CAZypedia needs your help! We have many unassigned GH, PL, CE, AA, GT, and CBM pages in need of Authors and Responsible Curators.
Scientists at all career stages, including students, are welcome to contribute to CAZypedia. Read more here, and in the 10th anniversary article in Glycobiology.
New to the CAZy classification? Read this first.
*
Consider attending the 15th Carbohydrate Bioengineering Meeting in Ghent, 5-8 May 2024.

Difference between revisions of "Auxiliary Activity Family 3"

From CAZypedia
Jump to navigation Jump to search
m (Add EC link example)
m (Text replacement - "\^\^\^(.*)\^\^\^" to "$1")
 
(28 intermediate revisions by 3 users not shown)
Line 1: Line 1:
 
<!-- RESPONSIBLE CURATORS: Please replace the {{UnderConstruction}} tag below with {{CuratorApproved}} when the page is ready for wider public consumption -->
 
<!-- RESPONSIBLE CURATORS: Please replace the {{UnderConstruction}} tag below with {{CuratorApproved}} when the page is ready for wider public consumption -->
{{UnderConstruction}}
+
{{CuratorApproved}}
* [[Author]]: ^^^Roland Ludwig^^^ and ^^^Daniel Kracher^^^
+
* [[Author]]: [[User:Roland Ludwig|Roland Ludwig]] and [[User:Daniel Kracher|Daniel Kracher]]
* [[Responsible Curator]]:  ^^^Roland Ludwig^^^
+
* [[Responsible Curator]]:  [[User:Roland Ludwig|Roland Ludwig]]
 
----
 
----
  
Line 28: Line 28:
  
  
= Familiy members =
+
== General properties and substrate specificities ==
AA3 members have been divided into three subfamilies based on activity.
+
[[Image:AA3_1.png|thumb|right|400px|'''Figure 1. Reaction cycle of AA3 flavoenzymes'''. In the reductive half-reaction substrate oxidation leads to the formation of a two-electron reduced FADH<sub>2</sub> (flavohydroquinone). During the oxidative half-reaction the flavin cofactor is re-oxidised via electron transfer to a suitable electron acceptor. Exemplary reactions for each AA3 subfamily are shown.]]
 +
Enzymes of the CAZy family AA3 are widespread and catalyse the oxidation of alcohols or carbohydrates with the concomitant formation of hydrogen peroxide or hydroquinones <cite>Levasseur2013</cite>. AA3 enzymes are most abundant in wood-degrading fungi where they typically display a high multigenicity <cite> Suetzl2018 </cite>. The main function of fungal AA3 enzymes is the stimulation of lignocellulose degradation in cooperation with other AA-enzymes such as peroxidases ([[AA2]]) <cite>Ferreira2015 Hernandez2012 Daniel1994</cite> or lytic polysaccharide monooxygenases ([[AA9]]) <cite> Langston2011, Kracher2016, Garajova2016 Bissaro2017</cite>. In yeast, AA3 enzymes are involved in the catabolism of alcohols <cite> Goswami2013</cite>, and some AA3 genes identified in insects are thought to be relevant for immunity and development <cite> Iida2007, Sun2012</cite>. Recently, a bacterial AA3 enzyme with unknown biological function was isolated and characterized <cite> Mendes2017 </cite>. The functionally diverse enzymes of family AA3 all belong to the structurally related glucose-methanol-choline (GMC) family of oxidoreductases and require a flavin-adenine dinucleotide (FAD) cofactor for catalytic activity. Based on their sequences members of the AA3 family were divided into four subfamilies in the CAZy database (Figure 1). Family AA3_1 contains the flavodehydrogenase domain of cellobiose dehydrogenase ([{{EClink}}1.1.99.18 EC 1.1.99.18]), family AA3_2 includes aryl alcohol oxidase ([{{EClink}}1.1.3.7 EC 1.1.3.7]), glucose oxidase ([{{EClink}}1.1.3.4 EC 1.1.3.4]), glucose dehydrogenase ([{{EClink}}1.1.5.9 EC 1.1.5.9]) and pyranose dehydrogenase ([{{EClink}}1.1.99.29 EC 1.1.99.29]), family AA3_3 consists of alcohol (methanol) oxidases ([{{EClink}}1.1.3.13 EC 1.1.3.13]) and family AA3_4 comprises pyranose dehydrogenases([{{EClink}}1.1.3.10 EC 1.1.3.10]).
  
== Subfamily 1 - CDH ==
+
== Subfamily AA3_1: Cellobiose dehydrogenase ==
 +
Cellobiose dehydrogenases (CDHs) are extracellular flavocytochromes that were first described in 1974 <cite>Westermark1974</cite>. CDHs are exclusively found in wood-degrading and phytopathogenic fungi belonging to the phyla Basidiomycota (Class-I CDHs) and Ascomycota (Class-II and -III CDHs) <cite>Kracher2016b</cite>. They oxidise a wide variety of lignocellulose-derived saccharides to their corresponding sugar lactones. CDHs show a high preference for soluble, β-(1,4)-interlinked saccharides and scarcely oxidise monosaccharides.
 +
A common feature of all CDHs is their complex bipartite structure, which comprises an N-terminal cytochrome-binding domain (CYT) and a larger, catalytic flavodehydrogenase (DH) domain encoded within a single polypeptide chain <cite>Hallberg2002</cite> (Figure 2A). Both domains are connected by a flexible linker which typically comprises 15 – 30 amino acids. An important in vivo function of CDH is the reduction and activation of family [[AA9]] lytic polysaccharide monooxygenases via its haem ''b'' domain <cite>Phillips2011, Tan2015 Courtade2016</cite>. Recently, the holoenzyme structures of ''Neurospora crassa'' CDH (pdb: [{{PDBlink}}4qi7 4QI7]) and ''Myriococcum thermophilum'' CDH (pdb: [{{PDBlink}}4qi6 4QI6]) were reported. Two structures of ''N. crassa'' CDH showed an “open-state” conformation in which DH and CYT were spatially separated, whereas a structure of ''M. thermophilum'' CDH showed a “closed-state” conformation in which the propionate arm of the cytochrome domain interacted with the catalytic centre in DH <cite> Tan2015</cite>. Analysis by small angle scattering also suggested a number of possible intermediate conformers that exist in solution <cite>Tan2015, Bodenheimer2018</cite>. While the closed-state allows interdomain electron transfer from DH to CYT, reduction of electron acceptors (e.g. [[AA9]] enzymes) might occur in the open-state, in which the haem cofactor is fully accessible.
  
=== Kinetics and Mechanism ===
+
== Subfamily AA3_2 ==
Cellobiose dehydrogenase (CDH, [{{EClink}}1.1.99.15 EC 1.1.99.15] cellobiose:quinone oxidoreductase) is a flavohaemoprotein (or flavocytochrome) secreted by lignocellulose degrading, phytopathogenic and saprotrophic fungi belonging to the phyla of Basidiomycota and Ascomycota <cite>Zamocky2006</cite>. An important in vivo function of CDH is the reduction of lytic polysaccharide monooxygenases belonging to the auxiliary activity families AA9 and AA10 <cite>Phillips2011 Loose2016</cite>. CDHs oxidize a variety of lignocellulose-derived saccharides to their corresponding 1,5-lactones. All CDHs show a high catalytic efficiency towards soluble, β-(1,4)-linked cellulose derived di- oand oligosaccharides like cellobiose, cellotriose, cellotetraose, cellopentaose, cellohexaose and the structural analogue lactose. Also hemicellulose- and starch- derived oligosaccharides, such as xylo- or manno or maltooligosaccharides are oxidized by a number of CDHs, although with lower catalytic efficiency <cite>Zamocky2006 Zhang2011</cite>. CDHs typically discriminate monosaccharides and show a low catalytic efficiency towards glucose, galactose and mannose <cite>Zamocky2006 Henriksson1998</cite>.
+
==== Aryl alcohol oxidase/dehydrogenase ====
 
+
[[Image:AA3_2.png|thumb|right|400px|'''Figure 2. Structures of AA3 family members.''' A, Cellobiose dehydrogenase from ''Myriococcum thermophilum'' (pdb: [{{PDBlink}}4qi6 4QI6]). The haem ''b''-containing domain is shown in red. B, pyranose oxidase from ''Phanerochaete chrysosporium'' (pdb: [{{PDBlink}}4mif 4MIF]); C, aryl-alcohol oxidase from ''Pleurotus eryngii'' (pdb: [{{PDBlink}}3fim 3FIM]); D, methanol oxidase from ''Pichia pastoris'' (pdb: [{{PDBlink}}5hsa 5HSA]); E, glucose oxidase from ''Aspergillus niger'' (pdb: [{{PDBlink}}4cf4 4CF4]). The GMC-oxidoreductase α/β-fold is colour-coded.]]
CDH has a bipartite structure, which comprises an N-terminal cytochrome domain and a catalytic flavodehydrogenase domain (of the GMC-oxidoreductase family) within a single polypeptide chain. The domains are connected by a papain-sensitive interdomain protein linker which typically comprises 15 - 20 amino acids. The oxidation of carbohydrates occurs at the flavodehydrogenase domain, which binds substrates in a binding (B-) site and thereby orients the anomeric carbon atom of the glucose moiety bound in the cytalytic (C-)site towards the N5 atom of the isoalloxazine ring <cite>Hallberg2003</cite>. This orientation allows oxidative attack of cellobiose by the catalytic histidine by a general hydride transfer mechanism, in which the catalytic His initially abstracts a proton from the C1 hydroxyl group <cite>WongnateandChaiyen2013</cite>. Transfer of the anomeric (C1) hydrogen to the N5 atom of the FAD cofactor then results in a 2-electron reduced FAD (hydroquinone form), while cellobiose is oxidized to cellobionolactone.
+
Aryl alcohol oxidoreductases (AAO) are secreted by basidiomycete fungi and proposed to play a crucial role in lignin degradation. The X-ray crystal structure of the most widely studied AAO from ''Pleurotus eryngii'' (pdb: [{{PDBlink}}3fim 3FIM]) showed a non-covalently bound FAD cofactor in the active-site <cite>Fernandez2009</cite> (Figure 2C). Access to the active site is restricted by three aromatic residues, which interact with both the alcohol substrate and oxygen. The substrate scope of AAO includes secreted benzyl alcohols from the secondary metabolism (e.g. veratryl alcohol), or related alcohols that accumulate during lignin degradation <cite>Hernandez2012</cite>. During the reductive half-reaction, the primary alcohol of the substrate is two-electron oxidised to form the corresponding aldehyde. In addition, aromatic aldehydes can undergo further oxidation yielding the corresponding acids. The concomitantly formed hydrogen peroxide is considered essential for the activity of lignin degrading peroxidases. Recently, aryl-alcohol quinone oxidoreductases (AAQO) with a high sequence identity to ''Pleurotus eryngii'' AAO were identified in the genome of the basidiomycete ''Pycnoporus cinnabarinus'' <cite>Mathieu2016</cite>. While these enzymes showed a substrate spectrum similar to known AAOs, oxygen reduction was insignificant or very low relative to other characterised AAOs. However, reoxidation of FADH<sub>2</sub> was very efficient with benzoquinone or phenoxy-radicals, which are formed by fungal laccases, suggesting a different functional role of AAQOs.
 
 
Stopped-flow spectrophotometry showed that mixing of CDH with cellobiose results in rapid reduction of the FAD cofactor in the flavodehydrogenase domain, followed by a slower interdomain electron transfer to the ''b''-type heme in the cytochrome domain <cite>Igarashi2002</cite>. The electron transfer from FAD to heme ''b'' was further strengthened by the observed midpoint potentials of the cofactors. The heme ''b'' redox potential is always higher than that of FAD, typically between +100 - +150 mV. However, the cytochrome domain is not the only electron acceptor and several substituted quinones, redox dyes and metal complexes have been reported as electron acceptors directly interacting with the FAD <cite>Zamocky2006 Henriksson1998</cite>.
 
 
 
Based on phylogenetic, catalytic and molecular differences CDHs are classified into three classes <cite>Zamocky2004</cite>. Class-I CDHs are of basidiomycetous origin, feature a cellulose binding site in the flavodehydrogenase domain, show a strong discrimination of monosaccharides and have an acidic pH optimum. Class-II CDHs are of ascomycetous origin. Subclass CDH-IIA features a carbohydrate binding module (CBM1), whicle CDH-IIB lacks a CDM or a cellulose binding site. Ascomycetous CDHs show a broader substrate spectrum can have acidic or alkaline pH optima <cite>Harreither2011</cite>. No member From the phylogenetic group of Class-III CDHs has been characterized yet.
 
 
 
=== Catalytic Residues ===
 
The FAD cofactor in the flavodehydrogenase domain is noncovalently bound. The substrates anomeric carbon atom is oriented towards N5 of the FAD isoalloxazine ring at a distance of approx. 2.9 Å <cite>Hallberg2003</cite> and close to the general base histidine (HXXX in [{{PDBlink}}1kdg 1KDG], HYYY in [{{PDBlink}}4qm6 4QM6], ZZZ in [{{PDBlink}}4qm7 4QM7]). Important residues for substrate binding are XXDXX and essential for the interdomain electron transfer is R7XX <cite>Tan2015</cite>.
 
 
 
=== Three-dimensional structures ===
 
The crystal structure of the isolated cytochrome domain from ''P. chrysosporium'' CDH was reported in 2000 at a resolution of 1.9 Å (PDB ID [{{PDBlink}}1d7c 1D7C], <cite>Hallberg2000</cite>). The fragment folds into a compact globular structure which resembles the antibody fold of the heavy-chain Fab fragment. The propionate arm of the heme b is partly solvent exposed. The heme iron is hexa-coordinated by Met and His as axial ligands <cite>Cox1992</cite>. The crystal structure of the ''P. chrysosporium'' flavodehydrogenase domain (1.6 Å resolution, 1KDG) was reported in 2002 <cite>Hallberg2002</cite>. The flavin binding domain features the typical, flavin-binding βαβ-motif - the Rossmann-fold. The FAD moiety is non-covalently bound to the enzyme.
 
Full-length structures of ''Neurospora crassa'' and ''Myriococcum thermophilum'' CDH were reported in 2015 <cite>Tan2015</cite>. Structures of ''N. crassa'' CDH were resolved in the open-state, in which the cytochrome domain was spatially separated from the flavodehydrogenase domain. The structure of ''M. thermophilum'' CDH was resolved in the closed-state. It is suggested that the closed-state allows interdomain electron transfer from FAD to heme ''b'', whereas the subsequent reduction of LPMO occurs in the open-state, in which the heme ''b'' cofactor is accessible. CDH is a glycosylated enzyme featuring high-mannose type N-glycosides and O-glycosylation.
 
   
 
=== Family Firsts ===
 
;First CDH identified: In 1974 from the white-rot fungus ''Phanerochaete chrysosporium '' <cite>Westermark1974</cite>
 
;First demonstration of interaction with AA9 (LPMO): Patent [{{PatentLink}}US20100159536 US20100159536], and later published in more detail <cite>Langston2011</cite>
 
;First stereochemistry determination: The nature of the product was determined by XX... ().
 
;First catalytic nucleophile identification: <cite>Hallberg2002</cite>.
 
;First 3-D structure: ''P. chrysosporium'' cytochrome domain was crystallized in 2000 <cite>Hallberg2000</cite> ''P. chrysosporium'' flavodehydrogenase domain was resolved in 20022 <cite>Hallberg2002</cite>, first full structures were reported in 2015 in the open-state (''N. crassa'', <cite>Tan2015</cite>) and in the closed-state (''M. thermophilum'', <cite>Tan2015</cite>).
 
 
 
 
 
== Subfamily 2 - P2O ==
 
 
 
=== Kinetics and Mechanism ===
 
Pyranose 2-oxidase (P2O; syn. pyranose oxidase, POX; EC 1.1.3.10; pyranose:oxygen 2-oxidoreductase) is widespread in lignin-degrading white-rot fungi and catalyzes the oxygen dependent oxidation of several monosaccharides at the C2 position to the corresponding 2-keto-aldoses and H<sub>2</sub>O<sub>2</sub> <cite>WongnateandChaiyen2013</cite>. The preferred substrate is D-glucose, which is oxidized to 2-keto-D-glucose. Unlike other GMC-oxidoreductases, the FAD in P2O is covalently linked via its isoalloxazine ring 8α-methyl group to the N12 atom of His167 (histidyl linkage, <cite>Sucharitakul2008</cite>). The supposed physiological role of P2O is to generate H<sub>2</sub>O<sub>2</sub> as a substrate for lignin-degrading peroxidases. P2O is localized in the hyphal periplasmatic space <cite>Giffhorn2000</cite>.
 
 
 
During the reductive half reaction, two electrons are transferred from the substrate to the FAD, leading to the formation of FADH<sub>2</sub> and the 2-keto sugar. During the oxidative half reaction, the two stored electrons are transferred to molecular oxygen to form H<sub>2</sub>O<sub>2</sub>. The catalytic reaction is proposed to be of the ping-pong bi-bi type, since the 2-keto-sugar product is released prior to the reaction with oxygen <cite>WongnateandChaiyen2013</cite>. The active site contains a conserved, catalytic His–Asn pair positioned below the FAD isoalloxazine ring, which is typical for GMC oxidoreductases. His XXX (PDB) or His YYY (PDB) is the catalytic base that facilitates deprotonation of the sugar substrate.     
 
 
 
=== Catalytic Residues ===
 
The FAD cofactor in the flavodehydrogenase domain is covalently bound. The carbohydrate C2 atom is oriented towards N5 of the FAD isoalloxazine ring at a distance of approx. XY9 Å (PDB, Ref.3). Indispensible for catalysis is the general base histidine (HXXX in PDB, HYYY in PDB). A strictly conserved arginine residue (Argxxx in PDB, Arg YYY in PDB) is close to the caltalytic histidine <cite>WongnateandChaiyen2013</cite>.
 
 
 
=== Three-dimensional structures ===
 
Crystal structures of P2O from ''Trametes multicolor'' and ''Phanerochaete chrysosporium'' were resolved (pdb: [{{PDBlink}}1tt0 1TT0], <cite>Hallberg2004</cite>;pdb: [{{PDBlink}}4mig 4MIG]; <cite>Hassan2013</cite>) P2O is a homotetrameric enzyme with a molecular masses of approx. 68,000 per subunit. Each subunit contains the flavin-binding Rossmann-fold typical for GMC- oxidoreductases. The substrate-binding subdomain comprises six-stranded central-β sheet and three α-helices. The homotetramer features a large internal cavity, from which the four active sites are accessible. P2O is a non-glycosylated enzyme.
 
 
 
=== Family Firsts ===
 
;First P2O identified: Pyranose oxidase was first isolated from the basidiomycete Polyporus obtusus in 1968 <cite>Ruelius1968</cite>.
 
;First 3-D structure: P2O from Trametes multicolor MB 49 (pdb: [{{PDBlink}}1tt0 1TT0]; <cite>Hallberg2004</cite>)     
 
 
 
   
 
== Subfamily 3 - GOx ==
 
 
 
=== Kinetics and Mechanism ===
 
Glucose oxidase (GOx, syn. GOD, EC 1.1.x.x., glucose:oxygen 1-oxidoreductase) catalyzes the regioselective oxidation of β-D-glucose to gluconic acid by utilizing molecular oxygen as an electron acceptor with simultaneous production of hydrogen peroxide. Glucose oxidases are produced by a number of fungi and insects. Proposed functions are related to the peroxide-producing abilities of the enzyme and include the preservation of honey, microbial defense as well as the support of H2O2 dependent ligninases in wood degrading fungi <cite>Wong2008</cite>.
 
 
 
In the oxidative half reaction of GOx, two-electron oxidation of D-glucose at C1 results in the formation of gluconolactone and reduced cofactor (FADH<sub>2</sub>). Reduction by glucose occurs via a hydride ion transfer to the N-5 position of the FAD. Oxygen reduction to H<sub>2</sub>O<sub>2</sub> involves the uptake of two electrons (from FADH<sub>2</sub>) and two protons. A highly conserved His (His516 in ''Aspergillus niger'' GOx) in the active site of the enzyme was identified by site directed mutagenesis as the catalytic base <cite>Roth2003</cite>.     
 
 
 
=== Catalytic Residues ===
 
Content is to be added here.
 
 
 
=== Three-dimensional structures ===
 
Content is to be added here.
 
 
 
=== Family Firsts ===
 
;First GOx identified: Glucose oxidase from Aspergillus niger in 1928 <cite>Mueller1928</cite>.
 
;First 3D structure: Glucose oxidase from ''Aspergillus niger'' ([{{PDBlink}}1gal 1GAL]; <cite>Hecht1993</cite>).
 
  
 +
==== Glucose oxidase and glucose dehydrogenase ====
 +
Glucose oxidases (GOX) (Figure 2D) and glucose dehydrogenases (GDH) catalyse the regioselective oxidation of β-<small>D</small>-glucose to <small>D</small>-glucono-1,5-lactone. Both GOX and GDH are structurally related, including a set of conserved active site residues for the binding of glucose. Therefore, a discrimination of these genes based on their sequences is difficult <cite>Suetzl2018</cite>. Glucose oxidases typically occur as homodimers whereas GDHs occur as monomers or dimers. The first crystal structure of GOX from ''Aspergillus niger'' was resolved in 1993 (pdb: [{{PDBlink}}1gal 1GAL] <cite>Hecht1993</cite>). A highly conserved arrangement of active-site residues, which form hydrogen bonds to all five hydroxyl groups of β-<small>D</small>-glucose, is responsible for the particularly high specificity towards <small>D</small>-glucose. The first crystal structure of GDH from ''Aspergillus flavus'' was reported in 2015, showing a slightly less conserved active site and fewer interactions with glucose (pdb: [{{PDBlink}}4ynt 4YNT] <cite>Yoshida2015</cite>). This may also explain the promiscuous reaction of GDH with β-<small>D</small>-xylose, which is not observed for GOX. Both enzymes show different cosubstrate specificities. Glucose oxidases reduce atmospheric oxygen to hydrogen peroxide in their oxidative half-reaction. Proposed native functions of the enzyme are closely related to its peroxide-producing abilities and include the preservation of honey, microbial defense as well as the support of H<sub>2</sub>O<sub>2</sub>-dependent ligninases in wood degrading fungi <cite>Wong2008</cite>. GDHs interact very slowly with oxygen and prefer electron acceptors like quinones or phenoxy-radicals. The detailed physiological role of GDH remains elusive, but has been previously linked to the neutralization of plant laccase activity by fungi during plant infection <cite>Sygmund2011</cite>.
  
 +
==== Pyranose dehydrogenase ====
 +
Pyranose dehydrogenase (PDH) is a secretory protein found in a limited number of litter degrading fungi belonging to the families Agaricaceae and Lycoperdaceae <cite> Volc2001 </cite>. While the overall sequential and structural similarities to other GMC-oxidoreductases are high, the enzyme differs in terms of its catalytic properties and contains a covalently linked FAD, which is not observed in other AA3_2 enzymes. PDHs are characterised by a broad substrate spectrum, which includes a number of monosaccharides (<small>L</small>-arabinose, <small>D</small>-glucose, and <small>D</small>-galactose, <small>D</small>-xylose) as well as some oligo- or polysaccharides. PDHs catalyse the C1, C2, or C3 oxidation of the sugar substrate, or introduce di-oxidations at C2/C3 or C3/C4, resulting in the formation of aldonolactones (C1 oxidation) or (di)ketosugars. The first crystal structure of the enzyme was resolved in 2013 (pdb: [{{PDBlink}}4h7u 4H7U]) and showed an open active site conformation, which in part may also explain its high substrate promiscuity <cite> Tan2013 </cite>. The electron acceptor specificity is similar to those of other sugar dehydrogenases and includes quinones and phenoxy radicals.
 +
 +
== Subfamily AA3_3: Alcohol oxidase ==
 +
Alcohol- or methanol oxidases (AOX) (Figure 2D) are intra- or extracellular enzymes found in yeasts and fungi. In methylotrophic yeasts, such as ''Pichia pastoris'', AOX is a catabolic enzyme essential for the utilization of methanol. AOX catalyses the selective oxidation of primary alcohols at CH-OH, yielding the corresponding carbonyl compounds <cite> Ozimek2005 </cite>. The enzyme accepts both saturated and unsaturated aliphatic alcohols with a chain length from C1 to C8, but shows a strong discrimination against secondary alcohols. To date, two structures of AOX from ''P. pastoris'' were resolved in 2016; One based on X-ray diffraction (pdb: [{{PDBlink}}4hsa 5HSA] <cite> Koch2016 </cite>), the other on cyro-electron microscopy (pdb: [{{PDBlink}}5i68 5I68] <cite> Vonck2016 </cite>). AOX are homo-octameric enzymes located in the peroxisomal matrix <cite> Ozimek2005 </cite>. Each subunit contains a non-covalently bound FAD-molecule modified with an arabinyl chain. In ''P. pastoris'', the degree of the FAD arabinylation depends on the alcohol concentration in the medium and is thought to modify the reactivity with methanol. Considerably less is known about fungal AOX, which have been found in wood-degrading or phytopathogenic fungi from the phylum Basidiomycota. Similarly to AOX from yeast, they showed an octameric architecture and displayed the highest catalytic efficiencies for methanol. The fungal AOX from ''Gloeophyllum trabeum'' was identified as extracellular protein despite the lack of a dedicated secretion signal peptide <cite> Daniel2007 </cite>. The native function of fungal AOX could therefore be related to its peroxide production, which, similar to other AA3_2 enzymes, can potentially stimulate fungal attack on lignocellulose by supplying H<sub>2</sub>O<sub>2</sub> to peroxidases. AOX secreted by the phytopathogenic basidiomycete ''Moniliophthora perniciosa'', the causative agent of Witches’ broom disease in the cocoa tree, is thought to play a role in the utilization of methanol derived from pectin demethylation <cite>Oliveira2012</cite>.
  
 +
== Subfamily AA3_4: Pyranose oxidase ==
 +
Pyranose oxidases (POX) (Figure 2B) are widespread in lignocellulose degrading fungi and catalyse the C2-oxidation of monosaccharides. POX is the most distantly related member of the AA3 family and does not show the strict conservation of structural motifs found within the AA3 family. Unlike other AA3 enzymes, POXs are associated with membrane-bound vesicles and other membrane structures in the periplasmic space of the fungal hyphae. The first crystal structure of POX was reported in 2004 (''Trametes multicolor'', pdb: [{{PDBlink}}1tt0 1TT0] <cite> Hallberg2004 </cite>). The enzyme is a homo-tetramer with each of the subunits carrying a covalently bound FAD molecule. A distinct “head domain” on each of the subdomains is thought to be involved in oligomerization or in interactions with cell wall-polysaccharides <cite> Hallberg2004 </cite>. Access to the active site of the enzyme is modulated by a flexible active-site loop, which hinders entrance of oligosaccharides. Substrate channels lead from the polypeptide surface to an internal large cavity formed by the four subunits. The preferred substrate of POX is either α- or β-<small>D</small>-glucose, but also <small>D</small>-galactose, <small>D</small>-xylose, or <small>D</small>-glucono-1,5-lactone are converted <cite> Leitner2001 </cite>. Substrates are oxidised at the C2 position to yield 2-ketoaldoses as products. In the oxidative half-reaction, the FADH<sub>2</sub> is regenerated by reduction of O<sub>2</sub> to H<sub>2</sub>O<sub>2</sub>. The reduction of oxygen by POX proceeds through a C4a-hydroperoxyflavin intermediate, which is characteristic for FAD-dependent monooxygenases but is untypical for sugar oxidases <cite> Sucharitakul2008 </cite>. Apart from oxygen, POX can also utilize alternative electron acceptors including a number of (substituted) quinones and (complexed) metal ions. Of note, catalytic efficiencies for some of these electron acceptors are higher than for oxygen, suggesting that these acceptors might be biologically more relevant substrates than oxygen. Recently, putative POX sequences were identified in Actinobacteria with overall sequence identities of 39–24% to fungal POX <cite> Mendes2017 </cite>. A functional POX from ''Arthrobacter siccitolerans'' was recombinantly expressed and characterized <cite> Mendes2017 </cite>. POX activity has been also confirmed in the supernatant of the bacterium ''Pantoea ananatis'' when cultivated on rice straw <cite> Ma2016 </cite>.
  
 
== References ==
 
== References ==
 
<biblio>
 
<biblio>
#Zamocky2006 pmid=16787264
+
#Levasseur2013 pmid=23514094
 +
#Suetzl2018 pmid=29411063
 +
#Ferreira2015 pmid=26297778
 +
#Hernandez2012 pmid=22249717
 +
#Daniel1994 pmid=16349330
 +
#Langston2011 pmid=21821740
 +
#Kracher2016 pmid=27127235
 +
#Garajova2016 pmid=27312718
 +
#Bissaro2017 pmid=28846668
 +
#Goswami2013 pmid=23525937
 +
#Iida2007 pmid=17498303
 +
#Sun2012 pmid=23022604
 +
#Mendes2017 Mendes S, Banha C, Madeira J, Santos D, Miranda V, Manzanera M, Ventura MR, van Berkel WJH, Martins LO. ''Characterization of a bacterial pyranose 2-oxidase from Arthrobacter siccitolerans.'' 2016 J Mol Catal B Enzym  133, S34–S43
 +
#Westermark1974 Westermark U, Eriksson KE. ''Cellobiose:Quinone Oxidoreductase, a New Wood-degrading Enzyme from White-rot Fungi.'' 1974 Acta Chem Scand 1974 28b:209–214.
 +
#Kracher2016b Kracher D, Ludwig R. ''Cellobiose dehydrogenase: An essential enzyme for lignocellulose degradation in nature - A review.'' 2016 Bodenkultur 67:145–163.
 +
#Hallberg2002 pmid=12493734
 
#Phillips2011 pmid=22004347
 
#Phillips2011 pmid=22004347
#Loose2016 pmid=27643617
 
#Zhang2011 pmid=20709172
 
#Henriksson1998 pmid=9546045
 
#Hallberg2003 pmid=12493734
 
#WongnateandChaiyen2013 pmid=23578136
 
#Igarashi2002 pmid=11939907
 
#Zamocky2004 pmid=15302401
 
 
#Tan2015 pmid=26151670
 
#Tan2015 pmid=26151670
#Hallberg2000 pmid=10673428
+
#Courtade2016 pmid=27152023
#Cox1992 pmid=1322830
+
#Bodenheimer2018 pmid=29374564
#Hallberg2002 pmid=12493734
+
#Fernandez2009 pmid=19923715
#Westermark1974 Westermark U & Eriksson K-E ''Cellobiose:Quinone Oxidoreductase, a New Wood-degrading Enzyme from White-rot Fungi.'' Acta Chem. Scand. 1974 28b, 209–214.
+
#Mathieu2016 pmid=26873317
#Langston2011 pmid=21821740
 
#Sucharitakul2008 pmid=18652479
 
#Giffhorn2000 pmid=11152063
 
#Ruelius1968 pmid=5725162
 
#Mueller1928 Mueller D ''Oxidation von Glukose mit Extrakten aus Aspegillus niger''. Biochem Z 199, 136–170
 
 
#Hecht1993 pmid=8421298
 
#Hecht1993 pmid=8421298
 +
#Yoshida2015 pmid=26311535
 +
#Wong2008 pmid=18330562
 +
#Sygmund2011 pmid=21903757
 +
#Volc2001 pmid=11511865
 +
#Tan2013 pmid=23326459
 +
#Ozimek2005 pmid=16169288
 +
#Koch2016 pmid=26905908
 +
#Vonck2016 pmid=27458710
 +
#Daniel2007 pmid=17660304
 +
#Oliveira2012 pmid=23022488
 
#Hallberg2004 pmid=15288786
 
#Hallberg2004 pmid=15288786
#Hassan2013 pmid=24282677
+
#Leitner2001 pmid=11472941
#Wong2008 pmid=18330562
+
#Sucharitakul2008 pmid=18652479
#Roth2003 pmid=12506204
+
#Ma2016 pmid=27761153
</cite></biblio>
+
 
 +
</biblio>
  
  
  
 
[[Category:Auxiliary Activity Families|AA003]]
 
[[Category:Auxiliary Activity Families|AA003]]

Latest revision as of 14:17, 18 December 2021

Approve icon-50px.png

This page has been approved by the Responsible Curator as essentially complete. CAZypedia is a living document, so further improvement of this page is still possible. If you would like to suggest an addition or correction, please contact the page's Responsible Curator directly by e-mail.


Auxiliary Activity Family AA3
Clan AA3
Mechanism FAD-dependent substrate oxidation
Active site residues known
CAZy DB link
http://www.cazy.org/AA3.html


General properties and substrate specificities

Figure 1. Reaction cycle of AA3 flavoenzymes. In the reductive half-reaction substrate oxidation leads to the formation of a two-electron reduced FADH2 (flavohydroquinone). During the oxidative half-reaction the flavin cofactor is re-oxidised via electron transfer to a suitable electron acceptor. Exemplary reactions for each AA3 subfamily are shown.

Enzymes of the CAZy family AA3 are widespread and catalyse the oxidation of alcohols or carbohydrates with the concomitant formation of hydrogen peroxide or hydroquinones [1]. AA3 enzymes are most abundant in wood-degrading fungi where they typically display a high multigenicity [2]. The main function of fungal AA3 enzymes is the stimulation of lignocellulose degradation in cooperation with other AA-enzymes such as peroxidases (AA2) [3, 4, 5] or lytic polysaccharide monooxygenases (AA9) [6, 7, 8, 9]. In yeast, AA3 enzymes are involved in the catabolism of alcohols [10], and some AA3 genes identified in insects are thought to be relevant for immunity and development [11, 12]. Recently, a bacterial AA3 enzyme with unknown biological function was isolated and characterized [13]. The functionally diverse enzymes of family AA3 all belong to the structurally related glucose-methanol-choline (GMC) family of oxidoreductases and require a flavin-adenine dinucleotide (FAD) cofactor for catalytic activity. Based on their sequences members of the AA3 family were divided into four subfamilies in the CAZy database (Figure 1). Family AA3_1 contains the flavodehydrogenase domain of cellobiose dehydrogenase (EC 1.1.99.18), family AA3_2 includes aryl alcohol oxidase (EC 1.1.3.7), glucose oxidase (EC 1.1.3.4), glucose dehydrogenase (EC 1.1.5.9) and pyranose dehydrogenase (EC 1.1.99.29), family AA3_3 consists of alcohol (methanol) oxidases (EC 1.1.3.13) and family AA3_4 comprises pyranose dehydrogenases(EC 1.1.3.10).

Subfamily AA3_1: Cellobiose dehydrogenase

Cellobiose dehydrogenases (CDHs) are extracellular flavocytochromes that were first described in 1974 [14]. CDHs are exclusively found in wood-degrading and phytopathogenic fungi belonging to the phyla Basidiomycota (Class-I CDHs) and Ascomycota (Class-II and -III CDHs) [15]. They oxidise a wide variety of lignocellulose-derived saccharides to their corresponding sugar lactones. CDHs show a high preference for soluble, β-(1,4)-interlinked saccharides and scarcely oxidise monosaccharides. A common feature of all CDHs is their complex bipartite structure, which comprises an N-terminal cytochrome-binding domain (CYT) and a larger, catalytic flavodehydrogenase (DH) domain encoded within a single polypeptide chain [16] (Figure 2A). Both domains are connected by a flexible linker which typically comprises 15 – 30 amino acids. An important in vivo function of CDH is the reduction and activation of family AA9 lytic polysaccharide monooxygenases via its haem b domain [17, 18, 19]. Recently, the holoenzyme structures of Neurospora crassa CDH (pdb: 4QI7) and Myriococcum thermophilum CDH (pdb: 4QI6) were reported. Two structures of N. crassa CDH showed an “open-state” conformation in which DH and CYT were spatially separated, whereas a structure of M. thermophilum CDH showed a “closed-state” conformation in which the propionate arm of the cytochrome domain interacted with the catalytic centre in DH [18]. Analysis by small angle scattering also suggested a number of possible intermediate conformers that exist in solution [18, 20]. While the closed-state allows interdomain electron transfer from DH to CYT, reduction of electron acceptors (e.g. AA9 enzymes) might occur in the open-state, in which the haem cofactor is fully accessible.

Subfamily AA3_2

Aryl alcohol oxidase/dehydrogenase

Figure 2. Structures of AA3 family members. A, Cellobiose dehydrogenase from Myriococcum thermophilum (pdb: 4QI6). The haem b-containing domain is shown in red. B, pyranose oxidase from Phanerochaete chrysosporium (pdb: 4MIF); C, aryl-alcohol oxidase from Pleurotus eryngii (pdb: 3FIM); D, methanol oxidase from Pichia pastoris (pdb: 5HSA); E, glucose oxidase from Aspergillus niger (pdb: 4CF4). The GMC-oxidoreductase α/β-fold is colour-coded.

Aryl alcohol oxidoreductases (AAO) are secreted by basidiomycete fungi and proposed to play a crucial role in lignin degradation. The X-ray crystal structure of the most widely studied AAO from Pleurotus eryngii (pdb: 3FIM) showed a non-covalently bound FAD cofactor in the active-site [21] (Figure 2C). Access to the active site is restricted by three aromatic residues, which interact with both the alcohol substrate and oxygen. The substrate scope of AAO includes secreted benzyl alcohols from the secondary metabolism (e.g. veratryl alcohol), or related alcohols that accumulate during lignin degradation [4]. During the reductive half-reaction, the primary alcohol of the substrate is two-electron oxidised to form the corresponding aldehyde. In addition, aromatic aldehydes can undergo further oxidation yielding the corresponding acids. The concomitantly formed hydrogen peroxide is considered essential for the activity of lignin degrading peroxidases. Recently, aryl-alcohol quinone oxidoreductases (AAQO) with a high sequence identity to Pleurotus eryngii AAO were identified in the genome of the basidiomycete Pycnoporus cinnabarinus [22]. While these enzymes showed a substrate spectrum similar to known AAOs, oxygen reduction was insignificant or very low relative to other characterised AAOs. However, reoxidation of FADH2 was very efficient with benzoquinone or phenoxy-radicals, which are formed by fungal laccases, suggesting a different functional role of AAQOs.

Glucose oxidase and glucose dehydrogenase

Glucose oxidases (GOX) (Figure 2D) and glucose dehydrogenases (GDH) catalyse the regioselective oxidation of β-D-glucose to D-glucono-1,5-lactone. Both GOX and GDH are structurally related, including a set of conserved active site residues for the binding of glucose. Therefore, a discrimination of these genes based on their sequences is difficult [2]. Glucose oxidases typically occur as homodimers whereas GDHs occur as monomers or dimers. The first crystal structure of GOX from Aspergillus niger was resolved in 1993 (pdb: 1GAL [23]). A highly conserved arrangement of active-site residues, which form hydrogen bonds to all five hydroxyl groups of β-D-glucose, is responsible for the particularly high specificity towards D-glucose. The first crystal structure of GDH from Aspergillus flavus was reported in 2015, showing a slightly less conserved active site and fewer interactions with glucose (pdb: 4YNT [24]). This may also explain the promiscuous reaction of GDH with β-D-xylose, which is not observed for GOX. Both enzymes show different cosubstrate specificities. Glucose oxidases reduce atmospheric oxygen to hydrogen peroxide in their oxidative half-reaction. Proposed native functions of the enzyme are closely related to its peroxide-producing abilities and include the preservation of honey, microbial defense as well as the support of H2O2-dependent ligninases in wood degrading fungi [25]. GDHs interact very slowly with oxygen and prefer electron acceptors like quinones or phenoxy-radicals. The detailed physiological role of GDH remains elusive, but has been previously linked to the neutralization of plant laccase activity by fungi during plant infection [26].

Pyranose dehydrogenase

Pyranose dehydrogenase (PDH) is a secretory protein found in a limited number of litter degrading fungi belonging to the families Agaricaceae and Lycoperdaceae [27]. While the overall sequential and structural similarities to other GMC-oxidoreductases are high, the enzyme differs in terms of its catalytic properties and contains a covalently linked FAD, which is not observed in other AA3_2 enzymes. PDHs are characterised by a broad substrate spectrum, which includes a number of monosaccharides (L-arabinose, D-glucose, and D-galactose, D-xylose) as well as some oligo- or polysaccharides. PDHs catalyse the C1, C2, or C3 oxidation of the sugar substrate, or introduce di-oxidations at C2/C3 or C3/C4, resulting in the formation of aldonolactones (C1 oxidation) or (di)ketosugars. The first crystal structure of the enzyme was resolved in 2013 (pdb: 4H7U) and showed an open active site conformation, which in part may also explain its high substrate promiscuity [28]. The electron acceptor specificity is similar to those of other sugar dehydrogenases and includes quinones and phenoxy radicals.

Subfamily AA3_3: Alcohol oxidase

Alcohol- or methanol oxidases (AOX) (Figure 2D) are intra- or extracellular enzymes found in yeasts and fungi. In methylotrophic yeasts, such as Pichia pastoris, AOX is a catabolic enzyme essential for the utilization of methanol. AOX catalyses the selective oxidation of primary alcohols at CH-OH, yielding the corresponding carbonyl compounds [29]. The enzyme accepts both saturated and unsaturated aliphatic alcohols with a chain length from C1 to C8, but shows a strong discrimination against secondary alcohols. To date, two structures of AOX from P. pastoris were resolved in 2016; One based on X-ray diffraction (pdb: 5HSA [30]), the other on cyro-electron microscopy (pdb: 5I68 [31]). AOX are homo-octameric enzymes located in the peroxisomal matrix [29]. Each subunit contains a non-covalently bound FAD-molecule modified with an arabinyl chain. In P. pastoris, the degree of the FAD arabinylation depends on the alcohol concentration in the medium and is thought to modify the reactivity with methanol. Considerably less is known about fungal AOX, which have been found in wood-degrading or phytopathogenic fungi from the phylum Basidiomycota. Similarly to AOX from yeast, they showed an octameric architecture and displayed the highest catalytic efficiencies for methanol. The fungal AOX from Gloeophyllum trabeum was identified as extracellular protein despite the lack of a dedicated secretion signal peptide [32]. The native function of fungal AOX could therefore be related to its peroxide production, which, similar to other AA3_2 enzymes, can potentially stimulate fungal attack on lignocellulose by supplying H2O2 to peroxidases. AOX secreted by the phytopathogenic basidiomycete Moniliophthora perniciosa, the causative agent of Witches’ broom disease in the cocoa tree, is thought to play a role in the utilization of methanol derived from pectin demethylation [33].

Subfamily AA3_4: Pyranose oxidase

Pyranose oxidases (POX) (Figure 2B) are widespread in lignocellulose degrading fungi and catalyse the C2-oxidation of monosaccharides. POX is the most distantly related member of the AA3 family and does not show the strict conservation of structural motifs found within the AA3 family. Unlike other AA3 enzymes, POXs are associated with membrane-bound vesicles and other membrane structures in the periplasmic space of the fungal hyphae. The first crystal structure of POX was reported in 2004 (Trametes multicolor, pdb: 1TT0 [34]). The enzyme is a homo-tetramer with each of the subunits carrying a covalently bound FAD molecule. A distinct “head domain” on each of the subdomains is thought to be involved in oligomerization or in interactions with cell wall-polysaccharides [34]. Access to the active site of the enzyme is modulated by a flexible active-site loop, which hinders entrance of oligosaccharides. Substrate channels lead from the polypeptide surface to an internal large cavity formed by the four subunits. The preferred substrate of POX is either α- or β-D-glucose, but also D-galactose, D-xylose, or D-glucono-1,5-lactone are converted [35]. Substrates are oxidised at the C2 position to yield 2-ketoaldoses as products. In the oxidative half-reaction, the FADH2 is regenerated by reduction of O2 to H2O2. The reduction of oxygen by POX proceeds through a C4a-hydroperoxyflavin intermediate, which is characteristic for FAD-dependent monooxygenases but is untypical for sugar oxidases [36]. Apart from oxygen, POX can also utilize alternative electron acceptors including a number of (substituted) quinones and (complexed) metal ions. Of note, catalytic efficiencies for some of these electron acceptors are higher than for oxygen, suggesting that these acceptors might be biologically more relevant substrates than oxygen. Recently, putative POX sequences were identified in Actinobacteria with overall sequence identities of 39–24% to fungal POX [13]. A functional POX from Arthrobacter siccitolerans was recombinantly expressed and characterized [13]. POX activity has been also confirmed in the supernatant of the bacterium Pantoea ananatis when cultivated on rice straw [37].

References

  1. Levasseur A, Drula E, Lombard V, Coutinho PM, and Henrissat B. (2013). Expansion of the enzymatic repertoire of the CAZy database to integrate auxiliary redox enzymes. Biotechnol Biofuels. 2013;6(1):41. DOI:10.1186/1754-6834-6-41 | PubMed ID:23514094 [Levasseur2013]
  2. Sützl L, Laurent CVFP, Abrera AT, Schütz G, Ludwig R, and Haltrich D. (2018). Multiplicity of enzymatic functions in the CAZy AA3 family. Appl Microbiol Biotechnol. 2018;102(6):2477-2492. DOI:10.1007/s00253-018-8784-0 | PubMed ID:29411063 [Suetzl2018]
  3. Ferreira P, Carro J, Serrano A, and Martínez AT. (2015). A survey of genes encoding H2O2-producing GMC oxidoreductases in 10 Polyporales genomes. Mycologia. 2015;107(6):1105-19. DOI:10.3852/15-027 | PubMed ID:26297778 [Ferreira2015]
  4. Hernández-Ortega A, Ferreira P, and Martínez AT. (2012). Fungal aryl-alcohol oxidase: a peroxide-producing flavoenzyme involved in lignin degradation. Appl Microbiol Biotechnol. 2012;93(4):1395-410. DOI:10.1007/s00253-011-3836-8 | PubMed ID:22249717 [Hernandez2012]
  5. Daniel G, Volc J, and Kubatova E. (1994). Pyranose Oxidase, a Major Source of H(2)O(2) during Wood Degradation by Phanerochaete chrysosporium, Trametes versicolor, and Oudemansiella mucida. Appl Environ Microbiol. 1994;60(7):2524-32. DOI:10.1128/aem.60.7.2524-2532.1994 | PubMed ID:16349330 [Daniel1994]
  6. Langston JA, Shaghasi T, Abbate E, Xu F, Vlasenko E, and Sweeney MD. (2011). Oxidoreductive cellulose depolymerization by the enzymes cellobiose dehydrogenase and glycoside hydrolase 61. Appl Environ Microbiol. 2011;77(19):7007-15. DOI:10.1128/AEM.05815-11 | PubMed ID:21821740 [Langston2011]
  7. Kracher D, Scheiblbrandner S, Felice AK, Breslmayr E, Preims M, Ludwicka K, Haltrich D, Eijsink VG, and Ludwig R. (2016). Extracellular electron transfer systems fuel cellulose oxidative degradation. Science. 2016;352(6289):1098-101. DOI:10.1126/science.aaf3165 | PubMed ID:27127235 [Kracher2016]
  8. Garajova S, Mathieu Y, Beccia MR, Bennati-Granier C, Biaso F, Fanuel M, Ropartz D, Guigliarelli B, Record E, Rogniaux H, Henrissat B, and Berrin JG. (2016). Single-domain flavoenzymes trigger lytic polysaccharide monooxygenases for oxidative degradation of cellulose. Sci Rep. 2016;6:28276. DOI:10.1038/srep28276 | PubMed ID:27312718 [Garajova2016]
  9. Bissaro B, Røhr ÅK, Müller G, Chylenski P, Skaugen M, Forsberg Z, Horn SJ, Vaaje-Kolstad G, and Eijsink VGH. (2017). Oxidative cleavage of polysaccharides by monocopper enzymes depends on H(2)O(2). Nat Chem Biol. 2017;13(10):1123-1128. DOI:10.1038/nchembio.2470 | PubMed ID:28846668 [Bissaro2017]
  10. Goswami P, Chinnadayyala SS, Chakraborty M, Kumar AK, and Kakoti A. (2013). An overview on alcohol oxidases and their potential applications. Appl Microbiol Biotechnol. 2013;97(10):4259-75. DOI:10.1007/s00253-013-4842-9 | PubMed ID:23525937 [Goswami2013]
  11. Iida K, Cox-Foster DL, Yang X, Ko WY, and Cavener DR. (2007). Expansion and evolution of insect GMC oxidoreductases. BMC Evol Biol. 2007;7:75. DOI:10.1186/1471-2148-7-75 | PubMed ID:17498303 [Iida2007]
  12. Sun W, Shen YH, Yang WJ, Cao YF, Xiang ZH, and Zhang Z. (2012). Expansion of the silkworm GMC oxidoreductase genes is associated with immunity. Insect Biochem Mol Biol. 2012;42(12):935-45. DOI:10.1016/j.ibmb.2012.09.006 | PubMed ID:23022604 [Sun2012]
  13. Mendes S, Banha C, Madeira J, Santos D, Miranda V, Manzanera M, Ventura MR, van Berkel WJH, Martins LO. Characterization of a bacterial pyranose 2-oxidase from Arthrobacter siccitolerans. 2016 J Mol Catal B Enzym 133, S34–S43

    [Mendes2017]
  14. Westermark U, Eriksson KE. Cellobiose:Quinone Oxidoreductase, a New Wood-degrading Enzyme from White-rot Fungi. 1974 Acta Chem Scand 1974 28b:209–214.

    [Westermark1974]
  15. Kracher D, Ludwig R. Cellobiose dehydrogenase: An essential enzyme for lignocellulose degradation in nature - A review. 2016 Bodenkultur 67:145–163.

    [Kracher2016b]
  16. Hallberg BM, Henriksson G, Pettersson G, Vasella A, and Divne C. (2003). Mechanism of the reductive half-reaction in cellobiose dehydrogenase. J Biol Chem. 2003;278(9):7160-6. DOI:10.1074/jbc.M210961200 | PubMed ID:12493734 [Hallberg2002]
  17. Phillips CM, Beeson WT, Cate JH, and Marletta MA. (2011). Cellobiose dehydrogenase and a copper-dependent polysaccharide monooxygenase potentiate cellulose degradation by Neurospora crassa. ACS Chem Biol. 2011;6(12):1399-406. DOI:10.1021/cb200351y | PubMed ID:22004347 [Phillips2011]
  18. Tan TC, Kracher D, Gandini R, Sygmund C, Kittl R, Haltrich D, Hällberg BM, Ludwig R, and Divne C. (2015). Structural basis for cellobiose dehydrogenase action during oxidative cellulose degradation. Nat Commun. 2015;6:7542. DOI:10.1038/ncomms8542 | PubMed ID:26151670 [Tan2015]
  19. Courtade G, Wimmer R, Røhr ÅK, Preims M, Felice AK, Dimarogona M, Vaaje-Kolstad G, Sørlie M, Sandgren M, Ludwig R, Eijsink VG, and Aachmann FL. (2016). Interactions of a fungal lytic polysaccharide monooxygenase with β-glucan substrates and cellobiose dehydrogenase. Proc Natl Acad Sci U S A. 2016;113(21):5922-7. DOI:10.1073/pnas.1602566113 | PubMed ID:27152023 [Courtade2016]
  20. Bodenheimer AM, O'Dell WB, Oliver RC, Qian S, Stanley CB, and Meilleur F. (2018). Structural investigation of cellobiose dehydrogenase IIA: Insights from small angle scattering into intra- and intermolecular electron transfer mechanisms. Biochim Biophys Acta Gen Subj. 2018;1862(4):1031-1039. DOI:10.1016/j.bbagen.2018.01.016 | PubMed ID:29374564 [Bodenheimer2018]
  21. Fernández IS, Ruíz-Dueñas FJ, Santillana E, Ferreira P, Martínez MJ, Martínez AT, and Romero A. (2009). Novel structural features in the GMC family of oxidoreductases revealed by the crystal structure of fungal aryl-alcohol oxidase. Acta Crystallogr D Biol Crystallogr. 2009;65(Pt 11):1196-205. DOI:10.1107/S0907444909035860 | PubMed ID:19923715 [Fernandez2009]
  22. Mathieu Y, Piumi F, Valli R, Aramburu JC, Ferreira P, Faulds CB, and Record E. (2016). Activities of Secreted Aryl Alcohol Quinone Oxidoreductases from Pycnoporus cinnabarinus Provide Insights into Fungal Degradation of Plant Biomass. Appl Environ Microbiol. 2016;82(8):2411-2423. DOI:10.1128/AEM.03761-15 | PubMed ID:26873317 [Mathieu2016]
  23. Hecht HJ, Kalisz HM, Hendle J, Schmid RD, and Schomburg D. (1993). Crystal structure of glucose oxidase from Aspergillus niger refined at 2.3 A resolution. J Mol Biol. 1993;229(1):153-72. DOI:10.1006/jmbi.1993.1015 | PubMed ID:8421298 [Hecht1993]
  24. Yoshida H, Sakai G, Mori K, Kojima K, Kamitori S, and Sode K. (2015). Structural analysis of fungus-derived FAD glucose dehydrogenase. Sci Rep. 2015;5:13498. DOI:10.1038/srep13498 | PubMed ID:26311535 [Yoshida2015]
  25. Wong CM, Wong KH, and Chen XD. (2008). Glucose oxidase: natural occurrence, function, properties and industrial applications. Appl Microbiol Biotechnol. 2008;78(6):927-38. DOI:10.1007/s00253-008-1407-4 | PubMed ID:18330562 [Wong2008]
  26. Sygmund C, Klausberger M, Felice AK, and Ludwig R. (2011). Reduction of quinones and phenoxy radicals by extracellular glucose dehydrogenase from Glomerella cingulata suggests a role in plant pathogenicity. Microbiology (Reading). 2011;157(Pt 11):3203-3212. DOI:10.1099/mic.0.051904-0 | PubMed ID:21903757 [Sygmund2011]
  27. Volc J, Kubátová E, Daniel G, Sedmera P, and Haltrich D. (2001). Screening of basidiomycete fungi for the quinone-dependent sugar C-2/C-3 oxidoreductase, pyranose dehydrogenase, and properties of the enzyme from Macrolepiota rhacodes. Arch Microbiol. 2001;176(3):178-86. DOI:10.1007/s002030100308 | PubMed ID:11511865 [Volc2001]
  28. Tan TC, Spadiut O, Wongnate T, Sucharitakul J, Krondorfer I, Sygmund C, Haltrich D, Chaiyen P, Peterbauer CK, and Divne C. (2013). The 1.6 Å crystal structure of pyranose dehydrogenase from Agaricus meleagris rationalizes substrate specificity and reveals a flavin intermediate. PLoS One. 2013;8(1):e53567. DOI:10.1371/journal.pone.0053567 | PubMed ID:23326459 [Tan2013]
  29. Ozimek P, Veenhuis M, and van der Klei IJ. (2005). Alcohol oxidase: a complex peroxisomal, oligomeric flavoprotein. FEMS Yeast Res. 2005;5(11):975-83. DOI:10.1016/j.femsyr.2005.06.005 | PubMed ID:16169288 [Ozimek2005]
  30. Koch C, Neumann P, Valerius O, Feussner I, and Ficner R. (2016). Crystal Structure of Alcohol Oxidase from Pichia pastoris. PLoS One. 2016;11(2):e0149846. DOI:10.1371/journal.pone.0149846 | PubMed ID:26905908 [Koch2016]
  31. Vonck J, Parcej DN, and Mills DJ. (2016). Structure of Alcohol Oxidase from Pichia pastoris by Cryo-Electron Microscopy. PLoS One. 2016;11(7):e0159476. DOI:10.1371/journal.pone.0159476 | PubMed ID:27458710 [Vonck2016]
  32. Daniel G, Volc J, Filonova L, Plíhal O, Kubátová E, and Halada P. (2007). Characteristics of Gloeophyllum trabeum alcohol oxidase, an extracellular source of H2O2 in brown rot decay of wood. Appl Environ Microbiol. 2007;73(19):6241-53. DOI:10.1128/AEM.00977-07 | PubMed ID:17660304 [Daniel2007]
  33. de Oliveira BV, Teixeira GS, Reis O, Barau JG, Teixeira PJ, do Rio MC, Domingues RR, Meinhardt LW, Paes Leme AF, Rincones J, and Pereira GA. (2012). A potential role for an extracellular methanol oxidase secreted by Moniliophthora perniciosa in Witches' broom disease in cacao. Fungal Genet Biol. 2012;49(11):922-32. DOI:10.1016/j.fgb.2012.09.001 | PubMed ID:23022488 [Oliveira2012]
  34. Hallberg BM, Leitner C, Haltrich D, and Divne C. (2004). Crystal structure of the 270 kDa homotetrameric lignin-degrading enzyme pyranose 2-oxidase. J Mol Biol. 2004;341(3):781-96. DOI:10.1016/j.jmb.2004.06.033 | PubMed ID:15288786 [Hallberg2004]
  35. Leitner C, Volc J, and Haltrich D. (2001). Purification and characterization of pyranose oxidase from the white rot fungus Trametes multicolor. Appl Environ Microbiol. 2001;67(8):3636-44. DOI:10.1128/AEM.67.8.3636-3644.2001 | PubMed ID:11472941 [Leitner2001]
  36. Sucharitakul J, Prongjit M, Haltrich D, and Chaiyen P. (2008). Detection of a C4a-hydroperoxyflavin intermediate in the reaction of a flavoprotein oxidase. Biochemistry. 2008;47(33):8485-90. DOI:10.1021/bi801039d | PubMed ID:18652479 [Sucharitakul2008]
  37. Ma J, Zhang K, Huang M, Hector SB, Liu B, Tong C, Liu Q, Zeng J, Gao Y, Xu T, Liu Y, Liu X, and Zhu Y. (2016). Involvement of Fenton chemistry in rice straw degradation by the lignocellulolytic bacterium Pantoea ananatis Sd-1. Biotechnol Biofuels. 2016;9:211. DOI:10.1186/s13068-016-0623-x | PubMed ID:27761153 [Ma2016]

All Medline abstracts: PubMed