Biological roles of the elements

Last updated

A large fraction of the chemical elements that occur naturally on the Earth's surface are essential to the structure and metabolism of living things. Four of these elements (hydrogen, carbon, nitrogen, and oxygen) are essential to every living thing and collectively make up 99% of the mass of protoplasm. [1] Phosphorus and sulfur are also common essential elements, essential to the structure of nucleic acids and amino acids, respectively. Chlorine, potassium, magnesium, calcium and phosphorus have important roles due to their ready ionization and utility in regulating membrane activity and osmotic potential. [2] The remaining elements found in living things are primarily metals that play a role in determining protein structure. Examples include iron, essential to hemoglobin; and magnesium, essential to chlorophyll. Some elements are essential only to certain taxonomic groups of organisms, particularly the prokaryotes. For instance, the lanthanide series rare earths are essential for methanogens. As shown in the following table, there is strong evidence that 19 of the elements are essential to all living things, and another 17 are essential to some taxonomic groups. Of these 17, most have not been extensively studied, and their biological importance may be greater than currently supposed.

The remaining elements are not known to be essential. There appear to be several causes of this.

Aluminum warrants special mention because it is the most abundant metal and the third most abundant element in the Earth's crust; [3] despite this, it is not essential for life. With this sole exception, the eight most highly abundant elements in the Earth's crust, making up over 90% of the crustal mass, [3] are also essential for life.

Essential elements [4] [5] [6] [7] [8]
H He
LiBe BCNOFNe
NaMg AlSiPSClAr
KCaScTiVCrMnFeCoNiCuZnGaGeAsSeBrKr
RbSrYZrNbMoTcRuRhPdAgCdInSnSbTeIXe
CsBa*LuHfTaWReOsIrPtAuHgTlPbBiPoAtRn
FrRa**LrRfDbSgBhHsMtDsRgCnNhFlMcLvTsOg
 
 *LaCePrNdPmSmEuGdTbDyHoErTmYb
 **AcThPaUNpPuAmCmBkCfEsFmMdNo
Legend:
  Quantity elements
  Essential trace elements
  Essentiality or function in mammals debated
  No evidence for biological action in mammals, but essential in some lower organisms.
(In the case of the lanthanides, the definition of an essential nutrient as being indispensable and irreplaceable is not completely applicable due to their extreme similarity. The stable early lanthanides La–Nd are known to stimulate the growth of various lanthanide-using organisms, and Sm–Gd show lesser effects for some such organisms. The later elements in the lanthanide series do not appear to have such effects.) [9]

The following list identifies in rank order the possible biological roles of the chemical elements, ranging from a score of 5 for elements essential to all living things, to a score of 1 for elements that have no known effects on living things. There are also letter scores for special functions of the elements. These rank scores are used to characterize each element in the following table.

RankBiological Importance
5Essential for all (or most) living things.
4Essential for some living things.
3Not essential, but has a pharmacologic role; helps to treat disease in some organisms.
2Benign: present in some organisms, sometimes bioaccumulating, but generally having no apparent effects (except possible harmful effects, notes "a" or "b").
1Extremely rare on the Earth's surface (less than 1×10−7%, i.e. less than 1/10 as common as the least common essential element, selenium), thus has low potential for any kind of biological role.
aToxic in some molecular forms.
bRadioactive.
cHas uses in medicine as a drug or implant.

The following table identifies the 94 chemical elements that occur naturally on the Earth's surface, their atomic numbers, their biological rank as defined above, and their general beneficial and harmful roles in living things.

Element Z RankBeneficial roleHarmful role
actinium 891bHas no known biological role. [10] Radioactive.
aluminum 132aHas no known biological role. [10] [11] The metal, or various compounds, can be toxic to humans. [12] In plants, aluminum can be the primary limitation on growth in acidic soils. [13]
antimony 512cHas no known biological role, but has a variety of uses in medicine, e.g. antibacterial. [14] Some compounds are highly toxic to humans. [10]
argon 182None known. [10] None known.
arsenic 334aEssential to some species, including humans, for whom it is necessary for the functioning of the nervous system. [15] Some marine algae and shrimp contain arsenic compounds. [10] Toxic to humans in some forms. [10]
astatine 851bNone known. [10] Radioactive.
barium 562acHas no known biological role, but a variety of plants concentrate it from the soil, and it has a variety of uses in medicine. [10] Some compounds are toxic. In humans, barium ion affects the nervous system. [16]
beryllium 42cHas no known biological role, but has medical use in certain dental alloys [17] Toxic to humans, esp. via inhalation. Can substitute for magnesium in certain key enzymes, causing malfunction. [10]
bismuth 832acHas no known biological role, but has a variety of uses in medicine, e.g. in antiulcer, antibacterial, anti‐HIV and radiotherapeutic uses. [14] [18] Slightly toxic, perhaps the least toxic heavy metal, though poisonings have been reported. [19]
boron 54In plants, it has important roles in nucleic acid metabolism, carbohydrate and protein metabolism, cell wall synthesis, cell wall structure, membrane integrity and function, and phenol metabolism. [20] Probably essential to animals, for reasons not well understood. [21] Toxic to both animals and plants. [22]
bromine 355Essential to membrane architecture and tissue development in animals. [23] May have antibiotic effects in some compounds when it substitutes for chlorine. [24] Bromine compounds are very common in and presumably essential to a variety of marine organisms, including bacteria, fungi, seaweeds, and diatoms. [25] [26] Most marine organobromine compounds are made by the action of a unique algal enzyme, vanadium bromoperoxidase [27] Toxic in excessive concentrations, causing the human disease bromism.
cadmium 484A carbonic anhydrase using cadmium has been found in some marine diatoms that inhabit environments with very low zinc availability; the cadmium evidently provides a similar function. [28] Many plants bioaccumulate cadmium, which deters herbivory. [29] Cadmium deprivation in goats and rats leads to depressed growth, but has not been shown to be essential. [21] Cadmium poisoning is widely recognized in humans, but has not been described in other organisms. In general, cadmium acts by substituting for calcium, zinc, or iron, and can disrupt biochemical pathways dependent upon those metals. [30]
calcium 205aUbiquitous, essential [31] Appears in various toxic organochemicals; contributes to diseases e.g. kidney stones. [32]
carbon 65cUbiquitous, essential. [10] Its oxide is a pollutant. [33]
cerium 584aThe methanol dehydrogenase of the methanotrophic bacterium Methylacidiphilum fumariolicum SolV requires a lanthanide cofactor, lanthanum, cerium, praseodymium, or neodymium (or possibly other lanthanides) [34] but it appears that any of these lanthanides can perform this function, so cerium is only essential if no other suitable lanthanides are available. Has medical uses, e.g. in burn treatment. [10] Can substitute for calcium with possible adverse effects, and in metallic form, is mildly toxic. [10]
caesium 552aHas no known biological role. [10] Can substitute for potassium (a biologically essential element) with possible adverse effects, [10] particularly if the substitution is of radioactive cesium, which was the primary biologically active isotope released in the 1986 Chernobyl nuclear disaster. [10] [35]
chlorine 174aChlorine salts are critical for many species, including humans. [10] Its ion is used as an electrolyte, as well as making the hydrochloric acid the stomach uses for digestion. [15] Elemental Cl2 is toxic. [10]
chromium 244Appears to be essential in humans. Affects insulin metabolism. [10] Also influences metabolism, replication and transcription of nucleic acids, and decreases the content of corticosteroids in plasma. [36] Toxic in some forms. [10]
cobalt 275Essential to the metabolism of all animals, as a key constituent of cobalamin, also known as vitamin B12. [10] Toxic in some forms, probably carcinogenic. [10]
copper 295aEssential in many ways; an important component of many enzymes, especially cytochrome c oxidase, which is present in nearly all living things. [10] [37] Some compounds are toxic; [10] the metal is highly toxic to viruses. [38]
dysprosium 662Has no known biological role. [10] Some salts have low toxicity. [39]
erbium 682aHas no known function in humans, and is not taken up by plants. [10] Soluble salts are mildly toxic. [39]
europium 632aHas no known function in humans, and is not taken up by plants. [10] Possible low toxicity in some forms. [10]
fluorine 93aAffects bone density in humans; creates fluoroapatite, which makes tooth enamel hard and relatively impervious to chemical action, compared to bone. [10] Improves growth in rats; has pharmacologic effects – helps to treat other deficiencies, e.g. of iron. Absence of fluorine has no clear adverse consequences in animals. [21] Excess fluorine in humans results in fluoride toxicity, and can substitute for iodine, causing goitre.
francium 871bDue to its very short half-life, there is almost no potential for a living thing to be exposed to it. Even synthesis cannot produce more than minute quantities before it decays, so there is no medical use. [10] Radioactive. [10]
gadolinium 642acHas no known function in humans, and is not taken up by plants. [10] There has been limited use in experimental medicine. [40] Soluble salts are mildly toxic. [10] See medical discussion in Gadolinium: Safety.
gallium 312acAlthough nonessential, plays a complex role in humans, including concentrating in bone, binding to plasma proteins, and concentrating in malignancies. [41] It is selectively taken up by plants, so there are a variety of possible roles in plant metabolism. [42] There is limited medical use. [10] Inhibits iron uptake and metabolism in a variety of plants and bacteria. [42]
germanium 322aSome plants will take it up, but it has no known metabolic role. [10] Some salts are deadly to some bacteria. [10]
gold 792aAlthough some plants bioaccumulate gold, no living organism is known to require it. There are medical uses, including treatment of rheumatoid arthritis and fabrication of dental implants. [10] Some gold salts used in medicine have adverse side effects.
hafnium 722Has no known biological role. [10] Salts have low toxicity. [10]
helium 22As with other noble gases, has no known biological role. [10] Has no known harmful role.
holmium 672aThis lanthanide has no known biological roles, and is not taken up by plants. [10] There are medical uses; for example, holmium-containing nanoparticles are biocompatible and facilitate NMR imaging. [43] Some salts are known to be toxic to humans. [39]
hydrogen 15Ubiquitous, essential. [10] None known. [10]
indium 492aHas no known biological role. [10] Highly toxic to humans in fairly small doses; [44] mildly toxic to plants, comparable to aluminum; [45] may inhibit growth of some bacteria.
iodine 535acIodine has a role in biochemical pathways of organisms from all biological kingdoms, indicating it is uniformly essential to life [46] Widely used in medicine, mainly for treatment of goitre and for its antibacterial properties. [10] Highly toxic to humans in its elemental form. [10]
iridium 771aDue to its extreme rarity, iridium has no biological role. [10] The chloride is moderately toxic to humans. [10]
iron 265Essential to almost all living things, usually as a ligand in a protein; it is most familiar as an essential element in the protein hemoglobin. [10] Toxic in some forms. [10]
krypton 361As with other noble gases, has no known biological role. [10] It is also the rarest non-radioactive element in the Earth's crust. [3] None known.
lanthanum 574acThe methanol dehydrogenase of the methanotrophic bacterium Methylacidiphilum fumariolicum SolV requires a lanthanide cofactor, lanthanum, cerium, praseodymium, or neodymium (or possibly other lanthanides) [34] but it appears that any of these lanthanides can perform this function, so lanthanum is only essential if no other suitable lanthanides are available. Among plants, Carya accumulates lanthanum and other lanthanides, perhaps as an adaptation to certain site-limiting environmental stresses. [47] The chloride is mildly toxic to humans. [10]
lead 823aPb deprivation leads to suboptimal growth of rats, along with anemia, and reduced function of a variety of enzymes; but results have been inconclusive, and the effects may be pharmacologic. [21] Toxic in some forms, teratogenic, and carcinogenic; historically, lead poisoning has frequently been widespread in human societies. [10] It seems to have been rarely documented in other organisms.
lithium 34aThere is some evidence that lithium deprivation adversely affects multiple functions, especially fertility and adrenal gland function, in rats and goats, [21] and some plants accumulate lithium. [10] However, it is not known to be essential for any organism. There are medical uses, especially in treatment of manic-depressive symptoms. [10] Toxic in some forms. [10]
lutetium 712aThis lanthanide has no known biological roles, and is not taken up by plants. [10] Mildly toxic to humans in some forms. [10]
magnesium 125aEssential for almost all living things; needed for chlorophyll, and is a co-factor for many other enzymes; has multiple medical uses. [10] Large doses can have toxic effects. [10]
manganese 255aEssential for almost all living things, although in very small amounts; it is a cofactor for many classes of enzymes. [10] [48] At least one of these, mitochondrial superoxide dismutase (MnSOD), is present in all aerobic Bacteria and in the mitochondria of all eukaryotes. [49] Large doses can have toxic effects. [10]
mercury 802acAlthough nearly ubiquitous in the environment, mercury has no known biological role. Traditionally used in medicine and dental fillings, it is now avoided due to toxic side effects. [10] Can inactivate certain enzymes, as a result, both the metal and some compounds (especially methylmercury) are harmful to most life forms; there is a long and complex history of mercury poisoning in humans. [10]
molybdenum 425Found in many enzymes; essential to all eukaryotes, and to some bacteria. [50] [51] Molybdenum in proteins is bound by molybdopterin or to other chemical moieties to give one of the molybdenum cofactors. [52] Metallic molybdenum is toxic if ingested. [53] [54]
neodymium 604The methanol dehydrogenase of the methanotrophic bacterium Methylacidiphilum fumariolicum SolV requires a lanthanide cofactor, lanthanum, cerium, praseodymium, or neodymium (or possibly other lanthanides) [34] but it appears that any of these lanthanides can perform this function, so neodymium is only essential if no other suitable lanthanides are available.Toxic in some forms. Anticoagulant. [10]
neon 102As with other noble gases, has no known biological role. [10] None known.
neptunium 931bHas no known biological role. [10] Radioactive. [10]
nickel 284As a component of urease, and many other enzymes as well, nickel is needed by most living things in all domains. [55] [56] Nickel hyperaccumulator plants use it to deter herbivory. [57] Toxic in some forms. [10]
niobium 412Has no known biological role, although it does bioaccumulate in human bone. [10] Is hypoallergenic and, both alone and in a niobium-titanium alloy, is used in some medical implants including prosthetics, orthopedic implants, and dental implants. [58] [59] Toxic in some forms. [10]
nitrogen 75Ubiquitous, essential for all forms of life; all proteins and nucleic acids contain substantial amounts of nitrogen. [10] Toxic in some forms. [10]
osmium 761aNone known. [10] Osmium is very rare, substantially more so than any element essential to life. [3] The oxide is toxic to humans. [10]
oxygen 85Ubiquitous, essential for all forms of life; essentially all biological molecules (not to mention water) contain substantial amounts of oxygen. [10] In high concentrations, oxygen toxicity can occur.
palladium 462aHas no known biological role. [10] Medically, it is used in some dental amalgams to decrease corrosion and increase the metallic lustre of the final restoration. [60] Toxic in some forms. [10]
phosphorus 155Ubiquitous, essential for all forms of life; all nucleic acids contain substantial amounts of phosphorus; it is also essential to adenosine triphosphate (ATP), the basis for all cellular energy transfer; and it performs many other essential roles in different organisms. [10] Toxic in some forms; pure phosphorus is poisonous to humans. [10]
platinum 782cHas no known biological role, but it is a component of the drug cisplatin, which is highly effective in treating some forms of cancer. [10] Toxic in some forms. Contact can promote an allergic reaction (platinosis) in humans. [10]
plutonium 941bcHas no known biological role, and is extremely rare in the Earth's crust. The isotope plutonium-238 is used as an energy source in some heart pacemakers. [10] Both toxic and radioactive.
polonium 841bHas no known biological role, and due to its short half-life, is nearly nonexistent outside of research facilities. [10] Both highly toxic and radioactive.
potassium 195aEssential for almost all living things, except perhaps some prokaryotes; performs numerous functions, most of which are related to the transport of potassium ions. [10] Potassium ion in excess causes paralysis and depresses central nervous system activity in humans. [10]
praseodymium 594The methanol dehydrogenase of the methanotrophic bacterium Methylacidiphilum fumariolicum SolV requires a lanthanide cofactor, lanthanum, cerium, praseodymium, or neodymium (or possibly other lanthanides) [34] but it appears that any of these lanthanides can perform this function, so praseodymium is only essential if no other suitable lanthanides are available.Some forms are mildly toxic to humans. [10]
promethium 611bHas no known biological role; as it is radioactive with a short half-life, it is very rare and is seldom present for long. [10] Radioactive. [10]
protactinium 911bHas no known biological role; as it is radioactive with a short half-life, it is very rare and is seldom present for long. [10] Both toxic and highly radioactive.
radium 881bcHas no known biological role; as it is radioactive it is very rare. There have been various medical uses in the past. [10] Radioactive; historically, there have been many cases of radium poisoning, most notably in the case of the Radium Girls.
radon 861bcHas no known biological role. [10] Historically, there have been various medical uses.Radioactive, [10] with a variety of documented harmful effects on human health.
rhenium 751Has no known biological role, [10] and is extremely rare in the Earth's crust.None known. [10]
rhodium 451Has no known biological role, [10] and is extremely rare in the Earth's crust.Toxic in some forms. [10]
rubidium 372cHas no known biological role, although it seems to substitute for potassium, and bioaccumulates in plants. It has seen limited medical use. [10] None known. [10]
ruthenium 441aHas no known biological role; it bioaccumulates, but does not appear to have any function. It is extremely rare. [10] There is a highly toxic oxide, RuO4, but it is not naturally occurring. [10]
samarium 622acHas no known biological role, although it can bioaccumulate in some plants. One radioisotope is approved for medical use. [10] Toxic in some forms. [10]
scandium 212aHas no known biological role, but can bioaccumulate in some plants, perhaps because it can substitute for aluminum in some compounds. [10] Some compounds may be carcinogenic; some forms are mildly toxic to humans. [10]
selenium 344Selenium, which is an essential element for animals and prokaryotes and is a beneficial element for many plants, is the least-common of all the elements essential to life. [3] [61] Selenium acts as the catalytic center of several antioxidant enzymes, such as glutathione peroxidase, [10] and plays a wide variety of other biological roles.Toxic in some forms. [10]
silicon 144cEssential for connective tissue and bone in birds and mammals. [21] Silica appears in many organisms; e.g. as frustules (shells) of diatoms, spicules of sponges, and phytoliths of plants. [10] Also has medical uses, e.g. cosmetic implants. [10] Silicosis is a lung disease caused by inhalation of silica dust.
silver 472cHas no known biological role, apart from medical use (antibiotic, mainly; also dental fillings). [10] Can produce a variety of toxic effects in humans and other animals; also toxic to various microorganisms. [10]
sodium 115Essential to animals and plants in many ways, such as osmoregulation and transmission of nerve impulses. [10] Essential to energy metabolism of some bacteria, particularly extremophiles. [62] Toxic in some forms, and since it is essential to living things, either a lack or an excess can have harmful results.
strontium 384cEssential to Acantharean radiolarians, which have skeletons of strontium sulfate. [63] Also essential to some stony corals. [10] Limited medical use in drugs such as strontium ranelate.Non-toxic; in humans, it often substitutes for calcium. [10]
sulfur 165Sulfur is essential and ubiquitous, partly because it is part of the amino acids cysteine and methionine. Many metals that appear as enzyme cofactors are bound by cysteine, and methionine is essential for protein synthesis.Toxic in some forms.
tantalum 731cHas no known biological role, but is biocompatible, used in medical implants, e.g. skull plates. [10] Has not been found to be toxic, though some patients with tantalum implants have shown a mildly allergic reaction. [10]
technetium 431bNonexistent (radioactive). [10] Nonexistent (radioactive). [10]
tellurium 521aIs not known to be essential to any organism, but is metabolized by humans, typically through methylation. [10] Toxic in some forms; the sodium salt is fatal to humans in small doses, and the oxide causes severe bad breath. [10]
terbium 652aHas no known biological role, but is probably similar to other lanthanides such as cerium and lanthanum, i.e., not known to be essential. [10] Terbium is also one of the rarer lanthanides.Toxic in some forms. [10]
thallium 812aHas no known biological role. Medically, it was used for many years to induce hair loss, but this has ended due to its numerous other toxic effects on human health. [10] Its role, if any, in living things other than humans has been very little explored.It is very toxic and there is evidence that the vapor is both teratogenic and carcinogenic. [64] It can displace potassium in humans affecting the central nervous system. Thallium poisoning has a long history in humans, especially as it has sometimes been a preferred poison.
thorium 901bHas no known biological role. [10] Radioactive.
thulium 692aNo known function in humans, and is not taken up by plants. [10] Toxic in some forms.
tin 504aIn mammals, deprivation causes impaired reproduction and other abnormal growth, [21] suggesting that it is an essential element. Tin may have a role in tertiary structure of proteins. Some plants are tin hyperaccumulators, possibly to deter herbivory.Toxic in some forms, especially the organotin compounds, which include many potent biocides.
titanium 222cPresent in most animals, possibly beneficial to plant growth, but not known to be essential; some plants are hyperaccumulators. [10] Common in medical implants. [10] The common compounds are nontoxic. [10]
tungsten 744aIs a (presumably essential) component of a few bacterial enzymes, and is the heaviest biologically essential element. [65] Appears to be essential in ATP metabolism of some thermophilic archaea. Can substitute for molybdenum in some proteins. Some plants hyperaccumulate it, though its function is unknown. [10] Toxic, at least to animals, in some forms. [66] [67]
uranium 924bSome bacteria reduce uranium and use it as a terminal electron acceptor for respiration with acetate as electron donor. [68] Some bacteria hyperaccumulate uranium. [10] Radioactive, and most compounds are also chemically toxic to humans. [10]
vanadium 234aCan mimic and potentiate the effect of various growth factors such as insulin and epidermal growth factor. Can also affect processes regulated by cAMP. [69] Also used by some bacteria. Dinitrogenases, essential for nitrogen metabolism, normally use molybdenum but in its absence vanadium (or iron) will substitute. [70] Vanadium is also an essential for a variety of peroxidases found in many taxonomic groups, including bromoperoxidases, haloperoxidases, and chloroperoxidases. [71] Some compounds are toxic, and are implicated in several human diseases of including diabetes, cancer, chlorosis, anemia, and tuberculosis. [69]
xenon 541Has no known biological role. [10] None known.
ytterbium 702aNo known function in humans, where it concentrates in bones. Not taken up by plants. [10] Toxic in some forms. [10]
yttrium 392aNot well understood. It occurs in most organisms and at widely varying concentrations, suggesting it does have a role, but not known whether essential. [10] Toxic in some forms, and it may be carcinogenic. [10]
zinc 305aEssential, involved in numerous aspects of cellular metabolism (more than 200 different proteins). Some plants are hyperaccumulators. There are also medical uses, e.g. in dentistry. [10] Some compounds are toxic. [10]
zirconium 402aSome plants have high uptake, but it doesn't appear to be essential or even to have a role; benign. [10] Compounds generally have low toxicity. [10]

See also

Related Research Articles

<span class="mw-page-title-main">Molybdenum</span> Chemical element, symbol Mo and atomic number 42

Molybdenum is a chemical element; it has symbol Mo and atomic number 42. The name derived from Ancient Greek Μόλυβδος molybdos, meaning lead, since its ores were confused with lead ores. Molybdenum minerals have been known throughout history, but the element was discovered in 1778 by Carl Wilhelm Scheele. The metal was first isolated in 1781 by Peter Jacob Hjelm.

Nitrogen fixation is a chemical process by which molecular nitrogen (N
2
), which has a strong triple covalent bond, is converted into ammonia (NH
3
) or related nitrogenous compounds, typically in soil or aquatic systems but also in industry. The nitrogen in air is molecular dinitrogen, a relatively nonreactive molecule that is metabolically useless to all but a few microorganisms. Biological nitrogen fixation or diazotrophy is an important microbe-mediated process that converts dinitrogen (N2) gas to ammonia (NH3) using the nitrogenase protein complex (Nif).

<span class="mw-page-title-main">Vanadium</span> Chemical element, symbol V and atomic number 23

Vanadium is a chemical element; it has symbol V and atomic number 23. It is a hard, silvery-grey, malleable transition metal. The elemental metal is rarely found in nature, but once isolated artificially, the formation of an oxide layer (passivation) somewhat stabilizes the free metal against further oxidation.

<span class="mw-page-title-main">Metalloprotein</span> Protein that contains a metal ion cofactor

Metalloprotein is a generic term for a protein that contains a metal ion cofactor. A large proportion of all proteins are part of this category. For instance, at least 1000 human proteins contain zinc-binding protein domains although there may be up to 3000 human zinc metalloproteins.

<span class="mw-page-title-main">Mineral (nutrient)</span> Chemical element required as an essential nutrient by organisms to perform life functions

In the context of nutrition, a mineral is a chemical element. Some "minerals" are essential for life, most are not. Minerals are one of the four groups of essential nutrients, the others of which are vitamins, essential fatty acids, and essential amino acids. The five major minerals in the human body are calcium, phosphorus, potassium, sodium, and magnesium. The remaining elements are called "trace elements". The generally accepted trace elements are iron, chlorine, cobalt, copper, zinc, manganese, molybdenum, iodine, and selenium; there is some evidence that there may be more.

<span class="mw-page-title-main">Metallothionein</span> Family of proteins

Metallothionein (MT) is a family of cysteine-rich, low molecular weight proteins. They are localized to the membrane of the Golgi apparatus. MTs have the capacity to bind both physiological and xenobiotic heavy metals through the thiol group of its cysteine residues, which represent nearly 30% of its constituent amino acid residues.

<span class="mw-page-title-main">Group 5 element</span> Group of elements in the periodic table

Group 5 is a group of elements in the periodic table. Group 5 contains vanadium (V), niobium (Nb), tantalum (Ta) and dubnium (Db). This group lies in the d-block of the periodic table. This group is sometimes called the vanadium group or vanadium family after its lightest member; however, the group itself has not acquired a trivial name because it belongs to the broader grouping of the transition metals.

<span class="mw-page-title-main">Cofactor (biochemistry)</span> Non-protein chemical compound or metallic ion

A cofactor is a non-protein chemical compound or metallic ion that is required for an enzyme's role as a catalyst. Cofactors can be considered "helper molecules" that assist in biochemical transformations. The rates at which these happen are characterized in an area of study called enzyme kinetics. Cofactors typically differ from ligands in that they often derive their function by remaining bound.

<span class="mw-page-title-main">Nitrogenase</span> Class of enzymes

Nitrogenases are enzymes (EC 1.18.6.1EC 1.19.6.1) that are produced by certain bacteria, such as cyanobacteria (blue-green bacteria) and rhizobacteria. These enzymes are responsible for the reduction of nitrogen (N2) to ammonia (NH3). Nitrogenases are the only family of enzymes known to catalyze this reaction, which is a step in the process of nitrogen fixation. Nitrogen fixation is required for all forms of life, with nitrogen being essential for the biosynthesis of molecules (nucleotides, amino acids) that create plants, animals and other organisms. They are encoded by the Nif genes or homologs. They are related to protochlorophyllide reductase.

Bioinorganic chemistry is a field that examines the role of metals in biology. Bioinorganic chemistry includes the study of both natural phenomena such as the behavior of metalloproteins as well as artificially introduced metals, including those that are non-essential, in medicine and toxicology. Many biological processes such as respiration depend upon molecules that fall within the realm of inorganic chemistry. The discipline also includes the study of inorganic models or mimics that imitate the behaviour of metalloproteins.

In biochemistry, the metallome is the distribution of metal ions in a cellular compartment. The term was coined in analogy with proteome as metallomics is the study of metallome: the "comprehensive analysis of the entirety of metal and metalloid species within a cell or tissue type". Therefore, metallomics can be considered a branch of metabolomics, even though the metals are not typically considered as metabolites.

<span class="mw-page-title-main">Vanadate</span> Coordination complex of vanadium

In chemistry, a vanadate is an anionic coordination complex of vanadium. Often vanadate refers to oxoanions of vanadium, most of which exist in its highest oxidation state of +5. The complexes [V(CN)6]3− and [V2Cl9]3− are referred to as hexacyanovanadate(III) and nonachlorodivanadate(III), respectively.

<span class="mw-page-title-main">Aryl hydrocarbon receptor</span> Vertebrate transcription factor

The aryl hydrocarbon receptor is a protein that in humans is encoded by the AHR gene. The aryl hydrocarbon receptor is a transcription factor that regulates gene expression. It was originally thought to function primarily as a sensor of xenobiotic chemicals and also as the regulator of enzymes such as cytochrome P450s that metabolize these chemicals. The most notable of these xenobiotic chemicals are aromatic (aryl) hydrocarbons from which the receptor derives its name.

<span class="mw-page-title-main">Composition of the human body</span> Body composition elements

Body composition may be analyzed in various ways. This can be done in terms of the chemical elements present, or by molecular structure e.g., water, protein, fats, hydroxylapatite, carbohydrates and DNA. In terms of tissue type, the body may be analyzed into water, fat, connective tissue, muscle, bone, etc. In terms of cell type, the body contains hundreds of different types of cells, but notably, the largest number of cells contained in a human body are not human cells, but bacteria residing in the normal human gastrointestinal tract.

NAD<sup>+</sup> kinase Enzyme

NAD+ kinase (EC 2.7.1.23, NADK) is an enzyme that converts nicotinamide adenine dinucleotide (NAD+) into NADP+ through phosphorylating the NAD+ coenzyme. NADP+ is an essential coenzyme that is reduced to NADPH primarily by the pentose phosphate pathway to provide reducing power in biosynthetic processes such as fatty acid biosynthesis and nucleotide synthesis. The structure of the NADK from the archaean Archaeoglobus fulgidus has been determined.

<span class="mw-page-title-main">Vanadium nitrogenase</span> Enzyme necessary for the process of nitrogen fixation

Vanadium nitrogenase is a key enzyme for nitrogen fixation found in nitrogen-fixing bacteria, and is used as an alternative to molybdenum nitrogenase when molybdenum is unavailable. Vanadium nitrogenases are an important biological use of vanadium, which is uncommonly used by life. An important component of the nitrogen cycle, vanadium nitrogenase converts nitrogen gas to ammonia, thereby making otherwise inaccessible nitrogen available to plants. Unlike molybdenum nitrogenase, vanadium nitrogenase can also reduce carbon monoxide to ethylene, ethane and propane but both enzymes can reduce protons to hydrogen gas and acetylene to ethylene.

A transition metal oxo complex is a coordination complex containing an oxo ligand. Formally O2-, an oxo ligand can be bound to one or more metal centers, i.e. it can exist as a terminal or (most commonly) as bridging ligands (Fig. 1). Oxo ligands stabilize high oxidation states of a metal. They are also found in several metalloproteins, for example in molybdenum cofactors and in many iron-containing enzymes. One of the earliest synthetic compounds to incorporate an oxo ligand is potassium ferrate (K2FeO4), which was likely prepared by Georg E. Stahl in 1702.

Evolution of metal ions in biological systems refers to the incorporation of metallic ions into living organisms and how it has changed over time. Metal ions have been associated with biological systems for billions of years, but only in the last century have scientists began to truly appreciate the scale of their influence. Major and minor metal ions have become aligned with living organisms through the interplay of biogeochemical weathering and metabolic pathways involving the products of that weathering. The associated complexes have evolved over time.

<span class="mw-page-title-main">Vanadium cycle</span> Exchange of vanadium between continental crust and seawater

The global vanadium cycle is controlled by physical and chemical processes that drive the exchange of vanadium between its two main reservoirs: the upper continental crust and the ocean. Anthropogenic processes such as coal and petroleum production release vanadium to the atmosphere.

<span class="mw-page-title-main">Molybdenum in biology</span> Use of Molybdenum by organisms

Molybdenum is an essential element in most organisms. It is most notably present in nitrogenase which is an essential part of nitrogen fixation.

References

  1. Beaver, William C.; Noland, George B. (1970). General biology; the science of biology. St Louis: Mosby. ISBN   978-0-8016-0544-4.[ page needed ]
  2. Beaver, William C.; Noland, George B. (1970). General biology; the science of biology . St Louis: Mosby. p.  68. ISBN   978-0-8016-0544-4.
  3. 1 2 3 4 5 Abundance of elements in the Earth's crust and in the sea, CRC Handbook of Chemistry and Physics, 97th edition (2016–2017), p. 14-17.
  4. Ultratrace minerals. Authors: Nielsen, Forrest H. USDA, ARS Source: Modern nutrition in health and disease / editors, Maurice E. Shils ... et al. Baltimore: Williams & Wilkins, c1999., p. 283-303. Issue Date: 1999 URI:
  5. Szklarska D, Rzymski P (May 2019). "Is Lithium a Micronutrient? From Biological Activity and Epidemiological Observation to Food Fortification". Biol Trace Elem Res. 189 (1): 18–27. doi:10.1007/s12011-018-1455-2. PMC   6443601 . PMID   30066063.
  6. Enderle J, Klink U, di Giuseppe R, Koch M, Seidel U, Weber K, Birringer M, Ratjen I, Rimbach G, Lieb W (August 2020). "Plasma Lithium Levels in a General Population: A Cross-Sectional Analysis of Metabolic and Dietary Correlates". Nutrients. 12 (8): 2489. doi: 10.3390/nu12082489 . PMC   7468710 . PMID   32824874.
  7. McCall AS, Cummings CF, Bhave G, Vanacore R, Page-McCaw A, Hudson BG (June 2014). "Bromine is an essential trace element for assembly of collagen IV scaffolds in tissue development and architecture". Cell. 157 (6): 1380–92. doi:10.1016/j.cell.2014.05.009. PMC   4144415 . PMID   24906154.
  8. Zoroddu, Maria Antonietta; Aaseth, Jan; Crisponi, Guido; Medici, Serenella; Peana, Massimiliano; Nurchi, Valeria Marina (2019). "The essential metals for humans: a brief overview". Journal of Inorganic Biochemistry. 195: 120–129. doi:10.1016/j.jinorgbio.2019.03.013.
  9. Daumann, Lena J. (25 April 2019). "Essential and Ubiquitous: The Emergence of Lanthanide Metallobiochemistry". Angewandte Chemie International Edition. doi:10.1002/anie.201904090 . Retrieved 15 June 2019.
  10. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 Emsley, John (2003). Nature's building blocks: an A-Z guide to the elements . Oxford: Oxford University Press. ISBN   978-0-19-850340-8.
  11. Exley C. (2013) Aluminum in Biological Systems. In: Kretsinger R.H., Uversky V.N., Permyakov E.A. (eds) Encyclopedia of Metalloproteins. Springer, New York, NY[ page needed ]
  12. Exley, C. (June 2016). "The toxicity of aluminium in humans". Morphologie. 100 (329): 51–55. doi:10.1016/j.morpho.2015.12.003. PMID   26922890.
  13. Bojórquez-Quintal, Emanuel; Escalante-Magaña, Camilo; Echevarría-Machado, Ileana; Martínez-Estévez, Manuel (12 October 2017). "Aluminum, a Friend or Foe of Higher Plants in Acid Soils". Frontiers in Plant Science. 8: 1767. doi: 10.3389/fpls.2017.01767 . PMC   5643487 . PMID   29075280.
  14. 1 2 Guoqing, Zhang Zhipeng Zhong; Qiying, Jiang (2008). "Biological Activities of the Complexes of Arsenic, Antimony and Bismuth [J]". Progress in Chemistry. 9.
  15. 1 2 "Periodic Table of the Elements". Minerals Education Coalition. Minerals Education Coalition. Retrieved 7 April 2020.
  16. Patnaik, Pradyot (2003). Handbook of inorganic chemicals. McGraw-Hill. pp.  77–78. ISBN   978-0-07-049439-8.
  17. OSHA Hazard Information Bulletin HIB 02-04-19 (rev. 05-14-02) Preventing Adverse Health Effects From Exposure to Beryllium in Dental Laboratories
  18. Sun, Hongzhe; Li, Hougyan; Sadler, Peter J. (June 1997). "The Biological and Medicinal Chemistry of Bismuth". Chemische Berichte. 130 (6): 669–681. doi:10.1002/cber.19971300602.
  19. DiPalma, Joseph R. (April 2001). "Bismuth Toxicity, Often Mild, Can Result in Severe Poisonings". Emergency Medicine News. 23 (3): 16. doi:10.1097/00132981-200104000-00012.
  20. Ahmad, Waqar; Niaz, A.; Kanwal, S.; Rahmatullah; Rasheed, M. Khalid (2009). "Role of boron in plant growth: a review". Journal of Agricultural Research. 47 (3): 329–336.
  21. 1 2 3 4 5 6 7 Nielsen, Forrest H. (1984). "Ultratrace elements in nutrition". Annual Review of Nutrition. 4: 21–41. doi:10.1146/annurev.nu.04.070184.000321. PMID   6087860.
  22. Uluisik, Irem; Karakaya, Huseyin Caglar; Koc, Ahmet (1 January 2018). "The importance of boron in biological systems". Journal of Trace Elements in Medicine and Biology. 45: 156–162. doi:10.1016/j.jtemb.2017.10.008. hdl: 11147/7059 . PMID   29173473.
  23. McCall AS; Cummings CF; Bhave G; Vanacore R; Page-McCaw A; et al. (2014). "Bromine Is an Essential Trace Element for Assembly of Collagen IV Scaffolds in Tissue Development and Architecture". Cell. 157 (6): 1380–92. doi:10.1016/j.cell.2014.05.009. PMC   4144415 . PMID   24906154.
  24. Mayeno, A. N.; Curran, A. J.; Roberts, R. L.; Foote, C. S. (5 April 1989). "Eosinophils preferentially use bromide to generate halogenating agents". Journal of Biological Chemistry. 264 (10): 5660–5668. doi: 10.1016/S0021-9258(18)83599-2 . PMID   2538427.
  25. Moore, R. M.; Webb, M.; Tokarczyk, R.; Wever, R. (15 September 1996). "Bromoperoxidase and iodoperoxidase enzymes and production of halogenated methanes in marine diatom cultures". Journal of Geophysical Research: Oceans. 101 (C9): 20899–20908. Bibcode:1996JGR...10120899M. doi:10.1029/96JC01248.
  26. Gribble, Gordon W. (1999). "The diversity of naturally occurring organobromine compounds". Chemical Society Reviews. 28 (5): 335–346. doi:10.1039/A900201D.
  27. Butler, Alison; Carter-Franklin, Jayme N. (2004). "The role of vanadium bromoperoxidase in the biosynthesis of halogenated marine natural products". Natural Product Reports. 21 (1): 180–8. doi:10.1039/b302337k. PMID   15039842. S2CID   19115256.
  28. Lane, Todd W.; Saito, Mak A.; George, Graham N.; Pickering, Ingrid J.; Prince, Roger C.; Morel, François M. M. (4 May 2005). "A cadmium enzyme from a marine diatom". Nature. 435 (7038): 42. doi: 10.1038/435042a . PMID   15875011.
  29. Küpper, Hendrik; Leitenmaier, Barbara (2013). "Cadmium-Accumulating Plants". Cadmium: From Toxicity to Essentiality. Metal Ions in Life Sciences. Vol. 11. pp. 373–393. doi:10.1007/978-94-007-5179-8_12. ISBN   978-94-007-5178-1. PMID   23430779.
  30. Martelli, A.; Rousselet, E.; Dycke, C.; Bouron, A.; Moulis, J.-M. (November 2006). "Cadmium toxicity in animal cells by interference with essential metals". Biochimie. 88 (11): 1807–1814. doi:10.1016/j.biochi.2006.05.013. PMID   16814917.
  31. Brini, Marisa; Calì, Tito; Ottolini, Denis; Carafoli, Ernesto (2013). "Intracellular Calcium Homeostasis and Signaling". Metallomics and the Cell. Metal Ions in Life Sciences. Vol. 12. pp. 119–168. doi:10.1007/978-94-007-5561-1_5. ISBN   978-94-007-5560-4. PMID   23595672.
  32. "Calcium". Linus Pauling Institute, Oregon State University, Corvallis, Oregon. 1 September 2017. Retrieved 31 August 2019.
  33. Vaidyanathan, Gayathri (November 4, 2014). "The Worst Climate Pollution Is Carbon Dioxide". Scientific American. Scientific American. Retrieved 9 April 2020.
  34. 1 2 3 4 Pol, Arjan; Barends, Thomas R. M.; Dietl, Andreas; Khadem, Ahmad F.; Eygensteyn, Jelle; Jetten, Mike S. M.; Op den Camp, Huub J. M. (January 2014). "Rare earth metals are essential for methanotrophic life in volcanic mudpots". Environmental Microbiology. 16 (1): 255–264. doi:10.1111/1462-2920.12249. PMID   24034209.
  35. Venturi, Sebastiano (January 2021). "Cesium in Biology, Pancreatic Cancer, and Controversy in High and Low Radiation Exposure Damage—Scientific, Environmental, Geopolitical, and Economic Aspects". International Journal of Environmental Research and Public Health. 18 (17): 8934. doi: 10.3390/ijerph18178934 . PMC   8431133 . PMID   34501532. CC-BY icon.svg Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  36. Snitynskyĭ, VV; Solohub, LI; Antoniak, HL; Kopachuk, DM; Herasymiv, MH (1999). "[Biological role of chromium in humans and animals]". Ukrains'kyi Biokhimichnyi Zhurnal. 71 (2): 5–9. PMID   10609294.
  37. Castresana J, Lübben M, Saraste M, Higgins DG (June 1994). "Evolution of cytochrome oxidase, an enzyme older than atmospheric oxygen". The EMBO Journal. 13 (11): 2516–25. doi:10.1002/j.1460-2075.1994.tb06541.x. PMC   395125 . PMID   8013452.
  38. Morrison, Jim. "Copper's Virus-Killing Powers Were Known Even to the Ancients". Smithsonian Magazine. Smithsonian Magazine. Retrieved 5 May 2020.
  39. 1 2 3 Haley, Thomas J.; Koste, L.; Komesu, N.; Efros, M.; Upham, H. C. (1966). "Pharmacology and toxicology of dysprosium, holmium, and erbium chlorides". Toxicology and Applied Pharmacology. 8 (1): 37–43. doi:10.1016/0041-008X(66)90098-6. PMID   5921895.
  40. Yeung EW, Allen DG (August 2004). "Stretch-activated channels in stretch-induced muscle damage: role in muscular dystrophy". Clinical and Experimental Pharmacology & Physiology. 31 (8): 551–56. doi:10.1111/j.1440-1681.2004.04027.x. hdl: 10397/30099 . PMID   15298550. S2CID   9550616.
  41. Hayes, Raymond L. (January 1983). "The interaction of gallium with biological systems". International Journal of Nuclear Medicine and Biology. 10 (4): 257–261. doi:10.1016/0047-0740(83)90090-6. PMID   6363324.
  42. 1 2 Lutgen, Pierre (January 23, 2015). "Gallium, key element in the excellent Bamileke Artemisia?". Malaria World. Retrieved 9 April 2020.
  43. Atabaev, Timur; Shin, Yong; Song, Su-Jin; Han, Dong-Wook; Hong, Nguyen (7 August 2017). "Toxicity and T2-Weighted Magnetic Resonance Imaging Potentials of Holmium Oxide Nanoparticles". Nanomaterials. 7 (8): 216. doi: 10.3390/nano7080216 . PMC   5575698 . PMID   28783114.
  44. Bowen, H. J. M. 1979. Environmental chemistry of the elements. London: Academic Press.[ page needed ]
  45. Liebig, George F. Jr; Vanselow, Albert P.; Chapman, H. D. (September 1943). "Effects of gallium and indium on the growth of citrus plants in solution cultures". Soil Science. 56 (3): 173–186. Bibcode:1943SoilS..56..173L. doi:10.1097/00010694-194309000-00002. S2CID   93717588.
  46. Venturi, Sebastiano (1 September 2011). "Evolutionary Significance of Iodine". Current Chemical Biology. 5 (3): 155–162. doi:10.2174/187231311796765012.
  47. Wood, Bruce W.; Grauke, Larry J. (November 2011). "The Rare-earth Metallome of Pecan and Other Carya". Journal of the American Society for Horticultural Science. 136 (6): 389–398. doi: 10.21273/JASHS.136.6.389 .
  48. Law, N.; Caudle, M.; Pecoraro, V. (1998). Manganese Redox Enzymes and Model Systems: Properties, Structures, and Reactivity. Advances in Inorganic Chemistry. Vol. 46. p. 305. doi:10.1016/S0898-8838(08)60152-X. ISBN   978-0-12-023646-6.
  49. Miriyala, Sumitra; K. Holley, Aaron; St Clair, Daret K. (1 February 2011). "Mitochondrial Superoxide Dismutase - Signals of Distinction". Anti-Cancer Agents in Medicinal Chemistry. 11 (2): 181–190. doi:10.2174/187152011795255920. PMC   3427752 . PMID   21355846.
  50. Enemark, John H.; Cooney, J. Jon A.; Wang, Jun-Jieh; Holm, R. H. (2004). "Synthetic Analogues and Reaction Systems Relevant to the Molybdenum and Tungsten Oxotransferases". Chem. Rev. 104 (2): 1175–1200. doi:10.1021/cr020609d. PMID   14871153.
  51. Mendel, Ralf R.; Bittner, Florian (2006). "Cell biology of molybdenum". Biochimica et Biophysica Acta (BBA) - Molecular Cell Research. 1763 (7): 621–635. doi: 10.1016/j.bbamcr.2006.03.013 . PMID   16784786.
  52. Russ Hille; James Hall; Partha Basu (2014). "The Mononuclear Molybdenum Enzymes". Chem. Rev. 114 (7): 3963–4038. doi:10.1021/cr400443z. PMC   4080432 . PMID   24467397.
  53. "Material Safety Data Sheet – Molybdenum". The REMBAR Company, Inc. 2000-09-19. Archived from the original on March 23, 2007. Retrieved 2007-05-13.
  54. "CDC – NIOSH Pocket Guide to Chemical Hazards – Molybdenum". www.cdc.gov. Archived from the original on 2015-11-20. Retrieved 2015-11-20.
  55. Astrid Sigel; Helmut Sigel; Roland K. O. Sigel, eds. (2008). Nickel and Its Surprising Impact in Nature. Metal Ions in Life Sciences. Vol. 2. Wiley. ISBN   978-0-470-01671-8.[ page needed ]
  56. Zamble, Deborah; Rowińska-Żyrek, Magdalena; Kozlowski, Henryk (2017). The Biological Chemistry of Nickel. Royal Society of Chemistry. ISBN   978-1-78262-498-1.[ page needed ]
  57. Rascio, Nicoletta; Navari-Izzo, Flavia (February 2011). "Heavy metal hyperaccumulating plants: How and why do they do it? And what makes them so interesting?". Plant Science. 180 (2): 169–181. doi:10.1016/j.plantsci.2010.08.016. PMID   21421358.
  58. Xu, Jian; Weng, Xiao-Jun; Wang, Xu; Huang, Jia-Zhang; Zhang, Chao; Muhammad, Hassan; Ma, Xin; Liao, Qian-De (19 November 2013). "Potential Use of Porous Titanium–Niobium Alloy in Orthopedic Implants: Preparation and Experimental Study of Its Biocompatibility In Vitro". PLOS ONE. 8 (11): e79289. Bibcode:2013PLoSO...879289X. doi: 10.1371/journal.pone.0079289 . PMC   3834032 . PMID   24260188.
  59. Ramírez, G.; Rodil, S.E.; Arzate, H.; Muhl, S.; Olaya, J.J. (January 2011). "Niobium based coatings for dental implants". Applied Surface Science. 257 (7): 2555–2559. Bibcode:2011ApSS..257.2555R. doi:10.1016/j.apsusc.2010.10.021.
  60. Colon, Pierre; Pradelle-Plasse, Nelly; Galland, Jacques (2003). "Evaluation of the long-term corrosion behavior of dental amalgams: influence of palladium addition and particle morphology". Dental Materials. 19 (3): 232–9. doi:10.1016/S0109-5641(02)00035-0. PMID   12628436.
  61. Chauhan, Reshu; Awasthi, Surabhi; Srivastava, Sudhakar; Dwivedi, Sanjay; Pilon-Smits, Elizabeth A. H.; Dhankher, Om P.; Tripathi, Rudra D. (3 April 2019). "Understanding selenium metabolism in plants and its role as a beneficial element". Critical Reviews in Environmental Science and Technology. 49 (21): 1937–1958. doi:10.1080/10643389.2019.1598240. S2CID   133580188.
  62. Häse, Claudia C.; Fedorova, Natalie D.; Galperin, Michael Y.; Dibrov, Pavel A. (1 September 2001). "Sodium Ion Cycle in Bacterial Pathogens: Evidence from Cross-Genome Comparisons". Microbiology and Molecular Biology Reviews. 65 (3): 353–370. doi:10.1128/MMBR.65.3.353-370.2001. PMC   99031 . PMID   11528000.
  63. Rieder, Norbert; Ott, Hubert A.; Pfundstein, Peter; Schoch, Robert (February 1982). "X-ray Microanalysis of the Mineral Contents of Some Protozoa". The Journal of Protozoology. 29 (1): 15–18. doi:10.1111/j.1550-7408.1982.tb02875.x.
  64. Léonard, A; Gerber, G.B (August 1997). "Mutagenicity, carcinogenicity and teratogenicity of thallium compounds". Mutation Research/Reviews in Mutation Research. 387 (1): 47–53. doi:10.1016/S1383-5742(97)00022-7. PMID   9254892.
  65. Koribanics, N. M.; Tuorto, S. J.; Lopez-Chiaffarelli, N.; McGuinness, L. R.; Häggblom, M. M.; Williams, K. H.; Long, P. E.; Kerkhof, L. J. (2015). "Spatial Distribution of an Uranium-Respiring Betaproteobacterium at the Rifle, CO Field Research Site". PLOS ONE. 10 (4): e0123378. Bibcode:2015PLoSO..1023378K. doi: 10.1371/journal.pone.0123378 . PMC   4395306 . PMID   25874721.
  66. McMaster, J. & Enemark, John H. (1998). "The active sites of molybdenum- and tungsten-containing enzymes". Current Opinion in Chemical Biology. 2 (2): 201–207. doi:10.1016/S1367-5931(98)80061-6. PMID   9667924.
  67. Hille, Russ (2002). "Molybdenum and tungsten in biology". Trends in Biochemical Sciences. 27 (7): 360–367. doi:10.1016/S0968-0004(02)02107-2. PMID   12114025.
  68. Koribanics, Nicole M.; Tuorto, Steven J.; Lopez-Chiaffarelli, Nora; McGuinness, Lora R.; Häggblom, Max M.; Williams, Kenneth H.; Long, Philip E.; Kerkhof, Lee J.; Morais, Paula V (13 April 2015). "Spatial Distribution of an Uranium-Respiring Betaproteobacterium at the Rifle, CO Field Research Site". PLOS ONE. 10 (4): e0123378. Bibcode:2015PLoSO..1023378K. doi: 10.1371/journal.pone.0123378 . PMC   4395306 . PMID   25874721.
  69. 1 2 Chatterjee, Malay; Das, Subhadeep; Chatterjee, Mary; Roy, Kaushik (2013). "Vanadium in Biological Systems". Encyclopedia of Metalloproteins. pp. 2293–2297. doi:10.1007/978-1-4614-1533-6_134. ISBN   978-1-4614-1532-9.
  70. Bishop, P E; Joerger, R D (June 1990). "Genetics and Molecular Biology of Alternative Nitrogen Fixation Systems". Annual Review of Plant Physiology and Plant Molecular Biology. 41 (1): 109–125. doi:10.1146/annurev.pp.41.060190.000545.
  71. Wever, R.; Krenn, B. E. (1990). "Vanadium Haloperoxidases". Vanadium in Biological Systems. pp. 81–97. doi:10.1007/978-94-009-2023-1_5. ISBN   978-94-010-7407-0.