Cellobiose dehydrogenase (acceptor)

Last updated
cellobiose dehydrogenase (acceptor)
Identifiers
EC no. 1.1.99.18
CAS no. 54576-85-1
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

In enzymology, a cellobiose dehydrogenase (acceptor) (EC 1.1.99.18) is an enzyme that catalyzes the chemical reaction

Contents

cellobiose + acceptor cellobiono-1,5-lactone + reduced acceptor

Thus, the two substrates of this enzyme are cellobiose and acceptor, whereas its two products are cellobiono-1,5-lactone and reduced acceptor.

This enzyme belongs to the family of oxidoreductases, to be specific those acting on the CH-OH group of donor with other acceptors. The systematic name of this enzyme class is cellobiose:acceptor 1-oxidoreductase. Other names in common use include cellobiose dehydrogenase, cellobiose oxidoreductase, Phanerochaete chrysosporium cellobiose oxidoreductase, CBOR, cellobiose oxidase, cellobiose:oxygen 1-oxidoreductase, CDH, and cellobiose:(acceptor) 1-oxidoreductase. It employs sometimes one cofactor, FAD, but in most cases both a heme and a FAD located in separate domains. It makes the enzyme to one of the more complex extracellular oxidoreductases. It is produced by wood degrading organisms. [1]

Structural studies

To date, structures of the separated dehydrogenase (DH) and cytochrome (CYT) domains were reported (PDB accession codes 1NAA [2] and 1PL3 [3] ). In 2015, full-length structures of the enzyme were resolved for CDH from Crassicarpon hotsonii syn. Myriococcum thermophilum (ChCDH, PDB accession code 4QI6) and CDH from Neurospora crassa (NcCDH, PDB accession code 4QI7). [4]

The mobility of the CYT domain prevented for a long time the crystallization and X-ray structure elucidation of CDH. However, the crystal structures of the individual CYT [5] and DH [6] domains already showed structural complementarity and the position of the domain interface and provided a structural basis for the interdomain electron transfer observed in biochemical and bioelectrochemical studies. [7]

The crystal structure of the Phanerochaete chrysosporium DH domain was described by Hallberg et al. as peanut-shaped with dimensions of ~72 x 57 x 45 Å. The observed GMC-oxidoreductase protein fold features an FAD-binding and a substrate-binding subdomain. The FAD-binding subdomain primarily consists of the Rossmann fold, which is typical for NAD(P)- and FAD-dependent enzymes whereas the substrate-binding subdomain consists of a central twisted, seven-stranded β-sheet with three α-helices on one side of the sheet and the active-site on the other side. The substrate-binding subdomain hosts the active-site that consists of a substrate binding-site (B-site) and the catalytic-site (C-site). The B-site holds the non-reducing end of cellobiose in position whereas at the C-site oxidizes the reducing end of cellobiose.

The CYT domain consist of an ellipsoidal antiparallel β-sandwich with dimensions of 30 x 36 x 47 Å and a topology resembling the variable heavy chain of the antibody Fab fragment with a five-stranded inner β-sheet and a six-stranded outer β-sheet . This protein fold was a new fold observed for a cytochrome. Both individual crystal structures provided mechanistic insights. The DH domain was soaked with the inhibitor cellobionolactam which resolved the binding position of the substrate and gave a mechanistic explanation for the reductive half-reaction [52]. The CYT domain structure revealed an unusual axial haem b iron coordination by a His and a Met residue, which is the basis for intramolecular electron transfer between the two domains. [8] A mutational study on the iron-coordinating haem ligands performed by Rotseart et al. [9] showed that either the replacement of Met by His, or the switching of both ligands resulted in an IET inactive CYT domain.

In 2015, the elucidation of the full-length structures of CDH from Crassicarpon hotsonii (syn. Myriococcum thermophilum) (PDB ID: 4QI6) and Neurospora crassa (PDB ID: 4QI7) increased the understanding of the domain mobility for the role of CYT as built-in redox mediator. Tan and Kracher et al. [10] showed that CDH exists in a closed-state and an open-state conformation. In the closed-state the haem b accepts an electron from the FAD after substrate oxidation and it donates the electron to an external electron acceptor in the open-state. It was found that IET does not depend on the prominent Trp residue positioned directly between the FAD and the haem cofactor, but on a surface exposed Arg in the substrate channel, which stabilises the interaction with CYT through a haem propionate group. The heme b propionate-D is folded away, and a hydrogen bond to Tyr99 in the CYT domain prevents it from interacting directly with the DH active site. The closest edge-to-edge distance between heme b and FAD is 9 Å, which is well within the 14-Å limit for efficient electron transfer. [11] The mobile CYT domain of CDH is a unique feature among GMC-oxidoreductases and acquired at a later stage of evolution. [12] The CYT domain mobility is regulated by a flexible peptide linker of varying length and composition (e.g., N. crassa CDH IIA = 17 amino acids, N. crassa CDH IIB = 33 amino acids), which connects both domains. The movement of CYT has been observed by AFM [13] and SAXS [14] studies.

Phylogenetic classification

The CDH sequences evolved into four phylogenetic branches: [15] Class I CDH sequences are found exclusively in Basidiomycota whereas CDH sequences from Class II, Class III, and Class IV are found only in Ascomycota. Class I CDHs have a strong affinity for cellulose and bind presumably via a cellulose binding-site on the surface of the DH domain [16] that is different from the active-site. This cellulose binding-site is not present in other CDH classes. The binding to cellulose in Class II CDHs depends on the presence of a C-terminal family 1 carbohydrate binding module (CBM1). Some Class II CDH feature a C-terminal CBM1 and are classified as Class IIA CDHs, whereas Class IIB CDHs have no CBM1 and do not bind to cellulose. Other differences between Class I and Class II CDHs are in their substrate specificity and the pH optimum of the IET. Little can be said about Class III and Class IV CDHs, which have not yet been expressed and characterised. From sequence alignments a different active-site geometry can be deduced, but the substrate is unknown. Class IV CDHs are evolutionary furthest related to the other CDH classes and do not feature an electron transferring CYT domain, which suggests a different physiological role and the loss of its ability for DET.

Catalysis

CDH catalyses the 2e-/2H+ oxidation of the anomeric carbon atom (C1) of the disaccharide cellobiose to the cellobiono-δ-lactone hydrolyses further to cellobionic acid in water. Besides the natural substrate cellobiose and the cellodextrins cellotriose, cellotetraose, cellopentaose, and cellohexaose also the cellulose building unit glucose, as well as the hemicellulose building monomers or breakdown products galactose, mannose, mannobiose, mannopentaose, xylose, xylobiose and xylotriose, or starch derived maltose, maltotriose and maltotetraose have been reported as substrates of greatly varying catalytic efficiency for CDH. [17] A very good substrate not related to plant polysaccharides is lactose, because of its structural similarity to cellobiose. Cellobiose, cello-oligosaccharides, and lactose are the substrates for which CDH exhibits the highest catalytic efficiency, whereas monosaccharides are bad substrates with a very low affinity (high KM) to only the catalytic C-site of CDH. The extreme discrepancy of the catalytic efficiencies of P. chrysosporium CDH for cellobiose over glucose (87500 : 1) is being connected to the physiological role of the white-rot basidiomycete enzyme. [18] In ascomycete CDHs the discrimination of glucose is less strict [19] and a wider spectrum of mono- and oligosaccharides are converted. Of course, the affinity of CDH to di- and oligosaccharide substrates is much higher than for monosaccharides. However, Class II CDHs have KM-values for glucose of 10–100 mM, which is in the range of FAD-dependent glucose oxidase, also a GMC-oxidoreductase (5–100 mM). The lower catalytic rate of Class I and Class II CDHs (below 50 s−1) compared to glucose oxidase (300–2000 s−1) is probably an evolutionary adaption to the acquired CYT domain to optimize IET and prevent futile reactions of the substrate in an idle active-site.

CDH as a bioelectrocatalyst

CDH exhibits various properties that makes it a suitable electrocatalyst for biosensors or biofuel cells. [20] Unlike any other carbohydrate converting GMC-oxidoreductase it can be contacted via mediated electron transfer (MET) or direct electron transfer (DET). CDH-based 2nd and 3rd generation biosensors for the quantitation of commercially or medically relevant molecules such as glucose or lactose have been developed. The advantage 2nd generation biosensors is their high sensitivity, the advantage of 3rd generation biosensors is the avoidance of redox mediators in implantable blood glucose monitoring systems. CDH-based 3rd generation biosensors have a robust performance because of the good thermal and turnover stability of CDH and the mobile CYT domain, which can interact with many electrode materials and surface modifications and deliver reasonably high current densities.

CDH-based biosensors

CDH has been used as biosensor element for the detection of a number of different analytes ever since 1992, when Elmgren et al. reported on a new biosensor sensitive towards cello-oligosaccharides with degrees of polymerization of 2 to 6, lactose and maltose. [21] Research on CDH employed as biosensor since then until 2013 has been reviewed, [22] providing lists of CDH-based biosensors in both DET and MET mode for phenols (catechol, dopamine, noradrenaline, hydroquinone, aminophenol) and carbohydrates (cellobiose, lactose, maltose, glucose).

Both Class I and Class II CDHs have been employed as biocatalysts for biosensors. In accordance with their preferred substrates, Class I CDHs were mainly applied for the detection and quantitation of phenols, cellobiose, and lactose, whereas Class II CDHs have been applied for the detection of cellobiose, lactose, maltose and glucose. Among Class I CDHs, Phanerochaete chrysosporium CDH is the most comprehensively studied enzyme was employed for the detection of various phenol derivates, cellobiose and lactose. Phanerochaete sordida and Trametes villosa CDH have been used for the detection of lactose and Sclerotium rolfsii CDH for the detection of dopamine.

Among Class II CDHs, Crassicarpon hotsonii and Crassicarpon thermophilum CDH were employed as recognition element for cellobiose, lactose, glucose and maltose and Humicola insolens CDH for glucose and maltose. Detection of maltose was in all cases only reported to show that maltose is not interfering with glucose detection, which is important for blood glucose testing. [23]

In 2017, DirectSens GmbH launched the first lactose biosensor based on a CDH for very low lactose concentrations in lactose reduced milk products. LactoSens is the only third-generation biosensor on the market and distributed globally to dairy companies.

Related Research Articles

<span class="mw-page-title-main">Disaccharide</span> Complex sugar

A disaccharide is the sugar formed when two monosaccharides are joined by glycosidic linkage. Like monosaccharides, disaccharides are simple sugars soluble in water. Three common examples are sucrose, lactose, and maltose.

A dehydrogenase is an enzyme belonging to the group of oxidoreductases that oxidizes a substrate by reducing an electron acceptor, usually NAD+/NADP+ or a flavin coenzyme such as FAD or FMN. Like all catalysts, they catalyze reverse as well as forward reactions, and in some cases this has physiological significance: for example, alcohol dehydrogenase catalyzes the oxidation of ethanol to acetaldehyde in animals, but in yeast it catalyzes the production of ethanol from acetaldehyde.

<span class="mw-page-title-main">Rossmann fold</span>

The Rossmann fold is a tertiary fold found in proteins that bind nucleotides, such as enzyme cofactors FAD, NAD+, and NADP+. This fold is composed of alternating beta strands and alpha helical segments where the beta strands are hydrogen bonded to each other forming an extended beta sheet and the alpha helices surround both faces of the sheet to produce a three-layered sandwich. The classical Rossmann fold contains six beta strands whereas Rossmann-like folds, sometimes referred to as Rossmannoid folds, contain only five strands. The initial beta-alpha-beta (bab) fold is the most conserved segment of the Rossmann fold. The motif is named after Michael Rossmann who first noticed this structural motif in the enzyme lactate dehydrogenase in 1970 and who later observed that this was a frequently occurring motif in nucleotide binding proteins.

PEP group translocation, also known as the phosphotransferase system or PTS, is a distinct method used by bacteria for sugar uptake where the source of energy is from phosphoenolpyruvate (PEP). It is known to be a multicomponent system that always involves enzymes of the plasma membrane and those in the cytoplasm.

<span class="mw-page-title-main">Flavin adenine dinucleotide</span> Redox-active coenzyme

In biochemistry, flavin adenine dinucleotide (FAD) is a redox-active coenzyme associated with various proteins, which is involved with several enzymatic reactions in metabolism. A flavoprotein is a protein that contains a flavin group, which may be in the form of FAD or flavin mononucleotide (FMN). Many flavoproteins are known: components of the succinate dehydrogenase complex, α

Acyl-CoA dehydrogenases (ACADs) are a class of enzymes that function to catalyze the initial step in each cycle of fatty acid β-oxidation in the mitochondria of cells. Their action results in the introduction of a trans double-bond between C2 (α) and C3 (β) of the acyl-CoA thioester substrate. Flavin adenine dinucleotide (FAD) is a required co-factor in addition to the presence of an active site glutamate in order for the enzyme to function.

<span class="mw-page-title-main">Electron-transferring-flavoprotein dehydrogenase</span> Protein family

Electron-transferring-flavoprotein dehydrogenase is an enzyme that transfers electrons from electron-transferring flavoprotein in the mitochondrial matrix, to the ubiquinone pool in the inner mitochondrial membrane. It is part of the electron transport chain. The enzyme is found in both prokaryotes and eukaryotes and contains a flavin and FE-S cluster. In humans, it is encoded by the ETFDH gene. Deficiency in ETF dehydrogenase causes the human genetic disease multiple acyl-CoA dehydrogenase deficiency.

In enzymology, sarcosine dehydrogenase (EC 1.5.8.3) is a mitochondrial enzyme that catalyzes the chemical reaction N-demethylation of sarcosine to give glycine. This enzyme belongs to the family of oxidoreductases, specifically those acting on the CH-NH group of donor with other acceptors. The systematic name of this enzyme class is sarcosine:acceptor oxidoreductase (demethylating). Other names in common use include sarcosine N-demethylase, monomethylglycine dehydrogenase, and sarcosine:(acceptor) oxidoreductase (demethylating). Sarcosine dehydrogenase is closely related to dimethylglycine dehydrogenase, which catalyzes the demethylation reaction of dimethylglycine to sarcosine. Both sarcosine dehydrogenase and dimethylglycine dehydrogenase use FAD as a cofactor. Sarcosine dehydrogenase is linked by electron-transferring flavoprotein (ETF) to the respiratory redox chain. The general chemical reaction catalyzed by sarcosine dehydrogenase is:

<span class="mw-page-title-main">Glycerol dehydrogenase</span>

Glycerol dehydrogenase (EC 1.1.1.6, also known as NAD+-linked glycerol dehydrogenase, glycerol: NAD+ 2-oxidoreductase, GDH, GlDH, GlyDH) is an enzyme in the oxidoreductase family that utilizes the NAD+ to catalyze the oxidation of glycerol to form glycerone (dihydroxyacetone).

<span class="mw-page-title-main">Alcohol oxidase</span>

In enzymology, an alcohol oxidase (EC 1.1.3.13) is an enzyme that catalyzes the chemical reaction

In enzymology, a glucose 1-dehydrogenase is an enzyme that catalyzes the chemical reaction

In enzymology, a quinoprotein glucose dehydrogenase is an enzyme that catalyzes the chemical reaction

In enzymology, a 4-cresol dehydrogenase (hydroxylating) (EC 1.17.99.1) is an enzyme that catalyzes the chemical reaction

<span class="mw-page-title-main">Proline dehydrogenase</span>

In enzymology, proline dehydrogenase (PRODH) (EC 1.5.5.2, formerly EC 1.5.99.8) is an enzyme of the oxidoreductase family, active in the oxidation of L-proline to (S)-1-pyrroline-5-carboxylate during proline catabolism. The end product of this reaction is then further oxidized in a (S)-1-pyrroline-5-carboxylate dehydrogenase (P5CDH)-dependent reaction of the proline metabolism, or spent to produce ornithine, a crucial metabolite of ornithine and arginine metabolism. The systematic name of this enzyme class is L-proline:quinone oxidoreductase. Other names in common use include L-proline dehydrogenase, L-proline oxidase,and L-proline:(acceptor) oxidoreductase. It employs one cofactor, FAD, which requires riboflavin (vitamin B2).

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

Cytochromes c cytochromes, or heme-containing proteins, that have heme C covalently attached to the peptide backbone via one or two thioether bonds. These bonds are in most cases part of a specific Cys-X-X-Cys-His (CXXCH) binding motif, where X denotes a miscellaneous amino acid. Two thioether bonds of cysteine residues bind to the vinyl sidechains of heme, and the histidine residue coordinates one axial binding site of the heme iron. Less common binding motifs can include a single thioether linkage, a lysine or a methionine instead of the axial histidine or a CXnCH binding motif with n>2. The second axial site of the iron can be coordinated by amino acids of the protein, substrate molecules or water. Cytochromes c possess a wide range of properties and function as electron transfer proteins or catalyse chemical reactions involving redox processes. A prominent member of this family is mitochondrial cytochrome c.

<span class="mw-page-title-main">Flavocytochrome c sulfide dehydrogenase</span>

Flavocytochrome c sulfide dehydrogenase, also known as Sulfide-cytochrome-c reductase (flavocytochrome c) (EC 1.8.2.3), is an enzyme with systematic name hydrogen-sulfide:flavocytochrome c oxidoreductase. It is found in sulfur-oxidising bacteria such as the purple phototrophic bacteria Allochromatium vinosum. This enzyme catalyses the following chemical reaction:

<span class="mw-page-title-main">Glucose-methanol-choline oxidoreductase family</span>

In molecular biology, the glucose-methanol-choline oxidoreductase family is a family of enzymes with oxidoreductase activity.

<span class="mw-page-title-main">NADH:ubiquinone reductase (non-electrogenic)</span> Class of enzymes

NADH:ubiquinone reductase (non-electrogenic) (EC 1.6.5.9, NDH-2, ubiquinone reductase, coenzyme Q reductase, dihydronicotinamide adenine dinucleotide-coenzyme Q reductase, DPNH-coenzyme Q reductase, DPNH-ubiquinone reductase, NADH-coenzyme Q oxidoreductase, NADH-coenzyme Q reductase, NADH-CoQ oxidoreductase, NADH-CoQ reductase) is an enzyme with systematic name NADH:ubiquinone oxidoreductase. This enzyme catalyses the following chemical reaction:

YedZ of E. coli has been examined topologically and has 6 transmembrane segments (TMSs) with both the N- and C-termini localized to the cytoplasm. von Rozycki et al. 2004 identified homologues of YedZ in bacteria and animals. YedZ homologues exhibit conserved histidyl residues in their transmembrane domains that may function in heme binding. Some of the homologues encoded in the genomes of various bacteria have YedZ domains fused to transport, electron transfer and biogenesis proteins. One of the animal homologues is the 6 TMS epithelial plasma membrane antigen of the prostate (STAMP1) that is over-expressed in prostate cancer. Some animal homologues have YedZ domains fused C-terminal to homologues of NADP oxidoreductases.

<span class="mw-page-title-main">NDH-2</span>

NDH-2, also known as type II NADH:quinone oxidoreductase or alternative NADH dehydrogenase, is an enzyme which catalyzes the electron transfer from NADH to a quinone, being part of the electron transport chain. NDH-2 are peripheral membrane protein, functioning as dimers in vivo, with approximately 45 KDa per subunit and a single FAD as their cofactor.

References

  1. Westermark, Ulla; Eriksson, Karl-Erik; Daasvatn, Kari; Liaaen-Jensen, Synnøve; Enzell, Curt R.; Mannervik, Bengt (1974). "Cellobiose:Quinone Oxidoreductase, a New Wood-degrading Enzyme from White-rot Fungi". Acta Chemica Scandinavica. 28b: 209–214. doi: 10.3891/acta.chem.scand.28b-0209 .
  2. Martin Hallberg, B; Henriksson, Gunnar; Pettersson, Göran; Divne, Christina (2002-01-18). "Crystal structure of the flavoprotein domain of the extracellular flavocytochrome cellobiose dehydrogenase1". Journal of Molecular Biology. 315 (3): 421–434. doi:10.1006/jmbi.2001.5246. PMID   11786022.
  3. Rotsaert, Frederik A. J.; Hallberg, B. Martin; Vries, Simon de; Moenne-Loccoz, Pierre; Divne, Christina; Renganathan, V.; Gold, Michael H. (2003-08-29). "Biophysical and Structural Analysis of a Novel Heme b Iron Ligation in the Flavocytochrome Cellobiose Dehydrogenase". Journal of Biological Chemistry. 278 (35): 33224–33231. doi: 10.1074/jbc.M302653200 . ISSN   0021-9258. PMID   12796496.
  4. Tan TC, Kracher D, Gandini R, Sygmund C, Kittl R, Haltrich D, Hällberg BM, Ludwig R, Divne C (July 2015). "Structural basis for cellobiose dehydrogenase action during oxidative cellulose degradation". Nat. Commun. 6: 7542. Bibcode:2015NatCo...6.7542T. doi:10.1038/ncomms8542. PMC   4507011 . PMID   26151670.
  5. Hallberg, B. Martin; Bergfors, Terese; Bäckbro, Kristina; Pettersson, Göran; Henriksson, Gunnar; Divne, Christina (2000-01-01). "A new scaffold for binding haem in the cytochrome domain of the extracellular flavocytochrome cellobiose dehydrogenase". Structure. 8 (1): 79–88. doi: 10.1016/S0969-2126(00)00082-4 . ISSN   0969-2126. PMID   10673428.
  6. Hallberg, B. M.; Henriksson, Gunnar; Pettersson, Göran; Divne, Christina (2002-01-18). "Crystal structure of the flavoprotein domain of the extracellular flavocytochrome cellobiose dehydrogenase". Journal of Molecular Biology. 315 (3): 421–434. doi:10.1006/jmbi.2001.5246. ISSN   0022-2836. PMID   11786022.
  7. Zamocky, Marcel; Ludwig, R.; Peterbauer, C.; Hallberg, B. M.; Divne, C.; Nicholls, P.; Haltrich, D. (June 2006). "Cellobiose dehydrogenase--a flavocytochrome from wood-degrading, phytopathogenic and saprotropic fungi". Current Protein & Peptide Science. 7 (3): 255–280. doi:10.2174/138920306777452367. ISSN   1389-2037. PMID   16787264.
  8. Hallberg, B. Martin; Henriksson, Gunnar; Pettersson, Goran; Vasella, Andrea; Divne, Christina (2003-02-28). "Mechanism of the reductive half-reaction in cellobiose dehydrogenase". The Journal of Biological Chemistry. 278 (9): 7160–7166. doi: 10.1074/jbc.M210961200 . ISSN   0021-9258. PMID   12493734.
  9. Rotsaert, Frederik A. J.; Hallberg, B. Martin; de Vries, Simon; Moenne-Loccoz, Pierre; Divne, Christina; Renganathan, V.; Gold, Michael H. (2003-08-29). "Biophysical and structural analysis of a novel heme B iron ligation in the flavocytochrome cellobiose dehydrogenase". The Journal of Biological Chemistry. 278 (35): 33224–33231. doi: 10.1074/jbc.M302653200 . ISSN   0021-9258. PMID   12796496.
  10. Tan, Tien-Chye; Kracher, Daniel; Gandini, Rosaria; Sygmund, Christoph; Kittl, Roman; Haltrich, Dietmar; Hällberg, B. Martin; Ludwig, Roland; Divne, Christina (2015-07-07). "Structural basis for cellobiose dehydrogenase action during oxidative cellulose degradation". Nature Communications. 6: 7542. Bibcode:2015NatCo...6.7542T. doi:10.1038/ncomms8542. ISSN   2041-1723. PMC   4507011 . PMID   26151670.
  11. Page, C. C.; Moser, C. C.; Chen, X.; Dutton, P. L. (1999-11-04). "Natural engineering principles of electron tunnelling in biological oxidation-reduction". Nature. 402 (6757): 47–52. Bibcode:1999Natur.402...47P. doi:10.1038/46972. ISSN   0028-0836. PMID   10573417. S2CID   4431405.
  12. Zámocký, Marcel; Hallberg, Martin; Ludwig, Roland; Divne, Christina; Haltrich, Dietmar (2004-08-18). "Ancestral gene fusion in cellobiose dehydrogenases reflects a specific evolution of GMC oxidoreductases in fungi". Gene. 338 (1): 1–14. doi:10.1016/j.gene.2004.04.025. ISSN   0378-1119. PMID   15302401.
  13. Harada, Hirofumi; Onoda, Akira; Uchihashi, Takayuki; Watanabe, Hiroki; Sunagawa, Naoki; Samejima, Masahiro; Igarashi, Kiyohiko; Hayashi, Takashi (2017-09-01). "Interdomain flip-flop motion visualized in flavocytochrome cellobiose dehydrogenase using high-speed atomic force microscopy during catalysis". Chemical Science. 8 (9): 6561–6565. doi:10.1039/c7sc01672g. ISSN   2041-6520. PMC   5627353 . PMID   28989682.
  14. Bodenheimer, Annette M.; O'Dell, William B.; Oliver, Ryan C.; Qian, Shuo; Stanley, Christopher B.; Meilleur, Flora (April 2018). "Structural investigation of cellobiose dehydrogenase IIA: Insights from small angle scattering into intra- and intermolecular electron transfer mechanisms". Biochimica et Biophysica Acta (BBA) - General Subjects. 1862 (4): 1031–1039. doi: 10.1016/j.bbagen.2018.01.016 . ISSN   0304-4165. OSTI   1424446. PMID   29374564.
  15. Sützl, Leander; Foley, Gabriel; Gillam, Elizabeth M. J.; Bodén, Mikael; Haltrich, Dietmar (2019). "The GMC superfamily of oxidoreductases revisited: analysis and evolution of fungal GMC oxidoreductases". Biotechnology for Biofuels. 12: 118. doi: 10.1186/s13068-019-1457-0 . ISSN   1754-6834. PMC   6509819 . PMID   31168323.
  16. Hallberg, B. M.; Henriksson, Gunnar; Pettersson, Göran; Divne, Christina (2002-01-18). "Crystal structure of the flavoprotein domain of the extracellular flavocytochrome cellobiose dehydrogenase". Journal of Molecular Biology. 315 (3): 421–434. doi:10.1006/jmbi.2001.5246. ISSN   0022-2836. PMID   11786022.
  17. Scheiblbrandner, Stefan; Ludwig, Roland (February 2020). "Cellobiose dehydrogenase: Bioelectrochemical insights and applications". Bioelectrochemistry. 131: 107345. doi: 10.1016/j.bioelechem.2019.107345 . ISSN   1878-562X. PMID   31494387. S2CID   201215946.
  18. Henriksson, G.; Sild, V.; Szabó, I. J.; Pettersson, G.; Johansson, G. (1998-03-03). "Substrate specificity of cellobiose dehydrogenase from Phanerochaete chrysosporium". Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 1383 (1): 48–54. doi:10.1016/s0167-4838(97)00180-5. ISSN   0006-3002. PMID   9546045.
  19. Harreither, Wolfgang; Sygmund, Christoph; Augustin, Manfred; Narciso, Melanie; Rabinovich, Mikhail L.; Gorton, Lo; Haltrich, Dietmar; Ludwig, Roland (March 2011). "Catalytic properties and classification of cellobiose dehydrogenases from ascomycetes". Applied and Environmental Microbiology. 77 (5): 1804–1815. Bibcode:2011ApEnM..77.1804H. doi:10.1128/AEM.02052-10. ISSN   1098-5336. PMC   3067291 . PMID   21216904.
  20. Scheiblbrandner, Stefan; Ludwig, Roland (2020-02-01). "Cellobiose dehydrogenase: Bioelectrochemical insights and applications". Bioelectrochemistry. 131: 107345. doi: 10.1016/j.bioelechem.2019.107345 . ISSN   1567-5394. PMID   31494387. S2CID   201215946.
  21. Elmgren, Maja; Lindquist, Sten-Eric; Henriksson, Gunnar (1992-12-10). "Cellobiose oxidase crosslinked in a redox polymer matrix at an electrode surface—a new biosensor". Journal of Electroanalytical Chemistry. An International Journal Devoted to all Aspects of Electrode Kinetics, Interfacial Structure, Properties of Electrolytes, Colloid and Biological Electrochemistry. 341 (1): 257–273. doi:10.1016/0022-0728(92)80487-O. ISSN   1572-6657.
  22. Ludwig, Roland; Harreither, Wolfgang; Tasca, Federico; Gorton, Lo (2010-09-10). "Cellobiose dehydrogenase: a versatile catalyst for electrochemical applications". ChemPhysChem. 11 (13): 2674–2697. doi:10.1002/cphc.201000216. ISSN   1439-7641. PMID   20661990.
  23. Janssen, W.; Harff, G.; Caers, M.; Schellekens, A. (November 1998). "Positive interference of icodextrin metabolites in some enzymatic glucose methods". Clinical Chemistry. 44 (11): 2379–2380. doi: 10.1093/clinchem/44.11.2379 . ISSN   0009-9147. PMID   9799776.