shell growth in estuaries</span>

Shell growth in estuaries is an aspect of marine biology that has attracted a number of scientific research studies. Many groups of marine organisms produce calcified exoskeletons, commonly known as shells, hard calcium carbonate structures which the organisms rely on for various specialized structural and defensive purposes. The rate at which these shells form is greatly influenced by physical and chemical characteristics of the water in which these organisms live. Estuaries are dynamic habitats which expose their inhabitants to a wide array of rapidly changing physical conditions, exaggerating the differences in physical and chemical properties of the water.

<span class="mw-page-title-main">Microbiologically induced calcite precipitation</span>

Microbiologically induced calcium carbonate precipitation (MICP) is a bio-geochemical process that induces calcium carbonate precipitation within the soil matrix. Biomineralization in the form of calcium carbonate precipitation can be traced back to the Precambrian period. Calcium carbonate can be precipitated in three polymorphic forms, which in the order of their usual stabilities are calcite, aragonite and vaterite. The main groups of microorganisms that can induce the carbonate precipitation are photosynthetic microorganisms such as cyanobacteria and microalgae; sulfate-reducing bacteria; and some species of microorganisms involved in nitrogen cycle. Several mechanisms have been identified by which bacteria can induce the calcium carbonate precipitation, including urea hydrolysis, denitrification, sulfate production, and iron reduction. Two different pathways, or autotrophic and heterotrophic pathways, through which calcium carbonate is produced have been identified. There are three autotrophic pathways, which all result in depletion of carbon dioxide and favouring calcium carbonate precipitation. In heterotrophic pathway, two metabolic cycles can be involved: the nitrogen cycle and the sulfur cycle. Several applications of this process have been proposed, such as remediation of cracks and corrosion prevention in concrete, biogrout, sequestration of radionuclides and heavy metals.

<span class="mw-page-title-main">Marine biogenic calcification</span> Shell formation mechanism

Marine biogenic calcification refers to the production of calcium carbonate by organisms in the global ocean.

<span class="mw-page-title-main">Particulate inorganic carbon</span>

Particulate inorganic carbon (PIC) can be contrasted with dissolved inorganic carbon (DIC), the other form of inorganic carbon found in the ocean. These distinctions are important in chemical oceanography. Particulate inorganic carbon is sometimes called suspended inorganic carbon. In operational terms, it is defined as the inorganic carbon in particulate form that is too large to pass through the filter used to separate dissolved inorganic carbon.

References

  1. Klein, Cornelis; Hurlbut, Cornelius S. Jr. (1993). Manual of mineralogy: (after James D. Dana) (21st ed.). New York: Wiley. p. 405. ISBN   047157452X.
  2. 1 2 Sinkankas, John (1964). Mineralogy for amateurs. Princeton, N.J.: Van Nostrand. pp. 359–364. ISBN   0442276249.
  3. 1 2 Anthony, John W.; Bideaux, Richard A.; Bladh, Kenneth W.; Nichols, Monte C., eds. (2003). "Calcite" (PDF). Handbook of Mineralogy. Vol. V (Borates, Carbonates, Sulfates). Chantilly, VA, US: Mineralogical Society of America. ISBN   978-0962209741.
  4. 1 2 3 "Calcite". mindat.org. Retrieved 1 Nov 2021.
  5. Barthelmy, Dave. "Calcite Mineral Data". webmineral.com. Retrieved 6 May 2018.
  6. Yoshioka S.; Kitano Y. (1985). "Transformation of aragonite to calcite through heating". Geochemical Journal. 19 (4): 24–249. Bibcode:1985GeocJ..19..245Y. doi: 10.2343/geochemj.19.245 .
  7. Staudigel P. T.; Swart P. K. (2016). "Isotopic behavior during the aragonite-calcite transition: Implications for sample preparation and proxy interpretation". Chemical Geology. 442: 130–138. Bibcode:2016ChGeo.442..130S. doi:10.1016/j.chemgeo.2016.09.013.
  8. "calcite (n.)". Online Etymology Dictionary. Retrieved 6 May 2018.
  9. More about alabaster and travertine, brief guide explaining the different use of the same terms by geologists, archaeologists, and the stone trade. Oxford University Museum of Natural History, 2012
  10. 1 2 3 4 Hazen, Robert M. (2004). "Chiral crystal faces of common rock-forming minerals". In Palyi, C.; Zucchi, C.; Caglioti, L. (eds.). Progress in Biological Chirality . Oxford: Elsevier. pp.  137–151. ISBN   9780080443966.
  11. "Lublinite". mindat.org. Retrieved 6 May 2018.
  12. Jiang, Wenge; Pacella, Michael S.; Athanasiadou, Dimitra; Nelea, Valentin; Vali, Hojatollah; Hazen, Robert M.; Gray, Jeffrey J.; McKee, Marc D. (2017-04-13). "Chiral acidic amino acids induce chiral hierarchical structure in calcium carbonate". Nature Communications. 8 (1): 15066. Bibcode:2017NatCo...815066J. doi:10.1038/ncomms15066. ISSN   2041-1723. PMC   5399303 . PMID   28406143.
  13. Harstad, A. O.; Stipp, S. L. S. (2007). "Calcite dissolution; effects of trace cations naturally present in Iceland spar calcites". Geochimica et Cosmochimica Acta. 71 (1): 56–70. Bibcode:2007GeCoA..71...56H. doi:10.1016/j.gca.2006.07.037.
  14. Ropars, Guy; Lakshminarayanan, Vasudevan; Le Floch, Albert (2 October 2014). "The sunstone and polarised skylight: ancient Viking navigational tools?". Contemporary Physics . 55 (4): 302–317. Bibcode:2014ConPh..55..302R. doi:10.1080/00107514.2014.929797. S2CID   119962347.
  15. Elert, Glenn (2021). "Refraction". The Physics Hypertextbook.
  16. Thompson, D. W.; Devries, M. J.; Tiwald, T. E.; Woollam, J. A. (1998). "Determination of optical anisotropy in calcite from ultraviolet to mid-infrared by generalized ellipsometry". Thin Solid Films. 313–314 (1–2): 341–346. Bibcode:1998TSF...313..341T. doi:10.1016/S0040-6090(97)00843-2.
  17. 1 2 3 4 5 6 7 Medlin, W. L. (1959). "Thermoluminescent properties of calcite". The Journal of Chemical Physics. 30 (2): 451–458. Bibcode:1959JChPh..30..451M. doi:10.1063/1.1729973.
  18. Wolfgang, Dreybrodt (2004). "Dissolution: Carbonate rocks". Encyclopedia of Caves and Karst Science. pp. 295–298. Retrieved 26 December 2020.
  19. Sharp, W. E.; Kennedy, G. C. (March 1965). "The System CaO-CO 2 -H 2 O in the Two-Phase Region Calcite + Aqueous Solution". The Journal of Geology. 73 (2): 391–403. doi:10.1086/627069. S2CID   100971186.
  20. Blatt, Harvey; Middleton, Gerard; Murray, Raymond (1980). Origin of sedimentary rocks (2d ed.). Englewood Cliffs, N.J.: Prentice-Hall. pp. 448–449. ISBN   0136427103.}
  21. Dromgoole, Edward L.; Walter, Lynn M. (February 1990). "Iron and manganese incorporation into calcite: Effects of growth kinetics, temperature and solution chemistry". Chemical Geology. 81 (4): 311–336. Bibcode:1990ChGeo..81..311D. doi:10.1016/0009-2541(90)90053-A.
  22. Pedone, Vicki A.; Cercone, Karen Rose; Burruss, R.C. (October 1990). "Activators of photoluminescence in calcite: evidence from high-resolution, laser-excited luminescence spectroscopy". Chemical Geology. 88 (1–2): 183–190. Bibcode:1990ChGeo..88..183P. doi:10.1016/0009-2541(90)90112-K.
  23. 1 2 "Calcite Quarry, Michigan". Earth observatory. 7 August 2006. Retrieved 17 February 2023.
  24. Hudson Institute of Mineralogy. "Calcite from Canada". mindat.org. Retrieved 17 February 2023.
  25. Hudson Institute of Mineralogy. "Santa Eulalia Mining District, Aquiles Serdán Municipality, Chihuahua, Mexico". mindat.org. Retrieved 17 February 2023.
  26. 1 2 3 AZoMining (15 October 2013). "Calcite – occurrence, properties, and distribution". azomining.com. Retrieved 17 February 2023.
  27. 1 2 Kristjansson, L. (2002). "Iceland Spar: The Helgustadir Calcite Locality and its Influence on the Development of Science". Journal of Geoscience Education. 50 (4): 419–427. Bibcode:2002JGeEd..50..419K. doi:10.5408/1089-9995-50.4.419. S2CID   126987943.
  28. Reed, Kristina (Spring 2017). "Display Case" (PDF). La Sierra Digs. Vol. 5, no. 2. La Sierra University. Retrieved 6 February 2021.
  29. Perkins, Sid (3 April 2018). "Viking seafarers may have navigated with legendary crystals". Science. doi: 10.1126/science.aat7802 .
  30. Lister, Priscilla (December 5, 2010). "Borrego's calcite mine trail holds desert wonders". The San Diego Union-Tribune. Retrieved January 8, 2021.
  31. Klein & Hurlbut 1993, p. 408.
  32. Chen, Xianzhong; Luo, Yu; Zhang, Jingjing; Jiang, Kyle; Pendry, John B.; Zhang, Shuang (2011). "Macroscopic invisibility cloaking of visible light". Nature Communications. 2 (2): 176. arXiv: 1012.2783 . Bibcode:2011NatCo...2..176C. doi:10.1038/ncomms1176. PMC   3105339 . PMID   21285954.
  33. Mujah, D.; Shahin, M. A.; Cheng, L. (2017). "State-of-the-Art Review of Biocementation by Microbially Induced Calcite Precipitation (MICP) for Soil Stabilization". Geomicrobiology Journal. 34 (6): 524–537. Bibcode:2017GmbJ...34..524M. doi:10.1080/01490451.2016.1225866. S2CID   88584080.
  34. Castro-Alonso, M. J.; Montañez-Hernandez, L. E.; Sanchez-Muñoz, M. A.; Macias Franco, M. R.; Narayanasamy, R.; Balagurusamy, N. (2019). "Microbially induced calcium carbonate precipitation (MICP) and its potential in bioconcrete: microbiological and molecular concepts". Frontiers in Materials. 6: 126. Bibcode:2019FrMat...6..126C. doi: 10.3389/fmats.2019.00126 .
  35. Zúñiga-Barra, H.; Toledo-Alarcón, J.; Torres-Aravena, Á.; Jorquera, L.; Rivas, M.; Gutiérrez, L.; Jeison, D. (2022). "Improving the sustainable management of mining tailings through microbially induced calcite precipitation: A review". Minerals Engineering. 189: 107855–. doi:10.1016/j.mineng.2022.107855. S2CID   252986388.
  36. 1 2 Jimoh, O. A.; Ariffin, K. S.; Hussin, H. B.; Temitope, A. E. (2018). "Synthesis of precipitated calcium carbonate: a review". Carbonates and Evaporites. 33 (2): 331–346. doi:10.1007/s13146-017-0341-x. S2CID   133034902.
  37. Department of Nuclear Sciences and Applications, IAEA Environment Laboratories (16 July 2016). "Reference Sheet: Certified Reference Material: IAEA-603 (calcite) – Stable Isotope Reference Material for δ13C and δ18O" (PDF). IAEA. p. 2. Retrieved 28 February 2017.
  38. "IAEA-603, Calcite". Reference Products for Environment and Trade. International Atomic Energy Agency. Retrieved 27 February 2017.
  39. Chróścicka, A.; Jaegermann, Z.; Wychowański, P.; Ratajska, A.; Sadło, J.; Hoser, G.; Michałowski, S.; Lewandowska-Szumiel, M. (2016). "Synthetic Calcite as a Scaffold for Osteoinductive Bone Substitutes". Annals of Biomedical Engineering. 44 (7): 2145–2157. doi:10.1007/s10439-015-1520-3. PMC   4893069 . PMID   26666226.
  40. 1 2 3 Han, M.; Wang, Y.; Zhan, Y.; Lin, J.; Bai, X.; Zhang, Z. (2022). "Efficiency and mechanism for the control of phosphorus release from sediment by the combined use of hydrous ferric oxide, calcite and zeolite as a geo-engineering tool". Chemical Engineering Journal. 428: 131360–. doi:10.1016/j.cej.2021.131360.
  41. Angier, Natalie (3 March 2014). "When Trilobites Ruled the World". The New York Times . Retrieved 10 March 2014.
  42. Lynch, Amanda H.; Beringer, Jason; Kershaw, Peter; et al. (2007). "Using the Paleorecord to Evaluate Climate and Fire Interactions in Australia". Annual Review of Earth and Planetary Sciences . 35 (1): 215–239. Bibcode:2007AREPS..35..215L. doi:10.1146/annurev.earth.35.092006.145055.
  43. Rickwood, P. C. (1981). "The largest crystals" (PDF). American Mineralogist. 66: 885–907.
  44. McDougall, Raymond (2019-09-03). "Mineral Highlights from the Bancroft Area, Ontario, Canada". Rocks & Minerals. 94 (5): 408–419. Bibcode:2019RoMin..94..408M. doi:10.1080/00357529.2019.1619134. S2CID   201298402.
  45. Ravier, Edouard; Martinez, Mathieu; Pellenard, Pierre; et al. (December 2020). "The milankovitch fingerprint on the distribution and thickness of bedding-parallel veins (beef) in source rocks" (PDF). Marine and Petroleum Geology. 122: 104643. Bibcode:2020MarPG.12204643R. doi:10.1016/j.marpetgeo.2020.104643. S2CID   225177225.
  46. De Yoreo, J. J.; Vekilov, P. G. (2003). "Principles of crystal nucleation and growth". Reviews in Mineralogy and Geochemistry. 54 (1): 57–93. Bibcode:2003RvMG...54...57D. CiteSeerX   10.1.1.324.6362 . doi:10.2113/0540057.
  47. De Yoreo, J.; Gilbert, P.U.; Sommerdijk, N. A. J. M.; Penn, R. L.; Whitelam, S.; Joester, D.; Zhang, H.; Rimer, J. D.; Navrotsky, A.; Banfield, J. F.; Wallace, A. F.; Michel, F. M.; Meldrum, F. C.; Cölfen, H.; Dove, P. M. (2015). "Crystallization by particle attachment in synthetic, biogenic, and geologic environments" (PDF). Science. 349 (6247): aaa6760. doi:10.1126/science.aaa6760. PMID   26228157. S2CID   14742194.
  48. 1 2 Gebauer, Denis; Kellermeier, Matthias; Gale, Julian D.; Bergström, Lennart; Cölfen, Helmut (2014). "Pre-nucleation clusters as solute precursors in crystallisation". Chem. Soc. Rev. 43 (7): 2348–2371. doi:10.1039/C3CS60451A. hdl: 20.500.11937/6133 . PMID   24457316. S2CID   585569.
  49. Acosta, Marisa D.; Olsen, Ellen K.; Pickerel, Molly E. (2023-09-20). "Surface roughness and overgrowth dynamics: The effect of substrate micro-topography on calcite growth and Sr uptake". Chemical Geology. 634: 121585. doi: 10.1016/j.chemgeo.2023.121585 . ISSN   0009-2541.
  50. Drake, H.; Astrom, M. E.; Heim, C.; Broman, C.; Astrom, J.; Whitehouse, M.; Ivarsson, M.; Siljestrom, S.; Sjovall, P. (2015). "Extreme 13C depletion of carbonates formed during oxidation of biogenic methane in fractured granite". Nature Communications. 6: 7020. Bibcode:2015NatCo...6.7020D. doi:10.1038/ncomms8020. PMC   4432592 . PMID   25948095.
  51. Porter, S. M. (2007). "Seawater Chemistry and Early Carbonate Biomineralization". Science. 316 (5829): 1302. Bibcode:2007Sci...316.1302P. doi:10.1126/science.1137284. PMID   17540895. S2CID   27418253.
  52. Palmer, Timothy; Wilson, Mark (2004). "Calcite precipitation and dissolution of biogenic aragonite in shallow Ordovician calcite seas". Lethaia. 37 (4): 417–427. Bibcode:2004Letha..37..417P. doi:10.1080/00241160410002135.
  53. Harper, E.M.; Palmer, T.J.; Alphey, J.R. (1997). "Evolutionary response by bivalves to changing Phanerozoic sea-water chemistry". Geological Magazine. 134 (3): 403–407. Bibcode:1997GeoM..134..403H. doi:10.1017/S0016756897007061. S2CID   140646397.
  54. Meierhenrich, Uwe (2008). Amino acids and the asymmetry of life caught in the act of formation. Berlin: Springer. pp. 76–78. ISBN   9783540768869.
  55. 1 2 3 4 Tyrrell, T. (2008). "Calcium carbonate cycling in future oceans and its influence on future climates". Journal of Plankton Research. 30 (2): 141–156. doi:10.1093/plankt/fbm105.
  56. Sabine, Christopher L.; Feely, Richard A.; Gruber, Nicholas; Key, Robert M.; Lee, Kitack; Bullister, John L.; Wanninkhof, Rik; Wong, C. S.; Wallace, Douglas W. R.; Tilbrook, Bronte; Billero, Frank J.; Peng, Tsung-Hung; Kozyr, Alexander; Ono, Tsueno; Rios, A. F. (2004). "The oceanic sink for anthropogenic CO2". Science. 305(5682) (5682): 367–371. Bibcode:2004Sci...305..367S. doi:10.1126/science.1097403. hdl: 10261/52596 . PMID   15256665. S2CID   5607281.

Further reading