Catalase

Last updated
Catalase
PDB 7cat EBI.jpg
Identifiers
SymbolCatalase
Pfam PF00199
InterPro IPR011614
PROSITE PDOC00395
SCOP2 7cat / SCOPe / SUPFAM
OPM superfamily 370
OPM protein 3e4w
CDD cd00328
Available protein structures:
Pfam   structures / ECOD  
PDB RCSB PDB; PDBe; PDBj
PDBsum structure summary
Catalase
Identifiers
EC no. 1.11.1.6
CAS no. 9001-05-2
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
CAT
Catalase Structure.png
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases CAT , catalase
External IDs OMIM: 115500 MGI: 88271 HomoloGene: 55514 GeneCards: CAT
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_001752

NM_009804

RefSeq (protein)

NP_001743

NP_033934

Location (UCSC) Chr 11: 34.44 – 34.47 Mb Chr 2: 103.28 – 103.32 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

Catalase is a common enzyme found in nearly all living organisms exposed to oxygen (such as bacteria, plants, and animals) which catalyzes the decomposition of hydrogen peroxide to water and oxygen. [5] It is a very important enzyme in protecting the cell from oxidative damage by reactive oxygen species (ROS). Catalase has one of the highest turnover numbers of all enzymes; one catalase molecule can convert millions of hydrogen peroxide molecules to water and oxygen each second. [6]

Contents

Catalase is a tetramer of four polypeptide chains, each over 500 amino acids long. [7] It contains four iron-containing heme groups that allow the enzyme to react with hydrogen peroxide. The optimum pH for human catalase is approximately 7, [8] and has a fairly broad maximum: the rate of reaction does not change appreciably between pH 6.8 and 7.5. [9] The pH optimum for other catalases varies between 4 and 11 depending on the species. [10] The optimum temperature also varies by species. [11]

Structure

Human catalase forms a tetramer composed of four subunits, each of which can be conceptually divided into four domains. [12] The extensive core of each subunit is generated by an eight-stranded antiparallel β-barrel (β1-8), with nearest neighbor connectivity capped by β-barrel loops on one side and α9 loops on the other. [12] A helical domain at one face of the β-barrel is composed of four C-terminal helices (α16, α17, α18, and α19) and four helices derived from residues between β4 and β5 (α4, α5, α6, and α7). [12] Alternative splicing may result in different protein variants.

History

Catalase was first noticed in 1818 by Louis Jacques Thénard, who discovered hydrogen peroxide (H2O2). Thénard suggested its breakdown was caused by an unknown substance. In 1900, Oscar Loew was the first to give it the name catalase, and found it in many plants and animals. [13] In 1937 catalase from beef liver was crystallized by James B. Sumner and Alexander Dounce [14] and the molecular weight was measured in 1938. [15]

The amino acid sequence of bovine catalase was determined in 1969, [16] and the three-dimensional structure in 1981. [17]

Function

Molecular mechanism

While the complete mechanism of catalase is not currently known, [18] the reaction is believed to occur in two stages:

H2O2 + Fe(III)-E → H2O + O=Fe(IV)-E(.+)
H2O2 + O=Fe(IV)-E(.+) → H2O + Fe(III)-E + O2 [18]

Here Fe()-E represents the iron center of the heme group attached to the enzyme. Fe(IV)-E(.+) is a mesomeric form of Fe(V)-E, meaning the iron is not completely oxidized to +V, but receives some stabilising electron density from the heme ligand, which is then shown as a radical cation (.+).

As hydrogen peroxide enters the active site, it does not interact with the amino acids Asn148 (asparagine at position 148) and His75, causing a proton (hydrogen ion) to transfer between the oxygen atoms. The free oxygen atom coordinates, freeing the newly formed water molecule and Fe(IV)=O. Fe(IV)=O reacts with a second hydrogen peroxide molecule to reform Fe(III)-E and produce water and oxygen. [18] The reactivity of the iron center may be improved by the presence of the phenolate ligand of Tyr358 in the fifth coordination position, which can assist in the oxidation of the Fe(III) to Fe(IV). The efficiency of the reaction may also be improved by the interactions of His75 and Asn148 with reaction intermediates. [18] The decomposition of hydrogen peroxide by catalase proceeds according to first-order kinetics, the rate being proportional to the hydrogen peroxide concentration. [19]

Catalase can also catalyze the oxidation, by hydrogen peroxide, of various metabolites and toxins, including formaldehyde, formic acid, phenols, acetaldehyde and alcohols. It does so according to the following reaction:

H2O2 + H2R → 2H2O + R

The exact mechanism of this reaction is not known.

Any heavy metal ion (such as copper cations in copper(II) sulfate) can act as a noncompetitive inhibitor of catalase. However, "Copper deficiency can lead to a reduction in catalase activity in tissues, such as heart and liver." [20] Furthermore, the poison cyanide is a noncompetitive inhibitor [21] of catalase at high concentrations of hydrogen peroxide. [22] Arsenate acts as an activator. [23] Three-dimensional protein structures of the peroxidated catalase intermediates are available at the Protein Data Bank.

Cellular role

Hydrogen peroxide is a harmful byproduct of many normal metabolic processes; to prevent damage to cells and tissues, it must be quickly converted into other, less dangerous substances. To this end, catalase is frequently used by cells to rapidly catalyze the decomposition of hydrogen peroxide into less-reactive gaseous oxygen and water molecules. [24]

Mice genetically engineered to lack catalase are initially phenotypically normal. [25] However, catalase deficiency in mice may increase the likelihood of developing obesity, fatty liver, [26] and type 2 diabetes. [27] Some humans have very low levels of catalase (acatalasia), yet show few ill effects.

The increased oxidative stress that occurs with aging in mice is alleviated by over-expression of catalase. [28] Over-expressing mice do not exhibit the age-associated loss of spermatozoa, testicular germ and Sertoli cells seen in wild-type mice. Oxidative stress in wild-type mice ordinarily induces oxidative DNA damage (measured as 8-oxodG) in sperm with aging, but these damages are significantly reduced in aged catalase over-expressing mice. [28] Furthermore, these over-expressing mice show no decrease in age-dependent number of pups per litter. Overexpression of catalase targeted to mitochondria extends the lifespan of mice. [29]

In eukaryotes, catalase is usually located in a cellular organelle called the peroxisome. [30] Peroxisomes in plant cells are involved in photorespiration (the use of oxygen and production of carbon dioxide) and symbiotic nitrogen fixation (the breaking apart of diatomic nitrogen (N2) to reactive nitrogen atoms). Hydrogen peroxide is used as a potent antimicrobial agent when cells are infected with a pathogen. Catalase-positive pathogens, such as Mycobacterium tuberculosis , Legionella pneumophila , and Campylobacter jejuni , make catalase to deactivate the peroxide radicals, thus allowing them to survive unharmed within the host. [31]

Like alcohol dehydrogenase, catalase converts ethanol to acetaldehyde, but it is unlikely that this reaction is physiologically significant. [32]

Distribution among organisms

The large majority of known organisms use catalase in every organ, with particularly high concentrations occurring in the liver in mammals. [33] Catalase is found primarily in peroxisomes and the cytosol of erythrocytes (and sometimes in mitochondria [34] )

Almost all aerobic microorganisms use catalase. It is also present in some anaerobic microorganisms, such as Methanosarcina barkeri . [35] Catalase is also universal among plants and occurs in most fungi. [36]

One unique use of catalase occurs in the bombardier beetle. This beetle has two sets of liquids that are stored separately in two paired glands. The larger of the pair, the storage chamber or reservoir, contains hydroquinones and hydrogen peroxide, while the smaller, the reaction chamber, contains catalases and peroxidases. To activate the noxious spray, the beetle mixes the contents of the two compartments, causing oxygen to be liberated from hydrogen peroxide. The oxygen oxidizes the hydroquinones and also acts as the propellant. [37] The oxidation reaction is very exothermic (ΔH = −202.8 kJ/mol) and rapidly heats the mixture to the boiling point. [38]

Long-lived queens of the termite Reticulitermes speratus have significantly lower oxidative damage to their DNA than non-reproductive individuals (workers and soldiers). [39] Queens have more than two times higher catalase activity and seven times higher expression levels of the catalase gene RsCAT1 than workers. [39] It appears that the efficient antioxidant capability of termite queens can partly explain how they attain longer life.

Catalase enzymes from various species have vastly differing optimum temperatures. Poikilothermic animals typically have catalases with optimum temperatures in the range of 15-25 °C, while mammalian or avian catalases might have optimum temperatures above 35 °C, [40] [41] and catalases from plants vary depending on their growth habit. [40] In contrast, catalase isolated from the hyperthermophile archaeon Pyrobaculum calidifontis has a temperature optimum of 90 °C. [42]

Clinical significance and application

Hydrogen peroxide Wasserstoffperoxid.svg
Hydrogen peroxide

Catalase is used in the food industry for removing hydrogen peroxide from milk prior to cheese production. [43] Another use is in food wrappers, where it prevents food from oxidizing. [44] Catalase is also used in the textile industry, removing hydrogen peroxide from fabrics to make sure the material is peroxide-free. [45]

A minor use is in contact lens hygiene – a few lens-cleaning products disinfect the lens using a hydrogen peroxide solution; a solution containing catalase is then used to decompose the hydrogen peroxide before the lens is used again. [46]

Bacterial identification (catalase test)

Positive catalase reaction Catalase reaction.jpg
Positive catalase reaction

The catalase test is one of the three main tests used by microbiologists to identify species of bacteria. If the bacteria possess catalase (i.e., are catalase-positive), bubbles of oxygen are observed when a small amount of bacterial isolate is added to hydrogen peroxide. The catalase test is done by placing a drop of hydrogen peroxide on a microscope slide. An applicator stick is touched to the colony, and the tip is then smeared onto the hydrogen peroxide drop.

While the catalase test alone cannot identify a particular organism, it can aid identification when combined with other tests such as antibiotic resistance. The presence of catalase in bacterial cells depends on both the growth condition and the medium used to grow the cells.

Capillary tubes may also be used. A small sample of bacteria is collected on the end of the capillary tube, without blocking the tube, to avoid false negative results. The opposite end is then dipped into hydrogen peroxide, which is drawn into the tube through capillary action, and turned upside down, so that the bacterial sample points downwards. The hand holding the tube is then tapped on the bench, moving the hydrogen peroxide down until it touches the bacteria. If bubbles form on contact, this indicates a positive catalase result. This test can detect catalase-positive bacteria at concentrations above about 105 cells/mL, [50] and is simple to use.

Bacterial virulence

Neutrophils and other phagocytes use peroxide to kill bacteria. The enzyme NADPH oxidase generates superoxide within the phagosome, which is converted via hydrogen peroxide to other oxidising substances like hypochlorous acid which kill phagocytosed pathogens. [51] In individuals with chronic granulomatous disease (CGD), phagocytic peroxide production is impaired due to a defective NADPH oxidase system. Normal cellular metabolism will still produce a small amount of peroxide and this peroxide can be used to produce hypochlorous acid to eradicate the bacterial infection. However, if individuals with CGD are infected with catalase-positive bacteria, the bacterial catalase can destroy the excess peroxide before it can be used to produce other oxidising substances. In these individuals the pathogen survives and becomes a chronic infection. This chronic infection is typically surrounded by macrophages in an attempt to isolate the infection. This wall of macrophages surrounding a pathogen is called a granuloma. Many bacteria are catalase positive, but some are better catalase-producers than others. Some catalase-positive bacteria and fungi include: Nocardia, Pseudomonas, Listeria, Aspergillus, Candida, E. coli, Staphylococcus, Serratia, B. cepacia and H. pylori . [52]

Acatalasia

Acatalasia is a condition caused by homozygous mutations in CAT, resulting in a lack of catalase. Symptoms are mild and include oral ulcers. A heterozygous CAT mutation results in lower, but still present catalase. [53]

Gray hair

Low levels of catalase may play a role in the graying process of human hair. Hydrogen peroxide is naturally produced by the body and broken down by catalase. Hydrogen peroxide can accumulate in hair follicles and if catalase levels decline, this buildup can cause oxidative stress and graying. [54] These low levels of catalase are associated with old age. Hydrogen peroxide interferes with the production of melanin, the pigment that gives hair its color. [55] [56]

Interactions

Catalase has been shown to interact with the ABL2 [57] and Abl genes. [57] Infection with the murine leukemia virus causes catalase activity to decline in the lungs, heart and kidneys of mice. Conversely, dietary fish oil increased catalase activity in the heart, and kidneys of mice. [58]

Methods for determining catalase activity

In 1870, Schoenn discovered a formation of yellow color from the interaction of hydrogen peroxide with molybdate; [59] then, from the middle of the 20th century, this reaction began to be used for colorimetric determination of unreacted hydrogen peroxide in the catalase activity assay. [60] The reaction became widely used after publications by Korolyuk et al. (1988) [61] and Goth (1991). [62] The first paper describes serum catalase assay with no buffer in the reaction medium; the latter describes the procedure based on phosphate buffer as a reaction medium. Since phosphate ion reacts with ammonium molybdate, [62] the use of MOPS buffer as a reaction medium is more appropriate. [63]

Direct UV measurement of the decrease in the concentration of hydrogen peroxide is also widely used after the publications by Beers & Sizer [64] and Aebi. [65]

See also

Related Research Articles

Antioxidants are compounds that inhibit oxidation, a chemical reaction that can produce free radicals. Autoxidation leads to degradation of organic compounds, including living matter. Antioxidants are frequently added to industrial products, such as polymers, fuels, and lubricants, to extend their usable lifetimes. Foods are also treated with antioxidants to forestall spoilage, in particular the rancidification of oils and fats. In cells, antioxidants such as glutathione, mycothiol or bacillithiol, and enzyme systems like superoxide dismutase, can prevent damage from oxidative stress.

<span class="mw-page-title-main">Hydrogen peroxide</span> Chemical compound

Hydrogen peroxide is a chemical compound with the formula H2O2. In its pure form, it is a very pale blue liquid that is slightly more viscous than water. It is used as an oxidizer, bleaching agent, and antiseptic, usually as a dilute solution in water for consumer use, and in higher concentrations for industrial use. Concentrated hydrogen peroxide, or "high-test peroxide", decomposes explosively when heated and has been used both as a monopropellant and an oxidizer in rocketry.

<span class="mw-page-title-main">Peroxisome</span> Type of organelle

A peroxisome (IPA:[pɛɜˈɹɒksɪˌsoʊm]) is a membrane-bound organelle, a type of microbody, found in the cytoplasm of virtually all eukaryotic cells. Peroxisomes are oxidative organelles. Frequently, molecular oxygen serves as a co-substrate, from which hydrogen peroxide (H2O2) is then formed. Peroxisomes owe their name to hydrogen peroxide generating and scavenging activities. They perform key roles in lipid metabolism and the reduction of reactive oxygen species.

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

Superoxide dismutase (SOD, EC 1.15.1.1) is an enzyme that alternately catalyzes the dismutation (or partitioning) of the superoxide (O
2
) radical into ordinary molecular oxygen (O2) and hydrogen peroxide (H
2
O
2
). Superoxide is produced as a by-product of oxygen metabolism and, if not regulated, causes many types of cell damage. Hydrogen peroxide is also damaging and is degraded by other enzymes such as catalase. Thus, SOD is an important antioxidant defense in nearly all living cells exposed to oxygen. One exception is Lactobacillus plantarum and related lactobacilli, which use a different mechanism to prevent damage from reactive O
2
.

In chemistry, a superoxide is a compound that contains the superoxide ion, which has the chemical formula O−2. The systematic name of the anion is dioxide(1−). The reactive oxygen ion superoxide is particularly important as the product of the one-electron reduction of dioxygen O2, which occurs widely in nature. Molecular oxygen (dioxygen) is a diradical containing two unpaired electrons, and superoxide results from the addition of an electron which fills one of the two degenerate molecular orbitals, leaving a charged ionic species with a single unpaired electron and a net negative charge of −1. Both dioxygen and the superoxide anion are free radicals that exhibit paramagnetism. Superoxide was historically also known as "hyperoxide".

<span class="mw-page-title-main">Glutathione peroxidase</span> Enzyme family protecting the organism from oxidative damages

Glutathione peroxidase (GPx) is the general name of an enzyme family with peroxidase activity whose main biological role is to protect the organism from oxidative damage. The biochemical function of glutathione peroxidase is to reduce lipid hydroperoxides to their corresponding alcohols and to reduce free hydrogen peroxide to water.

<span class="mw-page-title-main">Cytochrome c peroxidase</span>

Cytochrome c peroxidase, or CCP, is a water-soluble heme-containing enzyme of the peroxidase family that takes reducing equivalents from cytochrome c and reduces hydrogen peroxide to water:

<span class="mw-page-title-main">Reactive oxygen species</span> Highly reactive molecules formed from diatomic oxygen (O₂)

In chemistry and biology, reactive oxygen species (ROS) are highly reactive chemicals formed from diatomic oxygen (O2), water, and hydrogen peroxide. Some prominent ROS are hydroperoxide (O2H), superoxide (O2-), hydroxyl radical (OH.), and singlet oxygen. ROS are pervasive because they are readily produced from O2, which is abundant. ROS are important in many ways, both beneficial and otherwise. ROS function as signals, that turn on and off biological functions. They are intermediates in the redox behavior of O2, which is central to fuel cells. ROS are central to the photodegradation of organic pollutants in the atmosphere. Most often however, ROS are discussed in a biological context, ranging from their effects on aging and their role in causing dangerous genetic mutations.

Lipid peroxidation is the conversion of lipids to peroxide and hydroperoxide derivatives. These derivatives, known as lipid peroxides or lipid oxidation products (LOPs), are susceptible to further reactions that are relevant to "DNA and protein modification, radiation damage, aging..." Lipid peroxidation mainly applies to unsaturated fats, especially polyunsaturated fats such as those derived from linoleic acid.

Respiratory burst is the rapid release of the reactive oxygen species (ROS), superoxide anion and hydrogen peroxide, from different cell types.

<span class="mw-page-title-main">Peroxiredoxin</span> Family of antioxidant enzymes

Peroxiredoxins are a ubiquitous family of antioxidant enzymes that also control cytokine-induced peroxide levels and thereby mediate signal transduction in mammalian cells. The family members in humans are PRDX1, PRDX2, PRDX3, PRDX4, PRDX5, and PRDX6. The physiological importance of peroxiredoxins is indicated by their relative abundance. Their function is the reduction of peroxides, specifically hydrogen peroxide, alkyl hydroperoxides, and peroxynitrite.

In enzymology, a lignin peroxidase (EC 1.11.1.14) is an enzyme that catalyzes the chemical reaction

In enzymology, a manganese peroxidase (EC 1.11.1.13) is an enzyme that catalyzes the chemical reaction

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

In enzymology, a NADH peroxidase (EC 1.11.1.1) is an enzyme that catalyzes the chemical reaction

<span class="mw-page-title-main">Lactoperoxidase</span> Mammalian protein found in Homo sapiens

Lactoperoxidase is a peroxidase enzyme secreted from mammary, salivary and other mucosal glands including the lungs, bronchii and nose that functions as a natural and the first line of defense against bacteria and viruses. Lactoperoxidase is a member of the heme peroxidase family of enzymes. In humans, lactoperoxidase is encoded by the LPO gene.

Haem peroxidases (or heme peroxidases) are haem-containing enzymes that use hydrogen peroxide as the electron acceptor to catalyse a number of oxidative reactions. Most haem peroxidases follow the reaction scheme:

All living cells produce reactive oxygen species (ROS) as a byproduct of metabolism. ROS are reduced oxygen intermediates that include the superoxide radical (O2) and the hydroxyl radical (OH•), as well as the non-radical species hydrogen peroxide (H2O2). These ROS are important in the normal functioning of cells, playing a role in signal transduction and the expression of transcription factors. However, when present in excess, ROS can cause damage to proteins, lipids and DNA by reacting with these biomolecules to modify or destroy their intended function. As an example, the occurrence of ROS have been linked to the aging process in humans, as well as several other diseases including Alzheimer's, rheumatoid arthritis, Parkinson's, and some cancers. Their potential for damage also makes reactive oxygen species useful in direct protection from invading pathogens, as a defense response to physical injury, and as a mechanism for stopping the spread of bacteria and viruses by inducing programmed cell death.

Catalase-peroxidase (EC 1.11.1.21, katG (gene)) is an enzyme with systematic name donor:hydrogen-peroxide oxidoreductase. This enzyme catalyses the following chemical reaction

  1. donor + H2O2 ⇌ oxidized donor + 2 H2O
  2. 2 H2O2 ⇌ O2 + 2 H2O

Oxidation response is stimulated by a disturbance in the balance between the production of reactive oxygen species and antioxidant responses, known as oxidative stress. Active species of oxygen naturally occur in aerobic cells and have both intracellular and extracellular sources. These species, if not controlled, damage all components of the cell, including proteins, lipids and DNA. Hence cells need to maintain a strong defense against the damage. The following table gives an idea of the antioxidant defense system in bacterial system.

<span class="mw-page-title-main">Eosinophil peroxidase</span> Protein-coding gene in the species Homo sapiens

Eosinophil peroxidase is an enzyme found within the eosinophil granulocytes, innate immune cells of humans and mammals. This oxidoreductase protein is encoded by the gene EPX, expressed within these myeloid cells. EPO shares many similarities with its orthologous peroxidases, myeloperoxidase (MPO), lactoperoxidase (LPO), and thyroid peroxidase (TPO). The protein is concentrated in secretory granules within eosinophils. Eosinophil peroxidase is a heme peroxidase, its activities including the oxidation of halide ions to bacteriocidal reactive oxygen species, the cationic disruption of bacterial cell walls, and the post-translational modification of protein amino acid residues.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000121691 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000027187 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. Chelikani P, Fita I, Loewen PC (January 2004). "Diversity of structures and properties among catalases". Cellular and Molecular Life Sciences. 61 (2): 192–208. doi:10.1007/s00018-003-3206-5. hdl: 10261/111097 . PMID   14745498. S2CID   4411482.
  6. Goodsell DS (2004-09-01). "Catalase". Molecule of the Month. RCSB Protein Data Bank. Retrieved 2016-08-23.
  7. Boon EM, Downs A, Marcey D. "Catalase: H2O2: H2O2 Oxidoreductase". Catalase Structural Tutorial Text. Retrieved 2007-02-11.
  8. Maehly AC, Chance B (1954). "The assay of catalases and peroxidases". Methods of Biochemical Analysis. Vol. 1. pp. 357–424. doi:10.1002/9780470110171.ch14. ISBN   978-0-470-11017-1. PMID   13193536.
  9. Aebi H (1984). "Catalase in vitro". Oxygen Radicals in Biological Systems. Methods in Enzymology. Vol. 105. pp.  121–6. doi:10.1016/S0076-6879(84)05016-3. ISBN   978-0-12-182005-3. PMID   6727660.
  10. "EC 1.11.1.6 - catalase". BRENDA: The Comprehensive Enzyme Information System. Department of Bioinformatics and Biochemistry, Technische Universität Braunschweig. Retrieved 2009-05-26.
  11. Toner K, Sojka G, Ellis R. "A Quantitative Enzyme Study; CATALASE". bucknell.edu. Archived from the original on 2000-06-12. Retrieved 2007-02-11.
  12. 1 2 3 Putnam CD, Arvai AS, Bourne Y, Tainer JA (February 2000). "Active and inhibited human catalase structures: ligand and NADPH binding and catalytic mechanism". Journal of Molecular Biology. 296 (1): 295–309. doi:10.1006/jmbi.1999.3458. PMID   10656833.
  13. Loew O (May 1900). "A New Enzyme of General Occurrence in Organisms". Science. 11 (279): 701–702. Bibcode:1900Sci....11..701L. doi:10.1126/science.11.279.701. JSTOR   1625707. PMID   17751716.
  14. Sumner JB, Dounce AL (April 1937). "Crystalline Catalase". Science. 85 (2206): 366–367. Bibcode:1937Sci....85..366S. doi:10.1126/science.85.2206.366. PMID   17776781.
  15. Sumner JB, Gralén N (March 1938). "The Molecular Weight of Crystalline Catalase". Science. 87 (2256): 284. Bibcode:1938Sci....87..284S. doi:10.1126/science.87.2256.284. PMID   17831682. S2CID   36931581.
  16. Schroeder WA, Shelton JR, Shelton JB, Robberson B, Apell G (May 1969). "The amino acid sequence of bovine liver catalase: a preliminary report". Archives of Biochemistry and Biophysics. 131 (2): 653–655. doi:10.1016/0003-9861(69)90441-X. PMID   4892021.
  17. Murthy MR, Reid TJ, Sicignano A, Tanaka N, Rossmann MG (October 1981). "Structure of beef liver catalase". Journal of Molecular Biology. 152 (2): 465–499. doi:10.1016/0022-2836(81)90254-0. PMID   7328661.
  18. 1 2 3 4 Boon EM, Downs A, Marcey D. "Proposed Mechanism of Catalase". Catalase: H2O2: H2O2 Oxidoreductase: Catalase Structural Tutorial. Retrieved 2007-02-11.
  19. Aebi H (1984). "Catalase in vitro". Oxygen Radicals in Biological Systems. Methods in Enzymology. Vol. 105. pp. 121–126. doi:10.1016/S0076-6879(84)05016-3. ISBN   9780121820053. PMID   6727660.
  20. Hordyjewska A, Popiołek Ł, Kocot J (August 2014). "The many "faces" of copper in medicine and treatment". Biometals. 27 (4): 611–621. doi:10.1007/s10534-014-9736-5. PMC   4113679 . PMID   24748564.
  21. Nonstationary Inhibition of Enzyme Action. The Cyanide Inhibition of Catalase
  22. Ogura Y, Yamazaki I (August 1983). "Steady-state kinetics of the catalase reaction in the presence of cyanide". Journal of Biochemistry. 94 (2): 403–408. doi:10.1093/oxfordjournals.jbchem.a134369. PMID   6630165.
  23. Kertulis-Tartar GM, Rathinasabapathi B, Ma LQ (October 2009). "Characterization of glutathione reductase and catalase in the fronds of two Pteris ferns upon arsenic exposure". Plant Physiology and Biochemistry. 47 (10): 960–965. doi:10.1016/j.plaphy.2009.05.009. PMID   19574057.
  24. Gaetani GF, Ferraris AM, Rolfo M, Mangerini R, Arena S, Kirkman HN (February 1996). "Predominant role of catalase in the disposal of hydrogen peroxide within human erythrocytes". Blood. 87 (4): 1595–1599. doi: 10.1182/blood.V87.4.1595.bloodjournal8741595 . PMID   8608252.
  25. Ho YS, Xiong Y, Ma W, Spector A, Ho DS (July 2004). "Mice lacking catalase develop normally but show differential sensitivity to oxidant tissue injury". The Journal of Biological Chemistry. 279 (31): 32804–32812. doi: 10.1074/jbc.M404800200 . PMID   15178682.
  26. Heit C, Marshall S, Singh S, Yu X, Charkoftaki G, Zhao H, et al. (February 2017). "Catalase deletion promotes prediabetic phenotype in mice". Free Radical Biology & Medicine. 103: 48–56. doi:10.1016/j.freeradbiomed.2016.12.011. PMC   5513671 . PMID   27939935.
  27. Góth L, Nagy T (September 2012). "Acatalasemia and diabetes mellitus". Archives of Biochemistry and Biophysics. 525 (2): 195–200. doi:10.1016/j.abb.2012.02.005. PMID   22365890.
  28. 1 2 Selvaratnam J, Robaire B (November 2016). "Overexpression of catalase in mice reduces age-related oxidative stress and maintains sperm production". Experimental Gerontology. 84: 12–20. doi:10.1016/j.exger.2016.08.012. PMID   27575890. S2CID   2416413.
  29. Schriner SE, Linford NJ, Martin GM, Treuting P, Ogburn CE, Emond M, et al. (June 2005). "Extension of murine life span by overexpression of catalase targeted to mitochondria". Science. 308 (5730): 1909–1911. Bibcode:2005Sci...308.1909S. doi:10.1126/science.1106653. PMID   15879174. S2CID   38568666.
  30. Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2002). "Peroxisomes". Molecular Biology of the Cell (4th ed.). New York: Garland Science. ISBN   978-0-8153-3218-3.
  31. Srinivasa Rao PS, Yamada Y, Leung KY (September 2003). "A major catalase (KatB) that is required for resistance to H2O2 and phagocyte-mediated killing in Edwardsiella tarda". Microbiology. 149 (Pt 9): 2635–2644. doi: 10.1099/mic.0.26478-0 . PMID   12949187.
  32. Lieber CS (January 1997). "Ethanol metabolism, cirrhosis and alcoholism". Clinica Chimica Acta; International Journal of Clinical Chemistry. 257 (1): 59–84. doi:10.1016/S0009-8981(96)06434-0. PMID   9028626.
  33. Ilyukha VA (2001). "Superoxide Dismutase and Catalase in the Organs of Mammals of Different Ecogenesis". Journal of Evolutionary Biochemistry and Physiology. 37 (3): 241–245. doi:10.1023/A:1012663105999. S2CID   38916410.
  34. Bai J, Cederbaum AI (2001). "Mitochondrial catalase and oxidative injury". Biological Signals and Receptors. 10 (3–4): 189–199. doi:10.1159/000046887. PMID   11351128. S2CID   33795198.
  35. Brioukhanov AL, Netrusov AI, Eggen RI (June 2006). "The catalase and superoxide dismutase genes are transcriptionally up-regulated upon oxidative stress in the strictly anaerobic archaeon Methanosarcina barkeri". Microbiology. 152 (Pt 6): 1671–1677. doi: 10.1099/mic.0.28542-0 . PMID   16735730.
  36. Hansberg W, Salas-Lizana R, Domínguez L (September 2012). "Fungal catalases: function, phylogenetic origin and structure". Archives of Biochemistry and Biophysics. 525 (2): 170–180. doi:10.1016/j.abb.2012.05.014. PMID   22698962.
  37. Eisner T, Aneshansley DJ (August 1999). "Spray aiming in the bombardier beetle: photographic evidence". Proceedings of the National Academy of Sciences of the United States of America. 96 (17): 9705–9709. Bibcode:1999PNAS...96.9705E. doi: 10.1073/pnas.96.17.9705 . PMC   22274 . PMID   10449758.
  38. Beheshti N, McIntosh AC (2006). "A biomimetic study of the explosive discharge of the bombardier beetle" (PDF). International Journal of Design & Nature. 1 (1): 1–9. Archived from the original (PDF) on 2011-07-26.
  39. 1 2 Tasaki E, Kobayashi K, Matsuura K, Iuchi Y (2017). "An Efficient Antioxidant System in a Long-Lived Termite Queen". PLOS ONE. 12 (1): e0167412. Bibcode:2017PLoSO..1267412T. doi: 10.1371/journal.pone.0167412 . PMC   5226355 . PMID   28076409.
  40. 1 2 Mitsuda H (1956-07-31). "Studies on Catalase" (PDF). Bulletin of the Institute for Chemical Research, Kyoto University. 34 (4): 165–192. Archived (PDF) from the original on 2022-10-09. Retrieved 27 September 2017.
  41. Çetinus ŞA, Öztop HN (June 2003). "Immobilization of catalase into chemically crosslinked chitosan beads". Enzyme and Microbial Technology. 32 (7): 889–894. doi:10.1016/S0141-0229(03)00065-6.
  42. Amo T, Atomi H, Imanaka T (June 2002). "Unique presence of a manganese catalase in a hyperthermophilic archaeon, Pyrobaculum calidifontis VA1". Journal of Bacteriology. 184 (12): 3305–3312. doi:10.1128/JB.184.12.3305-3312.2002. PMC   135111 . PMID   12029047.
  43. "Catalase". Worthington Enzyme Manual. Worthington Biochemical Corporation. Retrieved 2009-03-01.
  44. Hengge A (1999-03-16). "Re: how is catalase used in industry?". General Biology. MadSci Network. Retrieved 2009-03-01.
  45. "textile industry". Case study 228. International Cleaner Production Information Clearinghouse. Archived from the original on 2008-11-04. Retrieved 2009-03-01.
  46. USpatent 5521091,Cook JN, Worsley JL,"Compositions and method for destroying hydrogen peroxide on contact lens",issued 1996-05-28
  47. Rollins DM (2000-08-01). "Bacterial Pathogen List". BSCI 424 Pathogenic Microbiology. University of Maryland. Retrieved 2009-03-01.
  48. Johnson M. "Catalase Production". Biochemical Tests. Mesa Community College. Archived from the original on 2008-12-11. Retrieved 2009-03-01.
  49. Fox A. "Streptococcus pneumoniae and Staphylococci". University of South Carolina. Retrieved 2009-03-01.
  50. Martin AM (2012-12-06). Fisheries Processing: Biotechnological applications. Springer Science & Business Media. ISBN   9781461553038.
  51. Winterbourn CC, Kettle AJ, Hampton MB (June 2016). "Reactive Oxygen Species and Neutrophil Function". Annual Review of Biochemistry. 85 (1): 765–792. doi:10.1146/annurev-biochem-060815-014442. PMID   27050287.
  52. Le T, Bhushan V, Sochat M, Kallianos K, Chavda Y, Zureick AH (2017-01-06). First aid for the USMLE step 1 2017: a student-to-student guide (27th ed.). New York: McGraw-Hill Education. ISBN   978-1-259-83762-3. OCLC   986222844.
  53. "OMIM Entry - # 614097 - ACATALASEMIA". www.omim.org.
  54. "Gray hair cure? Scientists find root cause of discoloration". NBC News. 6 May 2013. Retrieved 2022-07-31.
  55. "Why Hair Turns Gray Is No Longer A Gray Area: Our Hair Bleaches Itself As We Grow Older". Science News. ScienceDaily. 2009-02-24. Retrieved 2009-03-01.
  56. Wood JM, Decker H, Hartmann H, Chavan B, Rokos H, Spencer JD, et al. (July 2009). "Senile hair graying: H2O2-mediated oxidative stress affects human hair color by blunting methionine sulfoxide repair". FASEB Journal. 23 (7): 2065–2075. arXiv: 0706.4406 . doi: 10.1096/fj.08-125435 . PMID   19237503. S2CID   16069417.
  57. 1 2 Cao C, Leng Y, Kufe D (August 2003). "Catalase activity is regulated by c-Abl and Arg in the oxidative stress response". The Journal of Biological Chemistry. 278 (32): 29667–29675. doi: 10.1074/jbc.M301292200 . PMID   12777400.
  58. Xi S, Chen LH (2000). "Effects of dietary fish oil on tissue glutathione and antioxidant defense enzymes in mice with murine aids". Nutrition Research. 20 (9): 1287–99. doi:10.1016/S0271-5317(00)00214-1.
  59. Isaacs ML (1922). "A colorimetric determination of hydrogen peroxide". Journal of the American Chemical Society. 44 (8): 1662–1663. doi:10.1021/ja01429a006.
  60. Peizer LR, Widelock D (1955). "A colorimetric test for measuring catalase activity of cultures of M. tuberculosis". American Review of Tuberculosis and Pulmonary Diseases. 71 (2): 305–313. doi:10.1164/artpd.1955.71.2.305 (inactive 31 January 2024). PMID   14350192.{{cite journal}}: CS1 maint: DOI inactive as of January 2024 (link)
  61. Koroliuk MA, Ivanova LI, Maĭorova IG, Tokarev VE (1988). "[A method of determining catalase activity]". Laboratornoe Delo (1): 16–19. PMID   2451064.
  62. 1 2 Góth L (February 1991). "A simple method for determination of serum catalase activity and revision of reference range". Clinica Chimica Acta; International Journal of Clinical Chemistry. 196 (2–3): 143–151. doi:10.1016/0009-8981(91)90067-m. PMID   2029780.
  63. Razygraev AV (2023). "Catalase enzymatic activity in adult mosquitoes (Diptera: Culicidae): taxonomic distribution of the continuous trait suggests its relevance for phylogeny research". Zootaxa. 5339 (2): 159–176. doi:10.11646/zootaxa.5339.2.3. PMID   38221060. S2CID   261383164.
  64. Beers RF, Sizer IW (March 1952). "A spectrophotometric method for measuring the breakdown of hydrogen peroxide by catalase". The Journal of Biological Chemistry. 195 (1): 133–140. doi: 10.1016/S0021-9258(19)50881-X . PMID   14938361.
  65. Aebi H (1984-01-01). "Catalase in vitro". Oxygen Radicals in Biological Systems. Methods in Enzymology. Vol. 105. Academic Press. pp. 121–126. doi:10.1016/s0076-6879(84)05016-3. ISBN   9780121820053. PMID   6727660.