Cofactor (biochemistry)

Source: Wikipedia, the free encyclopedia.
iron–sulfur centers, and heme
.

A cofactor is a non-protein chemical compound or metallic ion that is required for an enzyme's role as a catalyst (a catalyst is a substance that increases the rate of a chemical reaction). 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.

Cofactors can be classified into two types:

essential nutrients in small amounts. (Some scientists limit the use of the term "cofactor" for inorganic substances; both types are included here.[2][3]
)

Coenzymes are further divided into two types. The first is called a "prosthetic group", which consists of a coenzyme that is tightly (or even covalently) and permanently bound to a protein.

holoenzyme.[5][page needed] (The International Union of Pure and Applied Chemistry (IUPAC) defines "coenzyme" a little differently, namely as a low-molecular-weight, non-protein organic compound that is loosely attached, participating in enzymatic reactions as a dissociable carrier of chemical groups or electrons; a prosthetic group is defined as a tightly bound, nonpolypeptide unit in a protein that is regenerated in each enzymatic turnover.[6]
)

Some enzymes or enzyme complexes require several cofactors. For example, the multienzyme complex pyruvate dehydrogenase[7] at the junction of glycolysis and the citric acid cycle requires five organic cofactors and one metal ion: loosely bound thiamine pyrophosphate (TPP), covalently bound lipoamide and flavin adenine dinucleotide (FAD), cosubstrates nicotinamide adenine dinucleotide (NAD+) and coenzyme A (CoA), and a metal ion (Mg2+).[8]

Organic cofactors are often

RNA world. It has been suggested that the AMP part of the molecule can be considered to be a kind of "handle" by which the enzyme can "grasp" the coenzyme to switch it between different catalytic centers.[9]

Classification

Cofactors can be divided into two major groups: organic cofactors, such as flavin or heme; and inorganic cofactors, such as the metal ions Mg2+, Cu+, Mn2+ and iron–sulfur clusters.

Organic cofactors are sometimes further divided into coenzymes and prosthetic groups. The term coenzyme refers specifically to enzymes and, as such, to the functional properties of a protein. On the other hand, "prosthetic group" emphasizes the nature of the binding of a cofactor to a protein (tight or covalent) and, thus, refers to a structural property. Different sources give slightly different definitions of coenzymes, cofactors, and prosthetic groups. Some consider tightly bound organic molecules as prosthetic groups and not as coenzymes, while others define all non-protein organic molecules needed for enzyme activity as coenzymes, and classify those that are tightly bound as coenzyme prosthetic groups. These terms are often used loosely.

A 1980 letter in Trends in Biochemistry Sciences noted the confusion in the literature and the essentially arbitrary distinction made between prosthetic groups and coenzymes group and proposed the following scheme. Here, cofactors were defined as an additional substance apart from protein and

substrate that is required for enzyme activity and a prosthetic group as a substance that undergoes its whole catalytic cycle attached to a single enzyme molecule. However, the author could not arrive at a single all-encompassing definition of a "coenzyme" and proposed that this term be dropped from use in the literature.[10]

Inorganic cofactors

Metal ions

thyroid hormones rather than as an enzyme cofactor.[15] Calcium is another special case, in that it is required as a component of the human diet, and it is needed for the full activity of many enzymes, such as nitric oxide synthase, protein phosphatases, and adenylate kinase, but calcium activates these enzymes in allosteric regulation, often binding to these enzymes in a complex with calmodulin.[16] Calcium is, therefore, a cell signaling molecule, and not usually considered a cofactor of the enzymes it regulates.[17]

Other organisms require additional metals as enzyme cofactors, such as vanadium in the nitrogenase of the nitrogen-fixing bacteria of the genus Azotobacter,[18] tungsten in the aldehyde ferredoxin oxidoreductase of the thermophilic archaean Pyrococcus furiosus,[19] and even cadmium in the carbonic anhydrase from the marine diatom Thalassiosira weissflogii.[20][21]

In many cases, the cofactor includes both an inorganic and organic component. One diverse set of examples is the heme proteins, which consist of a porphyrin ring coordinated to iron.[22]

Ion Examples of enzymes containing this ion
Cupric
Cytochrome oxidase
Ferrous or Ferric Catalase
Cytochrome (via Heme)
Nitrogenase
Hydrogenase
Magnesium Glucose 6-phosphatase
Hexokinase
DNA polymerase
Manganese Arginase
Molybdenum Nitrate reductase
Nitrogenase
Xanthine oxidase
Nickel Urease
Zinc Alcohol dehydrogenase
Carbonic anhydrase
DNA polymerase
A simple [Fe2S2] cluster containing two iron atoms and two sulfur atoms, coordinated by four protein cysteine residues.

Iron–sulfur clusters

Iron–sulfur clusters are complexes of iron and sulfur atoms held within proteins by cysteinyl residues. They play both structural and functional roles, including electron transfer, redox sensing, and as structural modules.[23]

Organic

Organic cofactors are small organic molecules (typically a molecular mass less than 1000 Da) that can be either loosely or tightly bound to the enzyme and directly participate in the reaction.[5][24][25][26] In the latter case, when it is difficult to remove without denaturing the enzyme, it can be called a prosthetic group. It is important to emphasize that there is no sharp division between loosely and tightly bound cofactors.[5] Indeed, many such as NAD+ can be tightly bound in some enzymes, while it is loosely bound in others.[5] Another example is thiamine pyrophosphate (TPP), which is tightly bound in transketolase or pyruvate decarboxylase, while it is less tightly bound in pyruvate dehydrogenase.[27] Other coenzymes, flavin adenine dinucleotide (FAD), biotin, and lipoamide, for instance, are tightly bound.[28] Tightly bound cofactors are, in general, regenerated during the same reaction cycle, while loosely bound cofactors can be regenerated in a subsequent reaction catalyzed by a different enzyme. In the latter case, the cofactor can also be considered a substrate or cosubstrate.

acyl groups. Most of these cofactors are found in a huge variety of species, and some are universal to all forms of life. An exception to this wide distribution is a group of unique cofactors that evolved in methanogens, which are restricted to this group of archaea.[30]

Vitamins and derivatives

Cofactor Vitamin Additional component Chemical group(s) transferred Distribution
Thiamine pyrophosphate[31] Thiamine (B1) pyrophosphate 2-carbon groups, α cleavage Bacteria, archaea and eukaryotes
NAD+ and NADP+[32]
Niacin
(B3)
ADP Electrons Bacteria, archaea and eukaryotes
Pyridoxal phosphate[33] Pyridoxine (B6) None Amino and carboxyl groups Bacteria, archaea and eukaryotes
Methylcobalamin[34] Vitamin B12 Methyl group
acyl groups
Bacteria, archaea and eukaryotes
Cobalamine[5]
Cobalamine
(B12)
None
alkyl groups
Bacteria, archaea and eukaryotes
Biotin[35] Biotin (H) None CO2 Bacteria, archaea and eukaryotes
Coenzyme A[36] Pantothenic acid (B5) ADP
acyl groups
Bacteria, archaea and eukaryotes
Tetrahydrofolic acid[37]
Folic acid
(B9)
Glutamate residues
Methyl, formyl, methylene
and formimino groups
Bacteria, archaea and eukaryotes
Menaquinone[38] Vitamin K None
Carbonyl group and electrons
Bacteria, archaea and eukaryotes
Ascorbic acid[39]
Vitamin C None Electrons Bacteria, archaea and eukaryotes
Flavin mononucleotide[40] Riboflavin (B2) None Electrons Bacteria, archaea and eukaryotes
Flavin adenine dinucleotide[40] Riboflavin (B2) ADP Electrons Bacteria, archaea and eukaryotes
Coenzyme F420[41] Riboflavin (B2) Amino acids Electrons Methanogens and some bacteria

Non-vitamins

Cofactor Chemical group(s) transferred Distribution
Adenosine triphosphate[42] Phosphate group Bacteria, archaea and eukaryotes
S-Adenosyl methionine[43] Methyl group Bacteria, archaea and eukaryotes
Coenzyme B[44] Electrons Methanogens
Coenzyme M[45][46] Methyl group Methanogens
Coenzyme Q[47]
Electrons Bacteria, archaea and eukaryotes
Cytidine triphosphate[48]
Diacylglycerols
and lipid head groups
Bacteria, archaea and eukaryotes
Glutathione[49][50] Electrons Some bacteria and most eukaryotes
Heme[51] Electrons Bacteria, archaea and eukaryotes
Lipoamide[5]
acyl groups
Bacteria, archaea and eukaryotes
Methanofuran[52] Formyl group Methanogens
Molybdopterin[53][54] Oxygen atoms Bacteria, archaea and eukaryotes
Nucleotide sugars[55] Monosaccharides Bacteria, archaea and eukaryotes
3'-Phosphoadenosine-5'-phosphosulfate[56] Sulfate group Bacteria, archaea and eukaryotes
Pyrroloquinoline quinone[57] Electrons Bacteria
Tetrahydrobiopterin[58] Oxygen atom and electrons Bacteria, archaea and eukaryotes
Tetrahydromethanopterin[59] Methyl group Methanogens

Cofactors as metabolic intermediates

The redox reactions of nicotinamide adenine dinucleotide.

Metabolism involves a vast array of chemical reactions, but most fall under a few basic types of reactions that involve the transfer of functional groups.[60] This common chemistry allows cells to use a small set of metabolic intermediates to carry chemical groups between different reactions.[61] These group-transfer intermediates are the loosely bound organic cofactors, often called coenzymes.

Each class of group-transfer reaction is carried out by a particular cofactor, which is the substrate for a set of enzymes that produce it, and a set of enzymes that consume it. An example of this are the

reductases in the cell that require electrons to reduce their substrates.[32]

Therefore, these cofactors are continuously recycled as part of metabolism. As an example, the total quantity of ATP in the human body is about 0.1 mole. This ATP is constantly being broken down into ADP, and then converted back into ATP. Thus, at any given time, the total amount of ATP + ADP remains fairly constant. The energy used by human cells requires the hydrolysis of 100 to 150 moles of ATP daily, which is around 50 to 75 kg. In typical situations, humans use up their body weight of ATP over the course of the day.[62] This means that each ATP molecule is recycled 1000 to 1500 times daily.

Evolution

Organic cofactors, such as

last universal ancestor, which lived about 4 billion years ago.[64][65]

Organic cofactors may have been present even earlier in the

RNA world, with early ribozymes evolving to bind a restricted set of nucleotides and related compounds.[67][68] Adenosine-based cofactors are thought to have acted as interchangeable adaptors that allowed enzymes and ribozymes to bind new cofactors through small modifications in existing adenosine-binding domains, which had originally evolved to bind a different cofactor.[9] This process of adapting a pre-evolved structure for a novel use is known as exaptation
.

A computational method, IPRO, recently predicted mutations that experimentally switched the cofactor specificity of Candida boidinii xylose reductase from NADPH to NADH.[69]

History

The first organic cofactor to be discovered was NAD+, which was identified by

alcoholic fermentation in unboiled yeast extracts. They called the unidentified factor responsible for this effect a coferment. Through a long and difficult purification from yeast extracts, this heat-stable factor was identified as a nucleotide sugar phosphate by Hans von Euler-Chelpin.[71] Other cofactors were identified throughout the early 20th century, with ATP being isolated in 1929 by Karl Lohmann,[72] and coenzyme A being discovered in 1945 by Fritz Albert Lipmann.[73]

The functions of these molecules were at first mysterious, but, in 1936, Otto Heinrich Warburg identified the function of NAD+ in hydride transfer.[74] This discovery was followed in the early 1940s by the work of Herman Kalckar, who established the link between the oxidation of sugars and the generation of ATP.[75] This confirmed the central role of ATP in energy transfer that had been proposed by Fritz Albert Lipmann in 1941.[76] Later, in 1949, Morris Friedkin and Albert L. Lehninger proved that NAD+ linked metabolic pathways such as the citric acid cycle and the synthesis of ATP.[77]

Protein-derived cofactors

In a number of enzymes, the moiety that acts as a cofactor is formed by post-translational modification of a part of the protein sequence. This often replaces the need for an external binding factor, such as a metal ion, for protein function. Potential modifications could be oxidation of aromatic residues, binding between residues, cleavage or ring-forming.[78] These alterations are distinct from other post-translation protein modifications, such as phosphorylation, methylation, or glycosylation in that the amino acids typically acquire new functions. This increases the functionality of the protein; unmodified amino acids are typically limited to acid-base reactions, and the alteration of resides can give the protein electrophilic sites or the ability to stabilize free radicals.[78] Examples of cofactor production include tryptophan tryptophylquinone (TTQ), derived from two tryptophan side chains,[79] and 4-methylidene-imidazole-5-one (MIO), derived from an Ala-Ser-Gly motif.[80] Characterization of protein-derived cofactors is conducted using X-ray crystallography and mass spectroscopy; structural data is necessary because sequencing does not readily identify the altered sites.

Non-enzymatic cofactors

The term is used in other areas of biology to refer more broadly to non-protein (or even protein) molecules that either activate, inhibit, or are required for the protein to function. For example,

hormones that bind to and activate receptor proteins are termed cofactors or coactivators, whereas molecules that inhibit receptor proteins are termed corepressors. One such example is the G protein-coupled receptor family of receptors, which are frequently found in sensory neurons. Ligand binding to the receptors activates the G protein, which then activates an enzyme to activate the effector.[81] In order to avoid confusion, it has been suggested that such proteins that have ligand-binding mediated activation or repression be referred to as coregulators.[82]

See also

References

  1. .
  2. ^ "coenzymes and cofactors". Archived from the original on 1999-08-26. Retrieved 2007-11-17.
  3. ^ "Enzyme Cofactors". Archived from the original on 2003-05-05. Retrieved 2007-11-17.
  4. .
  5. ^ .
  6. ^ de Bolster, M. W. G. (1997). GLOSSARY OF TERMS USED IN BIOINORGANIC CHEMISTRY (PDF). Pure & Appl. Chem.
  7. .
  8. ^ "Pyruvate Dehydrogenase Complex". Chemistry LibreTexts. 2013-10-02. Retrieved 2017-05-10.
  9. ^
    S2CID 10848692
    .
  10. .
  11. ^ "Biochemistry: Enzymes: Classification and catalysis (Cofactors)". vle.du.ac.in. Retrieved 2018-02-07.[permanent dead link]
  12. PMID 3905079
    .
  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. S2CID 28013583. Archived from the original
    on 16 December 2008.
  48. .
  49. .
  50. .
  51. .
  52. .
  53. .
  54. .
  55. .
  56. .
  57. .
  58. .
  59. .
  60. .
  61. .
  62. ^ Di Carlo SE, Collins HL (2001). "Estimating ATP resynthesis during a marathon run: a method to introduce metabolism". Advan. Physiol. Edu. 25 (2): 70–1.
  63. S2CID 44873410
    .
  64. .
  65. .
  66. .
  67. .
  68. .
  69. .
  70. .
  71. ^ "Fermentation of sugars and fermentative enzymes: Nobel Lecture, May 23, 1930" (PDF). Nobel Foundation. Retrieved 2007-09-30.
  72. S2CID 20328411
    .
  73. .
  74. .
  75. .
  76. .
  77. .
  78. ^ .
  79. .
  80. .
  81. ^ Lodish, Harvey; Berk, Arnold; Zipursky, S. Lawrence; Matsudaira, Paul; Baltimore, David; Darnell, James (2000-01-01). "G Protein–Coupled Receptors and Their Effectors". Molecular Cell Biology (4th ed.).
  82. PMID 18701638
    .

Further reading

External links