CO-methylating acetyl-CoA synthase

Last updated
CO-methylating acetyl-CoA synthase
PDB 1ru3 EBI.jpg
Monomeric Acetyl-CoA synthase
Identifiers
EC no. 2.3.1.169
Databases
IntEnz IntEnz view
BRENDA BRENDA entry
ExPASy NiceZyme view
KEGG KEGG entry
MetaCyc metabolic pathway
PRIAM profile
PDB structures RCSB PDB PDBe PDBsum
Gene Ontology AmiGO / QuickGO
Search
PMC articles
PubMed articles
NCBI proteins

Acetyl-CoA synthase (ACS), not to be confused with acetyl-CoA synthetase or acetate-CoA ligase (ADP forming), is a nickel-containing enzyme involved in the metabolic processes of cells. Together with carbon monoxide dehydrogenase (CODH), it forms the bifunctional enzyme Acetyl-CoA Synthase/Carbon Monoxide Dehydrogenase (ACS/CODH) found in anaerobic microorganisms such as archaea and bacteria. [1] The ACS/CODH enzyme works primarily through the Wood–Ljungdahl pathway which converts carbon dioxide to Acetyl-CoA. The recommended name for this enzyme is CO-methylating acetyl-CoA synthase. [2]

Contents

Chemistry

In nature, there are six different pathways where CO2 is fixed. Of these, the Wood–Ljungdahl pathway is the predominant sink in anaerobic conditions. Acetyl-CoA Synthase (ACS) and carbon monoxide dehydrogenase (CODH) are integral enzymes in this one pathway and can perform diverse reactions in the carbon cycle as a result. Because of this, the exact activity of these molecules has come under intense scrutiny over the past decade. [3]

Wood–Ljungdahl pathway

The Wood–Ljungdahl pathway consists of two different reactions that break down carbon dioxide. The first pathway involves CODH converting carbon dioxide into carbon monoxide through a two-electron transfer, and the second reaction involves ACS synthesizing acetyl-CoA using the carbon monoxide from CODH together with coenzyme-A (CoA) and a methyl group from a corrinoid iron-sulfur protein, CFeSP. [4] The two main overall reactions are as follows:

 

 

 

 

(1)

 

 

 

 

(2)

The acetyl-CoA produced can be used in a variety of ways depending on the needs of the organism. For example, acetate-forming bacteria use acetyl-CoA for their autotrophic growth processes, and methanogenic archae such as Methanocarcina barkeri convert the acetyl-CoA into acetate and use it as an alternative source of carbon instead of CO2. [5]

Autotrophic growth via the Wood-Ljungdahl pathway. Adapted from Ragsdale et al. ACS Autotrophic growth and acetate synthesis.jpg
Autotrophic growth via the Wood–Ljungdahl pathway. Adapted from Ragsdale et al.

Acetogenic bacteria use this method to generate acetate and acetic acid. Since the above two reactions are reversible, it opens up a diverse range of reactions in the carbon cycle. In addition to acetyl-CoA production, the reverse can occur with ACS producing CO and returning the methyl piece back to the corrinoid protein.

Along with the process of methanogenesis, organisms can subsequently convert the acetate to methane. Furthermore, the Wood-Ljungdahl pathway allows for the anaerobic oxidation of acetate where ATP is used to convert acetate into acetyl-CoA, which is then broken down by ACS to produce carbon dioxide that is released into the atmosphere. [6]

Other reactions

It has been discovered that the CODH/ACS enzyme in the bacteria M. theroaceticum can make dinitrogen (N2) from nitrous oxide in the presence of an electron-donating species. It can also catalyze the reduction of the pollutant, 2,4,6-trinitrotoluene (TNT) and catalyze the oxidation of n-butyl isocyanide. [3]

Structure

History

The first, and one of the most comprehensive, crystal structures of ACS/CODH from the bacteria M. thermoacetica was presented in 2002 by Drennan and colleagues. [7] In this paper they constructed a heterotetramer, with the active site "A-cluster" residing in the ACS subunit and the active site "C-cluster" in CODH subunit. Furthermore, they resolved the structure of the A-cluster active site and found an [Fe4S4]-X-Cu-X-Ni centre which is highly unusual in biology. This structural representation consisted of a [Fe4S4] unit bridged to a binuclear centre, where Ni(II) resided in the distal position (denoted as Nid) in a square-planar conformation and a Cu(I) ion resided in the proximal position in a distorted tetrahedral position with ligands of unknown identity. [7]

The debate towards the absolute structure and identity of the metals in the A-cluster active site of ACS continued, with a competing model presented. The authors suggested two different forms of the ACS enzyme, an "Open" form and a "Closed" form, with different metals occupying the proximal metal site (denoted as Mp) for each form. The general scheme of the enzyme followed closely with the first study's findings, but this new structure proposed a nickel ion in the "open" form and a zinc ion in the "closed" form. [4]

A later review article attempted to reconcile the different observations of Mp and stated that this proximal position in the active site of ACS was prone to substitution and could contain any one of Cu, Zn and Ni. The three forms of this A-cluster most likely hold a small amount of Ni and a relatively larger amount of Cu. [8]

Present (2014 onwards)

Bifunctional CODH/ACS unit in M.thermoacetica PDB 3I01.jpg
Bifunctional CODH/ACS unit in M.thermoacetica

It is now generally accepted that the ACS active site (A-cluster) is a Ni-Ni metal centre with both nickels having a +2 oxidation state. The [Fe4S4] cluster is bridged to the closer nickel, Np which is connected via a thiolate bridge to the farther nickel, Nid. Nid is coordinated to two cysteine molecules and two backbone amide compounds, and is in a square-planar coordination. The space next to the metal can accommodate substrates and products. Nip is in a T-shaped environment bound to three sulfur atoms, with an unknown ligand possibly creating a distorted tetrahedral environment. This ligand has been hypothesized to be a water molecule or an acetyl group in the surrounding area in the cell. Although the proximal nickel is labile and can be replaced with a Cu of Zn centre, experimental evidence suggests that activity of ACS is limited to the presence of nickel only. In addition, some studies have shown that copper can even inhibit the enzyme under certain conditions. [9]

The overall structure of the CODH/ACS enzyme consists of the CODH enzyme as a dimer at the centre with two ACS subunits on each side. The CODH core is made up of two Ni-Fe-S clusters (C-cluster), two [Fe4S4] clusters (B-cluster) and one [Fe4S4] D-cluster. The D-cluster bridges the two subunits with one C and one B cluster in each monomer, allowing rapid electron transfer. The A-cluster of ACS is in constant communication with the C-cluster in CODH. This active site is also responsible for the C-C and C-S bond formations in the product acetyl-CoA (and its reverse reaction). [8]

Furthermore, the crystal structures of the CODH/ACS complex of Carboxydothermus hydrogenoformans [10] and Clostridium autoethanogenum [11] have been solved. While the latter shows a more extended arrangement of the ACS subunits, the complex of C. hydrogenoformans is very similar to that of M. thermoacetica.

Acetyl-CoA Synthase active site structure ACS A cluster active site.jpg
Acetyl-CoA Synthase active site structure

The ACS enzyme contains three main subunits. The first is the active site itself with the NiFeS centre. The second is the portion that directly interacts with CODH in the Wood–Ljungdahl pathway. This part is made up of α-helices that go into a Rossmann fold. It also appears to interact with a ferredoxin compound which may activate the subunit during the CO transferring process from CODH to ACS. The final domain binds CoA and consists of six arginine residues with a tryptophan molecule. [3] [12]

Experiments between the C-cluster of CODH and the A-cluster of ACS reveal a long, hydrophobic channel connecting the two domains to allow for the transfer of carbon monoxide from CODH to ACS. This channel is most likely to protect the carbon monoxide molecules from the outside environment of the enzyme and to increase efficiency of acetyl-CoA production. [13]

Conformational changes

Studies in literature have been able to isolate the CODH/ACS enzyme in an "open" and "closed" configuration. This has led to the hypothesis that it undergoes four conformational changes depending on its activity. With the "open" position, the active site rotates itself to interact with the CFeSP protein in the methyl transfer step of the Wood–Ljungdahl pathway. The "closed" position opens up the channel between CODH and ACS to allow for the transfer of CO. These two configurations are opposite one another in that access to CO blocks off interaction with CFeSP, and when methylation occurs, the active site is buried and does not allow CO transfer. A second "closed" position is needed to block off water from the reaction. Finally, the A-cluster must be rotated once more to allow for the binding of CoA and release of the product. The exact trigger of these structural changes and the mechanistic details have yet to be resolved. [3] [6] [9]

Activity

Mechanism

Proposed diamagnetic (top) and paramagnetic (bottom) mechanisms. Adapted from Seravalli et al. ACS paramagetic and diamagnetic mechanisms.jpg
Proposed diamagnetic (top) and paramagnetic (bottom) mechanisms. Adapted from Seravalli et al.

Two competing mechanisms have been proposed for the formation of acetyl-CoA, the "paramagnetic mechanism" and the "diamagnetic mechanism". [3] Both are similar in terms of the binding of substrates and the general steps, but differ in the oxidation state of the metal centre. Nip is believed to be the substrate binding centre which undergoes redox. The farther nickel centre and the [Fe4S4] cluster are not thought to be involved in the process. [13]

In the paramagnetic mechanism, some type of complex (ferrodoxin, for example) activates the Nip atom, reducing it from Ni2+ to Ni1+. The nickel then binds to either carbon monoxide from CODH or the methyl group donated by the CFeSP protein in no particular order. [14] This is followed by migratory insertion to form an intermediate complex. CoA then binds to the metal and the final product, acetyl-CoA, is formed. [3] [9] Some criticisms of this mechanism are that it is unbalanced in terms of electron count and the activated Ni1+ intermediate cannot be detected with electron paramagnetic resonance. Furthermore, there is evidence of the ACS catalytic cycle without any external reducing complex, which refutes the ferrodoxin activation step. [15]

The second proposed mechanism, the diamagnetic mechanism, involves a Ni0 intermediate instead of a Ni1+. After addition of the methyl group and carbon monoxide, followed by insertion to produce the metal-acetyl complex, CoA attacks to produce the final product. [9] The order in which the carbon monoxide molecule and the methyl group bind to the nickel centre has been highly debated, but no solid evidence has demonstrated preference for one over the other. Although this mechanism is electronically balanced, the idea of a Ni0 species is highly unprecedented in biology. There has also been no solid evidence supporting the presence of a zerovalent Ni species. However, similar nickel species to ACS with a Ni0 centre have been made, so the diamagnetic mechanism is not an implausible hypothesis. [1]

Related Research Articles

<span class="mw-page-title-main">Citric acid cycle</span> Interconnected biochemical reactions releasing energy

The citric acid cycle—also known as the Krebs cycle, Szent-Györgyi-Krebs cycle or the TCA cycle (tricarboxylic acid cycle)—is a series of biochemical reactions to release the energy stored in nutrients through the oxidation of acetyl-CoA derived from carbohydrates, fats, and proteins. The chemical energy released is available under the form of ATP. The Krebs cycle is used by organisms that respire (as opposed to organisms that ferment) to generate energy, either by anaerobic respiration or aerobic respiration. In addition, the cycle provides precursors of certain amino acids, as well as the reducing agent NADH, that are used in numerous other reactions. Its central importance to many biochemical pathways suggests that it was one of the earliest components of metabolism. Even though it is branded as a 'cycle', it is not necessary for metabolites to follow only one specific route; at least three alternative segments of the citric acid cycle have been recognized.

<span class="mw-page-title-main">Acetyl-CoA</span> Chemical compound

Acetyl-CoA is a molecule that participates in many biochemical reactions in protein, carbohydrate and lipid metabolism. Its main function is to deliver the acetyl group to the citric acid cycle to be oxidized for energy production.

<span class="mw-page-title-main">Biological carbon fixation</span> Series of interconnected biochemical reactions

Biological carbon fixation or сarbon assimilation is the process by which inorganic carbon is converted to organic compounds by living organisms. The compounds are then used to store energy and as structure for other biomolecules. Carbon is primarily fixed through photosynthesis, but some organisms use a process called chemosynthesis in the absence of sunlight.

Iron–sulfur proteins are proteins characterized by the presence of iron–sulfur clusters containing sulfide-linked di-, tri-, and tetrairon centers in variable oxidation states. Iron–sulfur clusters are found in a variety of metalloproteins, such as the ferredoxins, as well as NADH dehydrogenase, hydrogenases, coenzyme Q – cytochrome c reductase, succinate – coenzyme Q reductase and nitrogenase. Iron–sulfur clusters are best known for their role in the oxidation-reduction reactions of electron transport in mitochondria and chloroplasts. Both Complex I and Complex II of oxidative phosphorylation have multiple Fe–S clusters. They have many other functions including catalysis as illustrated by aconitase, generation of radicals as illustrated by SAM-dependent enzymes, and as sulfur donors in the biosynthesis of lipoic acid and biotin. Additionally, some Fe–S proteins regulate gene expression. Fe–S proteins are vulnerable to attack by biogenic nitric oxide, forming dinitrosyl iron complexes. In most Fe–S proteins, the terminal ligands on Fe are thiolate, but exceptions exist.

An acetogen is a microorganism that generates acetate (CH3COO) as an end product of anaerobic respiration or fermentation. However, this term is usually employed in a narrower sense only to those bacteria and archaea that perform anaerobic respiration and carbon fixation simultaneously through the reductive acetyl coenzyme A (acetyl-CoA) pathway (also known as the Wood-Ljungdahl pathway). These genuine acetogens are also known as "homoacetogens" and they can produce acetyl-CoA (and from that, in most cases, acetate as the end product) from two molecules of carbon dioxide (CO2) and four molecules of molecular hydrogen (H2). This process is known as acetogenesis, and is different from acetate fermentation, although both occur in the absence of molecular oxygen (O2) and produce acetate. Although previously thought that only bacteria are acetogens, some archaea can be considered to be acetogens.

Acetogenesis is a process through which acetate is produced by prokaryote microorganisms either by the reduction of CO2 or by the reduction of organic acids, rather than by the oxidative breakdown of carbohydrates or ethanol, as with acetic acid bacteria.

<span class="mw-page-title-main">Glyoxylate cycle</span> Series of interconnected biochemical reactions

The glyoxylate cycle, a variation of the tricarboxylic acid cycle, is an anabolic pathway occurring in plants, bacteria, protists, and fungi. The glyoxylate cycle centers on the conversion of acetyl-CoA to succinate for the synthesis of carbohydrates. In microorganisms, the glyoxylate cycle allows cells to use two carbons, such as acetate, to satisfy cellular carbon requirements when simple sugars such as glucose or fructose are not available. The cycle is generally assumed to be absent in animals, with the exception of nematodes at the early stages of embryogenesis. In recent years, however, the detection of malate synthase (MS) and isocitrate lyase (ICL), key enzymes involved in the glyoxylate cycle, in some animal tissue has raised questions regarding the evolutionary relationship of enzymes in bacteria and animals and suggests that animals encode alternative enzymes of the cycle that differ in function from known MS and ICL in non-metazoan species.

<span class="mw-page-title-main">Methylcobalamin</span> Form of vitamin B12

Methylcobalamin (mecobalamin, MeCbl, or MeB12) is a cobalamin, a form of vitamin B12. It differs from cyanocobalamin in that the cyano group at the cobalt is replaced with a methyl group. Methylcobalamin features an octahedral cobalt(III) centre and can be obtained as bright red crystals. From the perspective of coordination chemistry, methylcobalamin is notable as a rare example of a compound that contains metal–alkyl bonds. Nickel–methyl intermediates have been proposed for the final step of methanogenesis.

Ethanol, an alcohol found in nature and in alcoholic drinks, is metabolized through a complex catabolic metabolic pathway. In humans, several enzymes are involved in processing ethanol first into acetaldehyde and further into acetic acid and acetyl-CoA. Once acetyl-CoA is formed, it becomes a substrate for the citric acid cycle ultimately producing cellular energy and releasing water and carbon dioxide. Due to differences in enzyme presence and availability, human adults and fetuses process ethanol through different pathways. Gene variation in these enzymes can lead to variation in catalytic efficiency between individuals. The liver is the major organ that metabolizes ethanol due to its high concentration of these enzymes.

Bioorganometallic chemistry is the study of biologically active molecules that contain carbon directly bonded to metals or metalloids. The importance of main-group and transition-metal centers has long been recognized as important to the function of enzymes and other biomolecules. However, only a small subset of naturally-occurring metal complexes and synthetically prepared pharmaceuticals are organometallic; that is, they feature a direct covalent bond between the metal(loid) and a carbon atom. The first, and for a long time, the only examples of naturally occurring bioorganometallic compounds were the cobalamin cofactors (vitamin B12) in its various forms. In the 21st century, as a result of the discovery of new systems containing carbon–metal bonds in biology, bioorganometallic chemistry is rapidly emerging as a distinct subdiscipline of bioinorganic chemistry that straddles organometallic chemistry and biochemistry. Naturally occurring bioorganometallics include enzymes and sensor proteins. Also within this realm are synthetically prepared organometallic compounds that serve as new drugs and imaging agents (technetium-99m sestamibi) as well as the principles relevant to the toxicology of organometallic compounds (e.g., methylmercury). Consequently, bioorganometallic chemistry is increasingly relevant to medicine and pharmacology.

In biochemistry, fatty acid synthesis is the creation of fatty acids from acetyl-CoA and NADPH through the action of enzymes called fatty acid synthases. This process takes place in the cytoplasm of the cell. Most of the acetyl-CoA which is converted into fatty acids is derived from carbohydrates via the glycolytic pathway. The glycolytic pathway also provides the glycerol with which three fatty acids can combine to form triglycerides, the final product of the lipogenic process. When only two fatty acids combine with glycerol and the third alcohol group is phosphorylated with a group such as phosphatidylcholine, a phospholipid is formed. Phospholipids form the bulk of the lipid bilayers that make up cell membranes and surrounds the organelles within the cells. In addition to cytosolic fatty acid synthesis, there is also mitochondrial fatty acid synthesis (mtFASII), in which malonyl-CoA is formed from malonic acid with the help of malonyl-CoA synthetase (ACSF3), which then becomes the final product octanoyl-ACP (C8) via further intermediate steps.

Acetyl-CoA synthetase (ACS) or Acetate—CoA ligase is an enzyme involved in metabolism of acetate. It is in the ligase class of enzymes, meaning that it catalyzes the formation of a new chemical bond between two large molecules.

<span class="mw-page-title-main">Wood–Ljungdahl pathway</span> A set of biochemical reactions used by some bacteria

The Wood–Ljungdahl pathway is a set of biochemical reactions used by some bacteria. It is also known as the reductive acetyl-coenzyme A (acetyl-CoA) pathway. This pathway enables these organisms to use hydrogen as an electron donor, and carbon dioxide as an electron acceptor and as a building block for biosynthesis.

In enzymology, carbon monoxide dehydrogenase (CODH) (EC 1.2.7.4) is an enzyme that catalyzes the chemical reaction

In enzymology, a carbon-monoxide dehydrogenase (ferredoxin) (EC 1.2.7.4) is an enzyme that catalyzes the chemical reaction

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

In enzymology, a pyruvate synthase is an enzyme that catalyzes the interconversion of pyruvate and acetyl-CoA. It is also called pyruvate:ferredoxin oxidoreductase (PFOR).

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

The enzyme benzylsuccinate synthase catalyzes the chemical reaction

<span class="mw-page-title-main">Fatty-acyl-CoA synthase</span>

Fatty-acyl-CoA Synthase, or more commonly known as yeast fatty acid synthase, is an enzyme complex responsible for fatty acid biosynthesis, and is of Type I Fatty Acid Synthesis (FAS). Yeast fatty acid synthase plays a pivotal role in fatty acid synthesis. It is a 2.6 MDa barrel shaped complex and is composed of two, unique multi-functional subunits: alpha and beta. Together, the alpha and beta units are arranged in an α6β6 structure. The catalytic activities of this enzyme complex involves a coordination system of enzymatic reactions between the alpha and beta subunits. The enzyme complex therefore consists of six functional centers for fatty acid synthesis.

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

In enzymology, a malate synthase (EC 2.3.3.9) is an enzyme that catalyzes the chemical reaction

[NiFe] hydrogenase is a type of hydrogenase, which is an oxidative enzyme that reversibly converts molecular hydrogen in prokaryotes including Bacteria and Archaea. The catalytic site on the enzyme provides simple hydrogen-metabolizing microorganisms a redox mechanism by which to store and utilize energy via the reaction

References

  1. 1 2 Lindahl PA (July 2004). "Acetyl-coenzyme A synthase: the case for a Ni(p)(0)-based mechanism of catalysis". Journal of Biological Inorganic Chemistry. 9 (5): 516–24. doi:10.1007/s00775-004-0564-x. PMID   15221478. S2CID   21597571.
  2. Springer handbook of enzymes. Vol. 30. pp. 459–466.
  3. 1 2 3 4 5 6 7 Can M, Armstrong FA, Ragsdale SW (April 2014). "Structure, function, and mechanism of the nickel metalloenzymes, CO dehydrogenase, and acetyl-CoA synthase". Chemical Reviews. 114 (8): 4149–74. doi:10.1021/cr400461p. PMC   4002135 . PMID   24521136.
  4. 1 2 Hegg EL (October 2004). "Unraveling the structure and mechanism of acetyl-coenzyme A synthase". Accounts of Chemical Research. 37 (10): 775–83. doi:10.1021/ar040002e. PMID   15491124.
  5. Riordan CG (July 2004). "Synthetic chemistry and chemical precedents for understanding the structure and function of acetyl coenzyme A synthase". Journal of Biological Inorganic Chemistry. 9 (5): 542–9. doi:10.1007/s00775-004-0567-7. PMID   15221481. S2CID   6536992.
  6. 1 2 Ragsdale SW, Kumar M (January 1996). "Nickel-Containing Carbon Monoxide Dehydrogenase/Acetyl-CoA Synthase". Chemical Reviews. 96 (7): 2515–2540. doi:10.1021/cr950058. PMID   11848835.
  7. 1 2 Doukov TI, Iverson TM, Seravalli J, Ragsdale SW, Drennan CL (October 2002). "A Ni-Fe-Cu center in a bifunctional carbon monoxide dehydrogenase/acetyl-CoA synthase". Science. 298 (5593): 567–72. Bibcode:2002Sci...298..567D. doi:10.1126/science.1075843. PMID   12386327. S2CID   39880131.
  8. 1 2 Drennan CL, Doukov TI, Ragsdale SW (July 2004). "The metalloclusters of carbon monoxide dehydrogenase/acetyl-CoA synthase: a story in pictures". Journal of Biological Inorganic Chemistry. 9 (5): 511–5. doi:10.1007/s00775-004-0563-y. PMID   15221484. S2CID   23263180.
  9. 1 2 3 4 Evans DJ (2005). "Chemistry relating to the nickel enzymes CODH and ACS". Coordination Chemistry Reviews. 249 (15–16): 1582–1595. doi:10.1016/j.ccr.2004.09.012.
  10. Ruickoldt, Jakob; Basak, Yudhajeet; Domnik, Lilith; Jeoung, Jae-Hun; Dobbek, Holger (2022-10-21). "On the Kinetics of CO 2 Reduction by Ni, Fe-CO Dehydrogenases". ACS Catalysis. 12 (20): 13131–13142. doi: 10.1021/acscatal.2c02221 . ISSN   2155-5435.
  11. Lemaire, Olivier N.; Wagner, Tristan (January 2021). "Gas channel rerouting in a primordial enzyme: Structural insights of the carbon-monoxide dehydrogenase/acetyl-CoA synthase complex from the acetogen Clostridium autoethanogenum". Biochimica et Biophysica Acta (BBA) - Bioenergetics. 1862 (1): 148330. doi: 10.1016/j.bbabio.2020.148330 . hdl: 21.11116/0000-0007-F1AD-6 .
  12. Kung Y, Drennan CL (April 2011). "A role for nickel-iron cofactors in biological carbon monoxide and carbon dioxide utilization" (PDF). Current Opinion in Chemical Biology. 15 (2): 276–83. doi:10.1016/j.cbpa.2010.11.005. PMC   3061974 . PMID   21130022.
  13. 1 2 Boer JL, Mulrooney SB, Hausinger RP (February 2014). "Nickel-dependent metalloenzymes". Archives of Biochemistry and Biophysics. 544: 142–52. doi:10.1016/j.abb.2013.09.002. PMC   3946514 . PMID   24036122.
  14. 1 2 Seravalli J, Ragsdale SW (March 2008). "Pulse-chase studies of the synthesis of acetyl-CoA by carbon monoxide dehydrogenase/acetyl-CoA synthase: evidence for a random mechanism of methyl and carbonyl addition". The Journal of Biological Chemistry. 283 (13): 8384–94. doi: 10.1074/jbc.M709470200 . PMC   2820341 . PMID   18203715.
  15. Sigel A, Sigel H, Sigel RK (2006). Nickel and its surprising impact in nature. Chichester, West Sussex, England: Wiley. pp. 377–380. ISBN   978-0-470-01671-8.