mineral deposit

mineral deposit

Introduction

      aggregate of a mineral in an unusually high concentration.

      About half of the known chemical elements (chemical element) possess some metallic (metal) properties. The term metal, however, is reserved for those chemical elements that possess two or more of the characteristic physical properties of metals (opacity, ductility, malleability, fusibility) and are also good conductors of heat and electricity. Approximately 40 metals are made available through the mining and smelting of the minerals (mineral) in which they occur.

      Certain kinds of mineral can be smelted more readily than others; these are commonly referred to as ore minerals. Ore minerals tend to be concentrated in small, localized rock masses that form as a result of special geologic processes, and such local concentrations are called mineral deposits. Mineral deposits are what prospectors seek. The terms ore mineral and mineral deposit were originally applied only to minerals and deposits from which metals are recovered, but present usage includes a few nonmetallic minerals, such as barite and fluorite, that are found in the same kinds of deposit as metallic minerals.

      No deposit consists entirely of a single ore mineral. There are always admixtures of valueless minerals, collectively called gangue. The more concentrated an ore mineral, the more valuable the mineral deposit. For every mineral deposit there is a set of conditions, such as the level of concentration and the size of the deposit, that must be reached if the deposit is to be worked at a profit. A mineral deposit that is sufficiently rich to be worked at a profit is called an ore deposit, and in an ore deposit the assemblage of ore minerals plus gangue is called the ore.

      All ore deposits are mineral deposits, but the reverse is not true. “Ore deposit” is an economic term, while “mineral deposit” is a geologic term. Whether a given mineral deposit is also an ore deposit depends on many factors other than the level of concentration and the size of the deposit; all factors that affect the mining, processing, and transporting of the ore must be considered as well. Among such factors are the shape of a deposit, its depth below the surface, its geographic remoteness, access to transportation, the political stability of the region, and market factors such as the price of the metal in world trade and the costs of borrowing the money needed to develop a mine. Because market factors change continually, a given mineral deposit may sometimes be an ore deposit, but at other times it may be uneconomic and hence not an ore deposit.

      Mineral deposits have been found both in rocks (rock) that lie beneath the oceans and in rocks that form the continents, although the only deposits that actually have been mined are in the continental rocks. (The mining of ocean deposits lies in the future.) The continental crust averages 35–40 kilometres (20–25 miles) in thickness, and below the crust lies the mantle. Mineral deposits may occur in the mantle, but with present technology it is not possible to discover them.

Geochemically abundant and scarce metals
      Metals used in industrial and technological applications can be divided into two classes on the basis of their abundance in the Earth's crust. The geochemically abundant metals, of which there are five (aluminum, iron, magnesium, manganese, and titanium), constitute more than 0.1 percent by weight of the Earth's crust, while the geochemically scarce metals, which embrace all other metals (including such familiar ones as copper, lead, zinc, gold, and silver), constitute less than 0.1 percent. In almost every rock, at least tiny amounts of all metals can be detected by sensitive chemical analysis. However, there are important differences in the way the abundant and scarce metals occur in common rocks. Geochemically abundant metals tend to be present as essential constituents in minerals. For example, basalt, a common igneous rock, consists largely of the minerals olivine and pyroxene (both magnesium-iron silicates), feldspar (sodium-calcium-aluminum silicate), and ilmenite (iron-titanium oxide). Careful chemical analysis of a basalt will reveal the presence of most of the geochemically scarce metals too, but no amount of searching will reveal minerals in which one or more of the scarce metals is an essential constituent.

      Geochemically scarce metals rarely form minerals in common rocks. Instead, they are carried in the structures of common rock-forming minerals (most of them silicates) through the process of atomic substitution. This process involves the random replacement of an atom in a mineral by a foreign atom of similar ionic radius and valence, without changing the atomic packing of the host mineral. Atoms of copper, zinc, and nickel, for example, can substitute for iron and magnesium atoms in olivine and pyroxene. However, since substitution of foreign atoms produces strains in an atomic packing, there are limits to this process, as determined by temperature, pressure, and various chemical parameters. Indeed, the substitution limits for most scarce metals in common silicate minerals are low—in many cases only a few hundred substituting atoms for every million host atoms—but even these limits are rarely exceeded in common rocks.

      One important consequence that derives from the way abundant and scarce metals occur in common rocks is that ore minerals of abundant metals can be found in many common rocks, while ore minerals of scarce metals can be found only where some special, restricted geologic process has formed localized enrichments that exceed the limits of atomic substitution.

Ore minerals (beneficiation)
      Two factors determine whether a given mineral is suitable to be an ore mineral. The first is the ease with which a mineral can be separated from the gangue and concentrated for smelting. Concentrating processes, which are based on the physical properties of the mineral, include magnetic separation, gravity separation, and flotation. The second factor is smelting—that is, releasing the metal from the other elements to which it is chemically bonded in the mineral. Smelting processes are discussed below, but of primary importance in this consideration of the suitability of an ore mineral is the amount of energy needed to break the chemical bonds and release the metal. In general, less energy is needed to smelt sulfide, oxide, or hydroxide minerals than is required to smelt a silicate mineral. For this reason, few silicate minerals are ore minerals. Because the great bulk of the Earth's crust (about 95 percent) is composed of silicate minerals, sulfide, oxide, and hydroxide ore minerals are at best only minor constituents of the Earth's crust—and in many cases they are very rare constituents.

      The preferred ore minerals of both geochemically abundant and geochemically scarce metals are native metals, sulfides, oxides, hydroxides, or carbonates. In a few cases, silicate minerals have to be used as ore minerals because the metals either do not form more desirable minerals or form desirable minerals that rarely occur in large deposits.

Native metals (native element)
      Only two metals, gold (gold processing) and platinum, are found principally in their native state, and in both cases the native metals are the primary ore minerals. Silver, copper, iron, osmium, and several other metals also occur in the native state, and a few occurrences are large enough—and sufficiently rich—to be ore deposits. One example is the rich deposits of native copper (copper processing) in the Lake Superior area of Michigan in the United States. Here the copper occurs in interbedded conglomerates (conglomerate) (a sedimentary rock consisting of pebbles and boulders) and basaltic lava flows, most of which are vesicular and have fragmental layers of basaltic rubble on top of each flow. In the basalt, the native copper and associated gangue minerals fill the vesicles and cavities in the rubble; in the conglomerate, the copper fills spaces between the pebbles and in part replaces some of the smaller rock fragments. Originally discovered and worked by native Americans who manufactured and traded ornaments of malleable copper, the great Michigan copper deposits were first mined in 1845 and continued in production for more than a century.

Sulfides (sulfide mineral)
      The largest group of ore minerals consists of sulfides. Because the physical and chemical properties of telluride, selenide, and arsenide minerals are very similar to those of sulfide minerals, and because these minerals tend to occur together, it is convenient to use the term sulfide to embrace these similar minerals. Copper, lead, zinc, nickel, molybdenum, silver, arsenic, antimony, bismuth, cobalt, and mercury all form sulfide ore minerals. Gold and silver form tellurides under certain circumstances, and platinum forms an important arsenide ore mineral.

      The principle of atomic substitution operates in all classes of minerals, and some of the rarest metals occur by atomic substitution in sulfide ore minerals of other scarce metals. For example, cadmium and indium are generally present in small amounts in the zinc sulfide sphalerite, the major ore mineral of zinc. In fact, most of the world's cadmium and indium is recovered as a by-product of the smelting of sphalerite concentrates to produce zinc (zinc processing).

Oxides (oxide mineral) and hydroxides
      Oxides and hydroxides are a large and diverse group of ore minerals. The major ore minerals of the geochemically abundant metals aluminum, iron, manganese, and titanium are either oxides or hydroxides, while the oxide-forming scarce metals are chromium, tin, tungsten, tantalum, niobium, and uranium. Vanadium (vanadium processing) is found mainly by atomic substitution in magnetite, a major oxide ore mineral of iron.

Carbonates (carbonate mineral) and silicates (silicate mineral)
      Carbonate minerals are widespread in the Earth's crust, but only a few are ore minerals. These are the carbonates of iron, manganese, magnesium, and the rare earths. The number of metals won from silicate ore minerals is small. Most important are beryllium, zirconium, and lithium, plus a certain amount of nickel recovered from the nickel silicate garnierite.

Formation of mineral deposits
      Mineral deposits form because some medium serves as a concentrating and transporting agent for the ore minerals, and some process subsequently causes the transporting agent to precipitate, or deposit, the minerals. Examples of concentrating and transporting agents are groundwater, seawater, and magma; examples of precipitating processes are boiling (as in a hot spring), the cooling of a hot solution, the crystallization of a magma, and a chemical reaction between a solution and the rocks through which it flows. The same kinds of concentrating and transporting agent and the same kinds of precipitating process are involved in the formation of deposits of both geochemically abundant and geochemically scarce metals.

      There are six principal concentrating and transporting agents. Together with the classes of deposit that they form, these agents are discussed below.

Magmatic (magma) concentration
      Magma is molten rock, together with any suspended mineral grains and dissolved gases, that forms when temperatures rise and melting occurs in the mantle or crust. When magma rises to the Earth's surface through fissures and volcanic vents, it is called lava. Lava cools and crystallizes quickly, so that igneous rocks (igneous rock) formed from lava tend to consist of tiny mineral grains. (Sometimes cooling can be so rapid that mineral grains cannot form and a glass results.) Underground magma, on the other hand, cools and crystallizes slowly, and the resulting igneous rocks tend to contain mineral grains at least one-half centimetre (about one-quarter inch) in diameter.

pegmatite deposits
      The crystallization of magma is a complex process because magma is a complex substance. Certain magmas, such as those which form granites, contain several percent water dissolved in them. When a granitic magma cools, the first minerals to crystallize tend to be anhydrous (e.g., feldspar), so an increasingly water-rich residue remains. Certain rare chemical elements, such as lithium, beryllium, and niobium, that do not readily enter into atomic substitution in the main granite minerals (feldspar, quartz, and mica) become concentrated in the water-rich residual magma. If the crystallization process occurs at a depth of about five kilometres or greater, the water-rich residual magma may migrate and form small bodies of igneous rock, satellitic to the main granitic mass, that are enriched in rare elements. Such small igneous bodies, called rare-metal pegmatites, are sometimes exceedingly coarse-grained, with individual grains of mica, feldspar, and beryl up to one metre across. Pegmatites have been discovered on all continents, providing an important fraction of the world's lithium, beryllium, cesium, niobium, and tantalum. Pegmatites also are the major source of sheet mica and important sources of gemstones, particularly tourmalines and the gem forms of beryl (aquamarine and emerald).

Carbonatite deposits
      Carbonatites are igneous rocks that consist largely of the carbonate minerals calcite and dolomite; (dolomite) they sometimes also contain the rare-earth ore minerals bastnaesite, parisite, and monazite, the niobium ore mineral pyrochlore, and (in the case of the carbonatite deposit at Palabora in South Africa) copper sulfide ore minerals. The origin of carbonatite magma is obscure. Most carbonatites occur close to intrusions of alkaline igneous rocks (those rich in potassium or sodium relative to their silica contents) or to the ultramafic igneous rocks (rocks with silica contents below approximately 50 percent by weight) known as kimberlites and lamproites. These associations suggest a common derivation, but details of the way that carbonatite magmas might concentrate geochemically scarce metals remain conjectural.

      Carbonatites have been found on all continents; they also range widely in age, from deposits in the East African Rift Valley (East African Rift System) that were formed during the present geologic age to South African deposits dating from the early Proterozoic Eon (2.5 billion to 543 million years ago). Many carbonatites are mined or contain such large reserves that they will be mined someday. Among the most important are Mountain Pass, Calif., U.S., a major source of rare earths; the Loolekop Complex, Palabora, S.Af., mined for copper and apatite (calcium phosphate, used as a fertilizer), plus by-products of gold, silver, and other metals; Jacupiranga, Brazil, a major resource of rare earths; Oka, Que., Can., a niobium-rich body; and the Kola Peninsula of Russia, mined for apatite, magnetite, and rare earths.

Magmatic cumulates
      Magmatic segregation is a general term referring to any process by which one or more minerals become locally concentrated (segregated) during the cooling and crystallization of a magma. Rocks formed as a result of magmatic segregation are called magmatic cumulates. While a magma may start as a homogeneous liquid, magmatic segregation during crystallization can produce an assemblage of cumulates with widely differing compositions. Extreme segregation can sometimes produce monomineralic cumulates; a dramatic example occurs in the Bushveld Igneous Complex of South Africa, where cumulus layers of chromite (iron-magnesium-chromium oxide, the only chromium ore mineral) are encased in cumulus layers of anorthite (calcium-rich feldspar).

      Mineral deposits that are magmatic cumulates are only found in mafic and ultramafic igneous rocks (i.e., rocks that are low in silica). This is due to the control exerted by silica on the viscosity of a magma: the higher the silica content, the more viscous a magma and the more slowly segregation can proceed. Highly viscous magmas, such as those of granitic composition, tend to cool and crystallize faster than segregation can proceed. In low-silica (and, hence, low-viscosity) magmas such as gabbro, basalt, and komatiite, mineral grains can float, sink, or be moved so rapidly by flowing magma that segregation can occur before crystallization is complete.

      As with most geologic processes that cannot be directly observed, a certain amount of uncertainty exists about how cumulates form. A mineral such as chromite, with a density considerably greater than the magma from which it crystallizes, will tend to sink as soon as it forms. As a result, geologists long held the opinion that cumulates of chromite and other dense minerals formed only by sinking. This simple picture was challenged in 1961 by E. Dale Jackson, a geologist employed by the U.S. Geological Survey, who studied chromite cumulates of the Stillwater Complex in Montana. The findings of Jackson and later workers suggested that cumulates can also be produced by such phenomena as in-place crystallization of monomineralic layers on the floor of a magma chamber or density currents (density current) carrying mineral grains from the walls and roof of a magma chamber to the floor. Opinion still remains open, but most geologists now agree that in-place crystallization and density currents are more important in the formation of magmatic cumulates than density sinking.

      Three oxide ore minerals form magmatic cumulates: chromite, magnetite, and ilmenite. The world's largest chromite deposits are all magmatic cumulates; the largest and richest of these is in the Bushveld Complex of South Africa. Cumulus deposits of magnetite make poor iron ores, because cumulus magnetites invariably contain elements such as titanium, manganese, and vanadium by atomic substitution—although vanadiferous magnetites are important as a source of vanadium (vanadium processing). In fact, much of the world's production of this metal comes from cumulus magnetites in the Bushveld Complex.

Immiscible melts
      A different kind of magmatic segregation involves liquid immiscibility. A cooling magma will sometimes precipitate droplets of a second magma that has an entirely different composition. Like oil and water, the two magmas will not mix (i.e., they are immiscible). The chemical principle governing precipitation of an immiscible liquid is the same as that governing crystallization of a mineral from a magma: when the concentration of a particular mineral within a parent magma reaches saturation, precipitation occurs. If saturation is reached at a temperature above the melting point of the mineral, a drop of liquid precipitates instead of a mineral grain. The composition of this immiscible drop is not exactly that of the pure mineral, because the liquid tends to scavenge and concentrate many elements from the parent magma, and this process can lead to rich ore deposits.

      Iron sulfide is the principal constituent of most immiscible magmas, and the metals scavenged by iron sulfide liquid are copper, nickel, and the platinum group. Immiscible sulfide drops can become segregated and form immiscible magma layers in a magma chamber in the same way that cumulus layers form; then, when layers of sulfide magma cool and crystallize, the result is a deposit of ore minerals of copper, nickel, and platinum-group metals in a gangue of an iron sulfide mineral. Among the ore deposits of the world formed in this way are the Merensky Reef of the Bushveld Complex, producer of a major fraction of the world's platinum-group metals; the Stillwater Complex, Montana, host to platinum-group deposits similar to the Merensky Reef; and the Norilsk deposits of Russia, containing large reserves of platinum-group metals.

      Under suitable conditions, immiscible sulfide liquids can also become segregated from flowing lavas. An example is offered by the Kambalda nickel deposit in Western Australia. At Kambalda a nickel ore mineral, pentlandite, together with valuable by-product minerals of copper and platinum-group metals, crystallized in an iron-sulfide-rich gangue from a sulfide liquid that had become segregated from a magnesium-rich lava called a komatiite (named for the Komati River in South Africa).

      A group of unique deposits formed by immiscible sulfide liquids is the Sudbury Igneous Complex in Ontario, Can., which formed about 1.85 billion years ago. Elliptical in outline (approximately 60 kilometres long by 28 kilometres wide), the complex has the shape of a funnel pointing down into the Earth. A continuous lower zone of a mafic rock called norite lies above a discontinuous zone of gabbro, some of which contains numerous broken fragments of the underlying basement rocks and some of which is rich in sulfide ore minerals of nickel, copper, and platinum-group metals. Many hypotheses have been suggested for the origin of the Sudbury Complex. Some consider it to be an intrusive igneous complex, and some consider it a combination of intrusive and extrusive igneous rocks, but the most widely held opinion derives from the work of the American scientist Robert S. Dietz (Dietz, Robert S.), who in 1964 suggested that the Sudbury structure is an astrobleme, the site of a large meteorite impact. The complex's sulfide ore bodies are thought to be derived from immiscible magmas formed in the Earth's mantle as a result of the impact (and possibly mixed with meteorite material), while the uppermost layers are thought to be rock debris remaining from the impact.

      A final and highly controversial group of deposits deserves mention because some geologists believe them to have formed from immiscible melts. These are magnetite deposits associated with volcanic rocks of dioritic affinity (i.e., igneous rocks intermediate in composition between granites and gabbros). There is no doubt that lava flows account for the presence of magnetite deposits in northern Chile, but there is great conjecture over whether magnetite bodies associated with these lavas formed as a result of a dioritic magma precipitating an immiscible oxide magma, as a magnetite lava formed through the melting of a previously formed sedimentary iron deposit, or as the source of a hydrothermal solution that deposited the magnetite. Many experts draw the latter conclusion. Considerable controversy also surrounds the origin of the famous Swedish iron ores at Kiruna and Gällivare. These magnetite-apatite bodies encased in volcanic rocks have been variously interpreted as having formed as immiscible oxide magmas, as iron-rich sediments that were subsequently metamorphosed, and as deposits arising from volcanic exhalations.

Hydrothermal solution
      Hydrothermal mineral deposits (hydrothermal mineral deposit) are those in which hot water serves as a concentrating, transporting, and depositing agent. They are the most numerous of all classes of deposit.

      Hydrothermal deposits are never formed from pure water, because pure water is a poor solvent of most ore minerals. Rather, they are formed by hot brines (brine), making it more appropriate to refer to them as products of hydrothermal solutions. Brines, and especially sodium-calcium chloride brines, are effective solvents of many sulfide and oxide ore minerals, and they are even capable of dissolving and transporting native metals such as gold and silver.

      The water in a hydrothermal solution can come from any of several sources. It may be released by a crystallizing magma; it can be expelled from a mass of rock undergoing metamorphism; or it may originate at the Earth's surface as rainwater or seawater and then trickle down to great depths through fractures and porous rocks, where it will be heated, react with adjacent rocks, and become a hydrothermal solution. Regardless of the origin and initial composition of the water, the final compositions of all hydrothermal solutions tend to converge, owing to reactions between solutions and the rocks they encounter.

      Hydrothermal solutions are sodium-calcium chloride brines with additions of magnesium and potassium salts, plus small amounts of many other chemical elements. The solutions range in concentration from a few percent to as much as 50 percent dissolved solids by weight. Existing hydrothermal solutions can be studied at hot springs, in subsurface brine reservoirs such as those in the Imperial Valley of California, the Cheleken Peninsula on the eastern edge of the Caspian Sea in Turkmenistan, in oil-field brines, and in submarine springs along the mid-ocean ridge. Fossil hydrothermal solutions can be studied in fluid inclusions, which are tiny samples of solution trapped in crystal imperfections by a growing mineral.

      Because hydrothermal solutions form as a result of many processes, they are quite common within the Earth's crust. Hydrothermal mineral deposits, on the other hand, are neither common nor very large compared to other geologic features. It is apparent from this that most solutions eventually mix in with the rest of the hydrosphere and leave few obvious traces of their former presence. Those solutions that do form mineral deposits (and thereby leave obvious evidence of their former presence) do so because some process causes them to deposit their dissolved loads in a restricted space or small volume of porous rock. It is most convenient, therefore, to discuss hydrothermal mineral deposits in the context of their settings.

Veins (vein)
      The simplest hydrothermal deposit to visualize is a vein, which forms when a hydrothermal solution flows through an open fissure and deposits its dissolved load. A great many veins occur close to bodies of intrusive igneous rocks because the igneous rocks serve as heat sources that create convectively driven flows in hydrothermal solutions. Precipitation of the minerals is usually caused by cooling of the hydrothermal solution, by boiling, or by chemical reactions between the solution and rocks lining the fissure. Some famous deposits are the tin-copper-lead-zinc veins of Cornwall, Eng.; the gold-quartz veins of Kalgoorlie, W. Aus., Australia, and Kirkland Lake, Ont., Can.; the tin-silver veins of Llallagua and Potosí, Bol.; and the silver-nickel-uranium veins of the Erzgebirge (Ore Mountains), Ger., which were first described by Georgius Agricola (Agricola, Georgius) in his book De re metallica (1556).

 Hydrothermal deposits formed at shallow depths below a boiling hot spring system (see figure—>) are commonly referred to as epithermal, a term retained from an old system of classifying hydrothermal deposits based on the presumed temperature and depth of deposition. Epithermal veins tend not to have great vertical continuity, but many are exceedingly rich and deserving of the term bonanza. Many of the famous silver and gold deposits of the western United States, such as Comstock (Comstock Lode) in Nevada and Cripple Creek in Colorado, are epithermal bonanzas.

Porphyry deposits
      Among the most distinctive hydrothermal deposits is a class known as porphyry copper deposits (porphyry copper deposit), so called because they are invariably associated with igneous intrusives that are porphyritic (meaning the rock is a mixture of coarse and fine mineral grains). Porphyry copper deposits (and their close relatives, porphyry molybdenum deposits) contain disseminated mineralization, meaning that a large volume of shattered rock contains a ramifying network of tiny quartz veins, spaced only a few centimetres apart, in which grains of the copper ore minerals chalcopyrite and bornite (or the molybdenum ore mineral molybdenite) occur with pyrite. The shattered rock serves as a permeable medium for the circulation of a hydrothermal solution, and the volume of rock that is altered and mineralized by the solution can be huge: porphyry coppers are among the largest of all hydrothermal deposits, with some giant deposits containing many billions of tons of ore. Although in most deposits the ore averages only between 0.5 and 1.5 percent copper by weight, the tonnages of ore mined are so large that more than 50 percent of all copper produced comes from porphyry coppers.

 Porphyry coppers are often associated with stratovolcanoes (see figure—>). As a result of the volcanism that rings the Pacific Ocean basin, porphyry coppers are conspicuous features of mineralization along the western borders of North and South America and in the Philippines. Among the major deposits are El Teniente, El Salvador, and Chuquicamata in Chile, Cananea in Mexico, and, in the United States, Bingham Canyon in Utah, Ely and Yerington in Nevada, and San Manuel in Arizona.

Skarns (skarn)
      When a limestone or marble is invaded by a high-temperature hydrothermal solution, the carbonate minerals calcite and dolomite react strongly with the slightly acid solution to form a class of mineral deposit called a skarn. Because solutions tend to have high temperatures close to a magma chamber, most skarns are found immediately adjacent to intrusive igneous rocks. The solutions introduce silica and iron, which combine with the calcium and magnesium in the parent rock to form silicate minerals such as diopside, tremolite, and andradite. The hydrothermal solutions may also deposit ore minerals of iron, copper, zinc, tungsten, or molybdenum.

      The mining of magnetite from a skarn deposit at Cornwall, Penn., U.S., commenced in 1737 and continued for two and a half centuries. copper skarns are found at many places, including Copper Canyon in Nevada and Mines Gaspé in Quebec, Can. Tungsten (tungsten processing) skarns supply much of the world's tungsten from deposits such as those at Sangdong, Korea; King Island, Tas., Australia; and Pine Creek, Calif., U.S.

Volcanogenic massive sulfides
      Wherever volcanism occurs beneath the sea, the potential exists for seawater to penetrate the volcanic rocks, become heated by a magma chamber, and react with the enclosing rocks—in the process concentrating geochemically scarce metals and so forming a hydrothermal solution. When such a solution forms a hot spring on the seafloor, it can suddenly cool and rapidly deposit its dissolved load. Mineral deposits formed by this process, which are called volcanogenic massive sulfide (VMS) deposits, are known in ancient seafloor rocks of all geologic ages. In addition, deposits forming today as a result of submarine hot-spring activity have been discovered at a number of places along the oceanic ridge (the most volcanically active zone on Earth), and in back-arc basins associated with subduction zones.

      VMS deposits constitute some of the richest deposits of copper, lead, and zinc known. Some of the most famous, found in Japan and called kuroko deposits, yield ores that contain as much as 20 percent combined copper, lead, and zinc by weight, plus important amounts of gold and silver. Other famous VMS deposits are the historic copper deposits of Cyprus and, in Canada, the Kidd Creek deposit in Ontario and the Noranda deposits of Quebec.

Mississippi Valley type
      The central plains of North America, running from the Appalachian Mountains on the east to the Rocky Mountains on the west, are underlain by nearly flat sedimentary rocks (sedimentary rock) that were laid down on a now-covered basement of igneous and metamorphic rocks. The cover of sedimentary rocks, which have been little changed since they were deposited, contains numerous strata of limestone, and within the limestones near the bottom of the pile is found a distinctive class of mineral deposit. Because the central plains coincide closely with the drainage basin of the Mississippi River, this class of deposit has come to be called the Mississippi Valley type (MVT).

 MVT deposits are always in limestones and are generally located near the edges of sedimentary basins or around the edges of what were islands or high points in the seafloor when the limestone was deposited (see figure—>). The hydrothermal solutions that introduced the ore minerals (principally the lead mineral galena and the zinc mineral sphalerite) apparently flowed through the sandstones and conglomerates that commonly underlie the limestones. Where they met a barrier to flow, such as a basement high or a basin edge, the solutions moved and reacted with the limestone, depositing ore minerals.

      Among the many famous MVT deposits are the great zinc deposits of Pine Point in Canada's Northwest Territories; the Tri-State zinc district centred on Joplin, Mo., U.S.; the Viburnum Trend of southeast Missouri; deposits in Cumberland, Eng., and in Trepča, Serbia; and the lead-zinc deposits of the central Irish plains.

Stratiform deposits
      A final class of hydrothermal deposit is called stratiform because the ore minerals are always confined within specific strata and are distributed in a manner that resembles particles in a sedimentary rock. Because stratiform deposits so closely resemble sedimentary rocks, controversy surrounds their origin. In certain cases, such as the White Pine copper deposits of Michigan, the historic Kupferschiefer deposits of Germany and Poland, and the important copper deposits of Zambia, research has demonstrated that the origin is similar to that of MVT deposits—that is, a hydrothermal solution moves through a porous aquifer at the base of a pile of sedimentary strata and, at certain places, deposits ore minerals in the overlying shales (shale). The major difference between stratiform deposits and MVT deposits is that, in the case of stratiform deposits, the host rocks are generally shales (fine-grained, clastic sedimentary rocks) containing significant amounts of organic matter and fine-grained pyrite.

      Several of the world's largest and most famous lead-zinc deposits are stratiform; they also are among the most controversial in origin because there are no obvious aquifers underlying the mineralized strata. Three examples are in Australia: Broken Hill in New South Wales, Mount Isa in Queensland, and McArthur River in the Northern Territory. Another example is the famous Canadian lead-zinc deposit at Sullivan, B.C. At Broken Hill, metamorphism has almost completely obscured the original geologic environment, but in the other three cases evidence suggests that hydrothermal fluids moved upward along a fault from deeper within a sedimentary basin, then reacted with a shale while it was still a mud on the seafloor. Details of the actual processes involved remain controversial.

      Groundwater is that part of subsurface water that is below the water table—that is, water in the zone of saturation. For the purpose of the present discussion, the difference between groundwater and hydrothermal solutions is that groundwater retains many of its original chemical characteristics and remains within one kilometre or less of the surface. Such waters form two important classes of deposit.

Roll-front deposits
      Uranium (uranium processing) occurs in two valence states, U4+ and U6+. Weathering of rocks converts uranium into the +6 state, in which state it forms the uranyl ion (UO2)2+. Uranyl compounds tend to be soluble in groundwater, whereas U4+ compounds are not. So long as the groundwater remains oxidizing, uranyl ions are stable and uranium can be transported by groundwater; however, when uranyl ions encounter a reducing agent such as organic matter, U4+ uranium is precipitated as uraninite and coffinite.

      Because groundwater flowing through an aquifer and meeting a reducing zone will deposit a zone, or front, of uraninite, and because the front tends to move slowly forward through the aquifer, dissolving as the oxidizing groundwater moves in and precipitating at the front of the zone, deposits formed in this fashion are known as roll-front uranium deposits. Such deposits have been extensively mined in the western United States, notably in Colorado, Wyoming, Utah, and Texas.

Caliche (calcrete) deposits
      A second class of uranium-bearing groundwater deposits forms in dry land areas where evaporation of groundwater during summer months is an important process. Evaporation causes precipitation of dissolved solids, and the most abundant dissolved solid in dry land groundwater is calcium carbonate. When deposited, this mineral forms a hard, calcareous cement known as caliche. If uranium is present in the groundwater, uranium minerals such as carnotite will also be precipitated and thus form a uraniferous caliche deposit. Extensive deposits of this kind have been identified in the Namib Desert in southwestern Africa and in desert areas of Western Australia.

seawater or lake water
      When either sea or lake waters evaporate, salts are precipitated. These salts include sodium chloride, potassium and magnesium chlorides, borax, and sodium carbonate. Such salts are important economically, but they are not used for the recovery of metals and thus do not warrant discussion here. One very important class of metallic mineral deposit, though, is also formed by precipitation from lake or seawater. This class of deposit comprises compounds of iron or manganese and is known as a chemical sediment, because the mineral constituents are transported in solution and then precipitated to form a sediment as a result of chemical reaction.

iron deposits
      By far the most important metal from an economic and technical point of view is iron. Sedimentary iron deposits, from which almost all iron is obtained, can therefore be viewed as one of the world's great mineral treasures. There are two major types of deposit. The first, and by far the most important, is banded iron formations (BIFs (banded-iron formation)), so called because they are finely layered alternations of cherty silica and an iron mineral, generally hematite, magnetite, or siderite.

      BIFs can be divided into two kinds. The first, and quantitatively most important, is found in sequences of sedimentary rocks deposited in the shallow waters of continental shelves or in ancient sedimentary basins. These deposits are typified by the vast BIFs around Lake Superior and are called Lake Superior-type deposits. Their individual sediment layers can be as thin as 0.5 millimetre (0.02 inch) or as thick as 2.5 centimetres (1 inch), but the alternation of a siliceous band and an iron mineral band is invariable. Several points about Lake Superior-type deposits are remarkable. First, individual thin bands have enormous continuity. During the 1980s, A.F. Trendall, working for the Geological Survey of Western Australia, studied deposits in the Hamersley Basin and found that individual thin layers could be traced for more than 100 kilometres. Such continuity suggests that evaporation played a major role in precipitating both the iron minerals and the silica. A second remarkable feature of Lake Superior-type deposits is that they only formed between 2.7 and 1.8 billion years ago. Such a narrow time frame suggests that the chemistry of the oceans and atmosphere at the time of formation differed greatly from that of the present (in today's ocean, iron is virtually insoluble because the oxidizing atmosphere causes the precipitation of insoluble ferric iron compounds).

      Lake Superior-type BIFs are known and mined on all continents. Among the most famous are the Lake Superior deposits of Michigan and Minnesota, the Labrador Trough deposits of Canada, Serra dos Carajas in Brazil, the Transvaal Basin deposits of South Africa, and the Hamersley Basin of Australia.

      A second kind of BIF, known as an Algoma type, formed over a much wider time range than the Lake Superior type (from 3.8 billion to a few hundred million years ago). Algoma-type BIFs are also finely layered intercalations of silica and an iron mineral, generally hematite or magnetite, but the individual layers lack the lateral continuity of Lake Superior-type BIFs. Algoma-type BIFs are found within rock sequences containing a significant proportion of submarine volcanic rocks, and for this reason it is generally accepted that such deposits formed as a result of submarine volcanism. Such a conclusion is supported by two simple observations: first, that many volcanogenic massive sulfide deposits, such as those in New Brunswick, Can., are found in the same stratigraphic horizons as Algoma-type iron deposits and, second, that in the modern ocean iron-rich, chemically precipitated siliceous layers can sometimes be observed surrounding seafloor hot springs. Important iron deposits of the Algoma type are also exploited in Western Australia and Liberia.

      Historically, a great deal of iron was mined from a second major type of chemically precipitated marine iron deposit. Containing pinhead-sized ooliths (oölite) (small, rounded, accretionary masses formed by repeated deposition of thin layers of an iron mineral), these oolitic iron deposits have been largely supplanted in importance by BIFs, but they once formed the backbone of the iron and steel industries in western Europe and North America. European oolitic iron deposits, commonly called Minette-type deposits, contain ooliths of siderite, a siliceous iron mineral known as chamosite, and goethite. The deposits were formed in shallow, near-shore marine environments and are most extensively developed in England, the Lorraine area of France, Belgium, and Luxembourg. In North America oolitic iron deposits contain ooliths of hematite, siderite, and chamosite and are called Clinton-type deposits. The geologic setting of Clinton-type deposits is very similar to Minette types, the most obvious difference being the presence of goethite in the Minettes and hematite in the Clintons. Clinton-type deposits are found in the Appalachians from Newfoundland to Alabama, and they are several hundred million years older than the Minette-type deposits. Because goethite dehydrates slowly and spontaneously to hematite, it is probable that the major difference between the two deposit types is age.

manganese deposits
      Manganese is very similar to iron in chemistry and in the way it is distributed and concentrated in rocks. Such is the case because manganese, like iron, has two important valence states, Mn2+ and Mn4+. In the +2 state, manganese forms soluble compounds and can be transported in solution. In the +4 state, however, it forms insoluble compounds, and any solution containing Mn2+ in solution will, on meeting an oxidizing environment, quickly precipitate a +4 compound such as pyrolusite, MnO2.

      Manganese forms chemical sediment deposits analogous to the Minette-type iron deposits; that is, the deposits form in shallow, near-shore environments and are oolitic. The most important of such deposits were formed just north of the Black Sea about 35 million years ago during the Oligocene Epoch. Named Chiatura and Nikopol after two cities in Georgia and Ukraine, they contain an estimated 70 percent of the world's known resources of high-grade manganese.

      Manganese deposits similar to Algoma-type iron deposits are widespread. Generally considered to have formed as a result of submarine volcanism, most are too poor to mine, but, where weathering has caused secondary enrichment (discussed below), small but very rich ore deposits have formed. Such deposits are mined in Brazil, Mexico, Gabon, and Ghana.

Rainwater (rain)
      Each of the deposit-forming processes discussed above involves the transport and deposition of ore minerals from solution. But solutions can also form deposits by dissolving and removing valueless material, leaving a residuum of less-soluble ore minerals. Deposits developed as residues from dissolution are called residual deposits. They occur most prominently in warm tropical regions subjected to high rainfall.

Laterites (laterite)
      Soils developed in warm tropical climates tend to be leached of all soluble material. Such soils are called laterites, and the insoluble residues remaining in them are hydroxide minerals of iron and aluminum. Most laterites are such intimate mixtures of iron and aluminum minerals that beneficiation to produce a pure concentrate of one or the other is not possible, but some residual deposits are naturally enriched in one metal or the other and under such circumstances are viable ores.

      Most iron-rich laterites are of little interest, because BIFs are much more desirable ores. However, aluminum-rich (aluminum) laterites, called bauxites (bauxite), are of considerable interest and are the principal ores of aluminum. Bauxites develop either on rocks that are initially low in iron or on iron-rich rocks under circumstances in which organic matter or some other special factor renders iron sufficiently soluble to be separated from the aluminum minerals. Bauxites that are currently forming in tropical regions in Australia, Brazil, West Africa, and elsewhere all contain gibbsite (Al[OH]3) as the ore mineral. Older bauxites contain boehmite and diaspore (both HAlO2), which form as a result of the slow, spontaneous dehydration of gibbsite.

 When mafic igneous rocks such as gabbros and peridotites are subjected to lateritic weathering, nickel released from atomic substitution in the primary igneous silicate minerals can be redeposited at and below the water table as the mineral garnierite, H4Ni3Si2O9 (see figure—>). Although garnierite is a silicate mineral (the most difficult type to smelt), an efficient method has been discovered to recover its nickel content, and it is therefore an excellent ore mineral. The most famous nickeliferous laterites are those of New Caledonia, which have been mined for many years. Other important deposits are known in Australia and Cuba.

Secondary enrichment
      An especially important class of residual deposit is formed by both the removal of valueless material in solution and the solution and redeposition of valuable ore minerals. Because solution and redeposition can produce highly enriched deposits, the process is known as a secondary enrichment.

      Secondary enrichment can affect most classes of ore deposit, but it is notably important in three circumstances. The first circumstance arises when gold-bearing (gold) rocks—even rocks containing only traces of gold—are subjected to lateritic weathering. Under such circumstances, the gold can be secondarily enriched into nuggets near the base of the laterite. The importance of secondary enrichment of gold in lateritic regions was realized only during the gold boom of the 1980s, especially in Australia.

      The second circumstance involves mineral deposits containing sulfide minerals (sulfide mineral), especially copper sulfides, that are subjected to weathering under desert conditions. Sulfide minerals are oxidized at the surface and produce sulfuric acid, and acidified rainwater then carries the copper, as copper sulfate, down to the water table. Below the water table, where sulfide minerals remain unoxidized, any iron sulfide grains present will react with the copper sulfate solution, putting iron into solution and precipitating a copper mineral. The net result is that copper is transferred from the oxidizing upper portion of the deposit to that portion at and just below the water table. Secondary enrichment of porphyry copper deposits in the southwestern United States, Mexico, Peru, and Chile is an important factor in making those deposits ores. Lead, zinc, and silver deposits are also subject to secondary enrichment under conditions of desert weathering.

      The third circumstance in which secondary enrichment is important involves BIFs (banded-iron formation) and sedimentary manganese deposits. A primary BIF may contain only 25 to 30 percent iron by weight, but, when subjected to intense weathering and secondary enrichment, portions of the deposit can be enriched to as high as 65 percent iron. Some primary BIFs are now mined and beneficiated under the name taconite, but in essentially all of these deposits mining actually commenced in the high-grade secondary-enrichment zone. Sedimentary manganese deposits, especially those formed as a result of submarine volcanism, must also be secondarily enriched before they become ores.

Flowing surface water
      When mineral grains of different density are moved by flowing water, the less dense grains will be most rapidly moved, and a separation of high-density and low-density grains can be effected. Mineral deposits formed as a result of gravity separation based on density are called placer deposits (placer deposit).

      For effective concentration, placer minerals must not only have a high density (greater than about 3.3 grams per cubic centimetre), they must also possess a high degree of chemical resistance to dissolution or reaction with surface water and be mechanically durable. The common sulfide ore minerals do not form placers, because they rapidly oxidize and break down. Ore minerals having suitable properties for forming placers are the oxides cassiterite (tin), chromite (chromium), columbite (niobium), ilmenite and rutile (titanium), magnetite (iron), monazite and xenotime (rare-earth metals), and zircon (zirconium). In addition, native gold and platinum have been mined from placers, and several gemstone minerals—in particular, diamond, ruby, and sapphire—also concentrate in placers.

Alluvial placers
 After a mineral-bearing soil reaches the bottom of a slope, it can be moved by stream water so that stream or alluvial placers form (see figure—>). Alluvial placers have played an especially important historical role in the production of gold. Indeed, more than half of the gold ever mined has come from placers, since the giant Witwatersrand gold deposits in South Africa are fossil placers more than two billion years old. Other fossil placers (i.e., deposits whose stream waters have long disappeared) have been discovered at Serra de Jacobina in Brazil (gold) and Blind River, Ont., Can. (uranium), but nowhere has a deposit been found equal to Witwatersrand. Just why and how such an extraordinarily large concentration of gold occurred there is a matter of continuing scientific controversy.

Beach placers
      When wave trains impinge obliquely on a beach, a net flow of water, called a longshore drift, occurs parallel to the beach. Such a current can produce a beach placer. Beach placers are a major source of ilmenite, rutile, monazite, and zircon. They have been extensively mined in India, Australia, Alaska (U.S.), and Brazil.

Metallogenic provinces (metallogenic province) and epochs
      Mineral deposits are not distributed uniformly through the Earth's crust. Rather, specific classes of deposit tend to be concentrated in particular areas or regions called metallogenic provinces. These groupings of deposits occur because deposit-forming processes, such as the emplacement of magma bodies and the formation of sedimentary basins, are themselves controlled by larger processes that shape the face of the Earth. The shape and location of such features as continents and oceans, volcanoes, sedimentary basins, and mountain ranges are controlled, either directly or indirectly, through the process known as plate tectonics—the lateral motion of segments of the lithosphere, the outermost 100-kilometre-thick layer of the Earth. For example, the distribution of hydrothermal mineral deposits, which form as a result of volcanism, is controlled by plate tectonics because most of the Earth's volcanism occurs along plate margins. In addition, porphyry copper deposits are formed as a result of volcanism along a subduction zone (i.e., the zone where one plate descends beneath another); this gives rise to metallogenic provinces parallel to subduction plate edges. Evidence indicates that plate tectonics has operated for at least two billion years, so that the locations and features of most metallogenic provinces formed over this period can be explained, at least in part, by this geologic process. Factors controlling the distribution of deposits formed more than two billion years ago are still a matter for research, but they too may have been linked to plate tectonics.

      Metallogenic epochs are units of geologic time during which conditions were particularly favourable for the formation of specific classes of mineral deposit. One conspicuous example of a metallogenic epoch is the previously mentioned 900-million-year period, from 2.7 to 1.8 billion years ago, when all of the great Lake Superior-type BIFs were formed. Because the iron in these deposits was deposited from seawater (an impossibility today, since the atmosphere is too oxidizing to allow seawater to transport iron), it is probable that a specific composition of the atmosphere and ocean peculiar to that period defined the BIF metallogenic epoch. Another great deposit-forming period occurred between about 2.8 and 2.65 billion years ago, when a large number of volcanogenic massive sulfide deposits formed; the probable cause of this metallogenic epoch was a period of extremely active submarine volcanism.

Brian J. Skinner

Additional Reading
Two texts for nonspecialists on the subject that also address questions of how the use of mineral resources affects the use of other resources and the environment are Brian J. Skinner, Earth Resources, 3rd ed. (1986); and James R. Craig, David J. Vaughan, and Brian J. Skinner, Resources of the Earth, 3rd ed. (2001). Three excellent references dealing with the broad topic of the genesis of mineral deposits, but written at a level requiring at least a beginning familiarity with geologic terminology, are Anthony M. Evans, An Introduction to Ore Geology, 2nd ed. (1987); Richard Edwards and Keith Atkinson, Ore Deposit Geology and Its Influence on Mineral Exploration (1986); and John M. Guilbert and Charles F. Park, Jr., The Geology of Ore Deposits (1986). More specialized texts deal with specific classes of mineral deposits or with specific processes that form mineral deposits. In all instances, such texts presume a working knowledge of the technical terms used in geology. Among the best specialized texts are Hubert Lloyd Barnes (ed.), Geochemistry of Hydrothermal Ore Deposits, 3rd ed. (1997); J. Barry Maynard, Geochemistry of Sedimentary Ore Deposits (1983); Anthony J. Naldrett, Magmatic Sulfide Deposits (1989); F.J. Sawkins, Metal Deposits in Relation to Plate Tectonics, 2nd rev. ed. (1990); B.R. Berger and P.M. Bethke (eds.), Geology and Geochemistry of Epithermal Systems (1985); J.A. Whitney and Anthony J. Naldrett (eds.), Ore Deposition Associated with Magmas (1989); and E.R. Force, J.J. Eidel, and J. Barry Maynard (eds.), Sedimentary and Diagenetic Mineral Deposits: A Basin Analysis Approach to Exploration (1991).Brian J. Skinner

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • mineral deposit — naudingųjų išgavų telkinys statusas T sritis ekologija ir aplinkotyra apibrėžtis Naudingųjų išgavų sankaupa Žemės plutoje, ribojama gretimų uolienų ir slūgsanti kaip klodas, gysla ar lęšis. atitikmenys: angl. accumulation; cluster; mineral… …   Ekologijos terminų aiškinamasis žodynas

  • hydrothermal mineral deposit — ▪ geology       any concentration of metallic minerals formed by the precipitation of solids from hot mineral laden water (hydrothermal solution). The solutions are thought to arise in most cases from the action of deeply circulating water heated …   Universalium

  • Mineral resource classification — is the classification of mineral deposits based on their geologic certainty and economic value. A McKelvey diagram showing the relationship of mineral resource classifications to economics and geologic certainty.[1] Mineral deposits can be… …   Wikipedia

  • Mineral lick — Salt lick redirects here. For other uses, see Salt lick (disambiguation). Gaur at a natural salt lick A mineral lick (also known as salt lick or natural lick) is a natural mineral deposit where animals in nutrient poor ecosystems can obtain… …   Wikipedia

  • deposit — de·pos·it || dɪ pÉ’zɪt n. sum of money that has been put in a bank for safekeeping; partial payment, pledge; (Geology) layer (of sand, mineral deposit, etc.) v. put a sum of money in a bank for safekeeping; pay in part; (Geology) set down,… …   English contemporary dictionary

  • deposit — 1. noun a) Sediment or rock that is not native to its present location or is different from the surrounding material. Sometimes refers to ore or gems. a mineral deposit b) asset that was left to the care of the other <!this definition could… …   Wiktionary

  • deposit — Synonyms and related words: accumulation, acquitment, acquittal, acquittance, allocation, alluvion, alluvium, amortization, amortizement, ash, assignment, bank, binder, bond, bosom, bottle up, bottomry, bundle away, bury, cache, cash, cash… …   Moby Thesaurus

  • deposit — v 1. put, place, set, lay, (of coins) insert, Archaic. repose; put down, set down, lay down. 2. precipitate, settle, dump. 3. bank, save, hoard; store, stow, reposit, lay in, lay away, put away, squirrel away. n 4. silt, alluvium, precipitation,… …   A Note on the Style of the synonym finder

  • Mineral exploration — is the process of finding ore (commercially viable concentrations of minerals) to mine. Mineral exploration is a much more intensive, organized and professional form of mineral prospecting and, though it frequently uses the services of… …   Wikipedia

  • Deposit — De*pos it, n. [L. depositum, fr. depositus, p. p. of deponere: cf. F. d[ e]p[^o]t, OF. depost. See {Deposit}, v. t., and cf. {Depot}.] 1. That which is deposited, or laid or thrown down; as, a deposit in a flue; especially, matter precipitated… …   The Collaborative International Dictionary of English

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”