リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

大学・研究所にある論文を検索できる 「Structural insight into the mechanism of angular dioxygenation in carbazole 1,9a-dioxygenase」の論文概要。リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

コピーが完了しました

URLをコピーしました

論文の公開元へ論文の公開元へ
書き出し

Structural insight into the mechanism of angular dioxygenation in carbazole 1,9a-dioxygenase

王, 翌霞 東京大学 DOI:10.15083/0002002073

2021.10.04

概要

Introduction
Carbazole (CAR) is a recalcitrant N-heterocyclic aromatic compound with mutagenicity and toxicity that is predominant in coal tar creosote, the release of which has caused environmental concerns. To achieve bioremediation of CAR in the environment, various CAR-degrading bacteria that use CAR as the sole source of nitrogen, carbon and energy, have been isolated from diverse niches and the corresponding catabolic gene (car) clusters from these bacteria have also been identified and characterized. It has been reported that angular dioxygenation of CAR, adding two hydroxyl groups with cis-confirmation to the angular position (C9a) carbon bound to the imino nitrogen and its adjacent C1 carbon, is most important by destroying the planar structure from which the toxicity derives, resulting in complete mineralization of CAR. Carbazole 1,9a-dioxygenase (CARDO), a member of three-component Rieske non-heme iron oxygenase (RO) consisting of ferredoxin reductase (Red, 37 kDa), ferredoxin (Fd, 13 kDa) and homo-trimeric terminal oxygenase (Oxy, 132 kDa), has been reported to initiate the degradation of CAR by angular dioxygenation to yield unstable cis-dihydrodiol, which subsequently converts to 2'-aminobiphenyl-2,3-diol (ABP-diol) spontaneously. To illuminate the mechanism of angular dioxygenation in CARDO, catalytic cycle of CARDO has been proposed on the basis of various well-determined structures of CARDO and other ROs (Fig. 1). As is shown in Fig. 1, Oxyox/red (1) in resting state contains oxidized Rieske cluster and reduced mononuclear iron in each subunit. Then reduced Fd (Fdred) binds to the resting Oxy to form the transient complex of Oxyox/red:Fdred. One electron is transferred from Fdred to Oxyox/red, resulting in the changes of redox state of both Oxy and Fd to for the complex of Oxyred/red:Fdox (2), followed by the release of oxidized Fd (Fdox). The Oxyred/red (3) then are able to bind with substrate and oxygen to initiate the dioxygenation of substrate. According to the possible binding sequence of substrate and oxygen, two pathways have been proposed: in pathway 1, aromatic substrate (eg. CAR) is available to the non-heme iron prior to oxygen molecule to form the binary Oxyred/red:CAR (4a); in pathway 2, oxygen molecule binds to the active site ahead of substrate, resulting in the formation of Oxyox/ox:O2 (4b). When both substrate and O2 are available to the active site, substrate transformation starts (Oxyox/ox:CAR:O2, (5)). The finally generated cis-dihydrodiol product binds at the active site with both oxygen atoms coordinated to the ferric iron (Oxyox/ox:ABP-diol, (6)). The active site of Oxyox/ox:ABP-diol is reduced by receiving the second electron from Fdred to form the complex of Oxyox/red:Fdox:ABP-diol binding with protonated ABP-diol (7). Two possible pathways have also been proposed based on the presumed release order of Fdox and product: in pathway 3, Fdox dissociates from the complex before the release of product (Oxyox/red:ABP-diol, (8a)); in pathway 4, ABP-diol escapes from the active site before the dissociation of Oxyox/red:Fdox complex (8b). Finally the Oxy returns to the resting state and initiates next turnover. To date, the structures of Oxyox/red (1), Oxyred/red:Fdox (2), Oxyred/red (3), Oxyox/red:Fdox (8b), as well as Fd in resting and reduced states (Fdox and Fdred in Fig. 1), have been solved. The current study continues on determining the structures of CARDO in substrate binding, O2 activation, product formation and release to illustrate the mechanism of angular dioxygenation in CARDO from structural insight.

Chapter 2. Substrate binding and product formation in Oxy of CARDO
In pathway 1, substrate binding was achieved by soaking crystals of previously obtained Oxyred/red (3) in dimethyl sulfoxide (DMSO) solution of CAR under anaerobic condition. The crystal structure of Oxyred/red:CAR (4a) was determined at 2.2 Å resolution in Photon Factory (Tsukuba, Japan) with the α3 doughnut-like quaternary structure. The structure of Oxyred/red:CAR shows CAR binds at the active site in such a conformation that the carbon atoms at the angular (C9a) and its adjacent (C1) are exposed to attack by O2, and is stabilized by hydrogen bonding of the imino nitrogen with the carbonyl oxygen of G178. Binding of planar CAR at the active site causes the large movement of residues L202-T214 and D229-V238 towards the entrance of the catalytic pocket, especially the significant shifts of residues F204 and I231, resulting in the closure of the catalytic pocket to trap CAR inside.
 In pathway 2, the crystals of Oxyred/red were moved out of the anaerobic chamber and exposed to pure O2 for 5 min to obtain the O2-bound Oxy (4b). The structure, determined at 2.2 Å resolution in Photon Factory, reveals that in absence of substrate, dioxygen binds skewed side-on to the active site, causing no significant conformational changes in the overall structure, compared with Oxyred/red, except for the flip of some side chains such as those of E233 and H234. Still, the catalytic pocket keeps open widely.
 When both substrate and O2 are available to the active site, substrate transformation starts. To obtain the substrate plus O2-bound structure, crystals of Oxyred/red:CAR were moved out of the anaerobic chamber and exposed to pure oxygen for 5 min before freezing. The structure determined at a resolution of 2.2 Å in Spring-8 shows different geometries in active sites A, B and C. Subunit A shows a positive density in the active site, which could not be modeled as planar CAR and individual dioxygen. Instead, the continuous electron density from the active site iron to the slightly puckered CAR suggests the presence of a covalent link between the iron and substrate. This suggests a presumed tricyclic alkylperoxo intermediate binds at the active site A (Oxyox/ox:CAR:O2; (5)), which was generated during the hydroxylation reaction. The tricyclic moiety of alkylperoxo intermediate is still stabilized by G178 and the peroxo moiety is stabilized with the 2-His-1-carboxylate facial triad motif. Furthermore, the formation of alkylperoxo intermediate initiates the opening of the previously well-closed catalytic pocket caused the binding of substrate. The continuous and distorted ligand density in the active sites B and C and the disappearance of electron density of former 5-membered CAR ring indicates the cleavage of O-O bond of the peroxo moiety and the ring fission between C9a and N9 to form ABP-diol (Oxyox/ox:ABP-diol, (6)). The product remains bound at the active site with both oxygen atoms coordinated to the iron in an average distance of 2.2 Å. Generation of ABP-diol has resulted in wide re-opening of the catalytic pocket by movements of residues L202-T214 and D229-V238 (especially F204 and I231) back to the initial positions, through which the product is ready to be released from the active site.

Chapter 3. Product release from the active site of Oxy in CARDO
To understand the mechanism of product release in CARDO, commercially available biphenyl-2,2',3-triol (BP-triol) and biphenyl-2,3-diol (BP-diol) were used as products instead of ABP-diol in the soaking experiments to obtain the product-bound structures. The previously obtained crystals of Oxyox/red (1) and Oxyox/red:Fdox complex (8b) were soaked in the DMSO solution of BP-triol or BP-diol for 5~10 min before cryo-cooling. The X-ray diffraction data were collected in Photon Factory.
 The determined structure of BP-diol bound Oxyox/red:Fdox complex (7) at a resolution of 2.0 Å, consists of one molecule of Oxy and three Fd molecules, and the asymmetric unit of the crystal contains one Oxyox/red:Fdox molecule. BP-diol with low occupancy binds at active site A with both O atoms coordinated to Fe(II), which is similar with the binding mode of ABP-diol to Fe(III) mentioned above. While the two hydroxyl groups of BP-diol in active sites B and C have rotated away from Fe(II), resulting in a water molecule lying between the Fe(II) and proximal ring of BP-diol, significantly different with that of ABP-diol at the non-heme Fe(III). Compared with Oxyox/ox:ABP-diol, some conformational changes have been observed at the boundary of Oxy and Fd. For example, side chains of R11, K13 and R118 of Oxy have flipped to form interactions with residues in Fd to stabilize the complex of Oxy and Fd.
 In pathway 3 for the product release, Fdox dissociates from Oxyox/red:product complex followed by the release of product. To get the product-bound Oxyox/red in the absence of Fdox (8a), crystals of Oxyox/red (1) were soaked with BP-triol or BP-diol solution for 5-10 min. The determined structure of Oxyox/red:BP-triol at a resolution of 2.3 Å reveals that this product binds at the active site with two hydroxyl groups coordinated to the iron, while the third hydroxyl group in the distal ring is stabilized by hydrogen bonds with G178 (3.4 Å) and two water molecules. While the structure of Oxyox/red:BP-diol determined at a resolution of 1.9 Å shows that this product analog binds at the active site through a water molecule at 1.8 Å from the Fe(II), which is similar with the binding mode of BP-diol in active sites B and C of Oxyox/red:Fdox:BP-diol. No obvious conformational changes are caused by the binding of BP-triol and BP-diol to the active site, except for the binding modes of BP-triol and BP-diol as described previously. Upon dissociation of Fdox from product-bound Oxyox/red, interactions that stabilize Oxy:Fd complex are disrupted by the flip of side chains of related residues, including R11, K13 and R118 in Oxy. Then the subsequent release of BP-triol or BP-diol has caused a shift of residues A259-D261 towards the entrance of the catalytic pocket up to 1.3 Å (N260), leading to contraction of the entrance of the catalytic pocket.
 In the pathway 4, cis-dihydrodiol product is released from the active site before the dissociation of Oxyox/red:Fdox complex. Comparison of Oxyox/red:Fdox:BP-diol and previously determined product-free Oxyox/red:Fdox (8b) shows that when the product is released from the active site, entrance of catalytic pocket is contracted caused by the movement of residues A259-D261, which is similar to the contraction caused by the product release from Oxyox/red.

Conclusions
On the basis of the series of Oxy structures bound with substrate, dioxygen, intermediate and products identified in this study, as well as those previously determined structures of Oxy and Fd in single and complex states, the mechanism of angular dioxygenation in CARDO has been elucidated (Fig. 2). (I) In the resting state, the mononuclear Fe(II) of Oxyox/red has a distorted pentahedral geometry. Fe(II) is accessible easily for the solvent and substrate through a largely open catalytic pocket. (II) Fdred transports an electron to the resting Oxy. (III) The obtained Oxyred/red shows a distorted octahedral coordination including two water molecules at the active site. The catalytic pocket opens widely. (IV) Coupling with reduced Rieske center, both substrate and O2 are possible to bind to the Fe(II). (IVa) Accessibility of planar CAR to the active site results in the closure of the catalytic pocket by shifts of residues L202-T214 and D229-V238, especially F204 and I231 to trap the substrate inside; (IVb) while dioxygen bound to the substrate-free active site in a skewed side-on fashion has barely effects on the overall structure. (V) Availability of both substrate and dioxygen to the active site results in the formation of alkyloperoxo intermediate, accompanied by the partial opening of catalytic pocket. (VI) The alkylperoxo intermediate is converted to the cis-dihydrodiol product, with two O atoms coordinated to the non-heme Fe(III), resulting in the wide re-open of the pocket. (VII) The protonated bicyclic product with high flexibility is released from the re-reduced non-heme Fe(II). Oxy returns back to the resting state to start next turnover.

参考文献

Alexander, M. (1999). Bioremediation and Biodegradation. Focus (Madison).

Arcos, J.C., and Argus, M.F. (1969). Molecular Geometry and Carcinogenic Activity of Aromatic Compounds. New Perspectives. Adv. Cancer Res. 11, 305–471.

Ashikawa, Y., Fujimoto, Z., Noguchi, H., Habe, H., Omori, T., Yamane, H., and Nojiri, H. (2005). Crystallization and preliminary X-ray diffraction analysis of the electron-transfer complex between the terminal oxygenase component and ferredoxin in the Rieske non-haem iron oxygenase system carbazole 1,9a-dioxygenase. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 61, 577–580.

Ashikawa, Y., Fujimoto, Z., Noguchi, H., Habe, H., Omori, T., Yamane, H., and Nojiri, H. (2006). Electron Transfer Complex Formation between Oxygenase and Ferredoxin Components in Rieske Nonheme Iron Oxygenase System. Structure 14, 1779–1789.

Ashikawa, Y., Fujimoto, Z., Usami, Y., Inoue, K., Noguchi, H., Yamane, H., and Nojiri, H. (2012). Structural insight into the substrate- and dioxygen-binding manner in the catalytic cycle of rieske nonheme iron oxygenase system, carbazole 1,9a-dioxygenase. BMC Struct. Biol. 12, 1.

Bassan, A., Blomberg, M.R.A., Borowski, T., and Siegbahn, P.E.M. (2004). Oxygen activation by rieske non-heme iron oxygenases, a theoretical insight. J. Phys. Chem. B 108, 13031–13041.

Batie, C.J., Ballou, D.P., and Correl, C.C. (1991). Phthalate dioxygenase reductase and related flavin-iron-sulfur containing electron transferases. In Chemistry and Biochemistry of Flavoenzymes Volume III, M. Franz, ed. (Boca Raton, FL: CRC Press), pp. 543–556.

Bertini, I., Cremonini, M.A., Ferretti, S., Lozzi, I., Luchinat, C., and Viezzoli, M.S. (1996). Arene hydroxylases: metalloenzymes catalysing dioxygenation of aromatic compounds. Coord. Chem. Rev. 151, 145–160.

Bimboim, H.C., and Doly, J. (1979). A rapid alkaline extraction procedure for screening recombinant plasmid DNA. Nucleic Acids Res.

Bingham, R.J., Findlay, J.B.C., Hsieh, S.Y., Kalverda, A.P., Kjellberg, A., Perazzolo, C., Phillips, S.E.V., Seshadri, K., Trinh, C.H., Turnbull, W.B., et al. (2004). Thermodynamics of Binding of 2-Methoxy-3-isopropylpyrazine and 2-Methoxy-3-isobutylpyrazine to the Major Urinary Protein. J. Am. Chem. Soc.

Bollinger, J.M., Chang, W.C., Matthews, M.L., Martinie, R.J., Boal, A.K., and Krebs, C. (2015). Mechanisms of 2-Oxoglutarate-Dependent Oxygenases: The Hydroxylation Pradigm and Beyond. In 2-Oxoglutarate-Dependent Oxygenases, C.J. Schofield, and R.P. Hausinger, eds. (The Royal Society of Chemistry, Cambridge), pp. 95–122.

Brinkmann, M., Blenkle, H., Salowsky, H., Bluhm, K., Schiwy, S., Tiehm, A., and Hollert, H. (2014). Genotoxicity of heterocyclic PAHs in the micronucleus assay with the fish liver cell line RTL-W1. PLoS One. Butler, C.S., and Mason, J.R. (1996). Structure-function Analysis of the Bacterial Aromatic Ring-hydroxylating Dioxygenases. R.K.B.T.-A. in M.P. Poole, ed. (Academic Press), pp. 47–84.

Capotorti, G., Digianvincenzo, P., Cesti, P., Bernardi, A., and Guglielmetti, G. (2004). Pyrene and benzo(a)pyrene metabolism by an Aspergillus terreus strain isolated from a polycylic aromatic hydrocarbons polluted soil. Biodegradation 15, 79–85.

Capyk, J.K., D’Angelo, I., Strynadka, N.C., and Eltis, L.D. (2009). Characterization of 3-ketosteroid 9α-hydroxylase, a Rieske oxygenase in the cholesterol degradation pathway of Mycobacterium tuberculosis. J. Biol. Chem. 284, 9937–9946.

Carredano, E., Karlsson, A., Kauppi, B., Choudhury, D., Parales, R.E., Parales, J. V., Lee, K., Gibson, D.T., Eklund, H., and Ramaswamy, S. (2000). Substrate binding site of naphthalene 1,2-dioxygenase: Functional implications of indole binding. J. Mol. Biol. 296, 701–712.

Cerniglia, C.E. (1992). Biodegradation of polycyclic aromatic hydrocarbons. Biodegradation 3, 351–368. Cheung, P.Y., and Kinkle, B.K. (2001). Mycobacterium Diversity and Pyrene Mineralization in Petroleum-Contaminated Soils. Appl. Environ. Microbiol. 67, 2222–2229.

Chiu, Y.-C., Okajima, T., Murakawa, T., Uchida, M., Taki, M., Hirota, S., Kim, M., Yamaguchi, H., Kawano, Y., Kamiya, N., et al. (2006). Kinetic and structural studies on the catalytic role of the aspartic acid residue conserved in copper amine oxidase. Biochemistry 45, 4105–4120.

Cho, J., Jeon, S., Wilson, S.A., Liu, L. V., Kang, E.A., Braymer, J.J., Lim, M.H., Hedman, B., Hodgson, K.O., Valentine, J.S., et al. (2011). Structure and reactivity of a mononuclear non-haem iron(III)-peroxo complex. Nature.

Chodera, J.D., and Mobley, D.L. (2013). Entropy-Enthalpy Compensation: Role and Ramifications in Biomolecular Ligand Recognition and Design. Annu. Rev. Biophys.

Costas, M., Mehn, M.P., Jensen, M.P., and Que, L. (2004). Dioxygen Activation at Mononuclear Nonheme Iron Active Sites: Enzymes, Models, and Intermediates. Chem. Rev.

D’Ordine, R.L., Rydel, T.J., Storek, M.J., Sturman, E.J., Moshiri, F., Bartlett, R.K., Brown, G.R., Eilers, R.J., Dart, C., Qi, Y., et al. (2009). Dicamba Monooxygenase: Structural Insights into a Dynamic Rieske Oxygenase that Catalyzes an Exocyclic Monooxygenation. J. Mol. Biol. 392, 481–497.

Daughtry, K.D., Xiao, Y., Stoner-Ma, D., Cho, E., Orville, A.M., Liu, P., and Allen, K.N. (2012). Quaternary ammonium oxidative demethylation: X-ray crystallographic, resonance raman, and uv-visible spectroscopic analysis of a Rieske-type demethylase. J. Am. Chem. Soc.

DeLano, W.L. (2002). The PyMOL Molecular Graphics System. Schrödinger LLC Wwwpymolorg Version 1., http://www.pymol.org. van der Donk, W.A., Krebs, C., and Bollinger, J.M. (2010). Substrate activation by iron superoxo intermediates. Curr. Opin. Struct. Biol.

Dumitru, R., Jiang, W.Z., Weeks, D.P., and Wilson, M.A. (2009). Crystal Structure of Dicamba Monooxygenase: A Rieske Nonheme Oxygenase that Catalyzes Oxidative Demethylation. J. Mol. Biol. 392, 498–510.

Dutson, S.M., Booth, G.M., Schaalje, G.B., Castle, R.N., and Seegmiller, R.E. (1997). Comparative developmental dermal toxicity and mutagenicity of carbazole and benzo[a]carbazole. Environ. Toxicol. Chem. Eisentraeger, A., Brinkmann, C., Hollert, H., Sagner, A., Tiehm, A., and Neuwoehner, J. (2008). Heterocyclic compounds: Toxic effects using algae, daphnids, and the Salmonella/microsome test taking methodical quantitative aspects into account. Environ. Toxicol. Chem.

Emsley, P., Lohkamp, B., Scott, W.G., and Cowtan, K. (2010). Features and development of Coot. Acta Crystallogr. Sect. D Biol. Crystallogr. 66, 486–501.

Ferraro, D.J., Gakhar, L., and Ramaswamy, S. (2005). Rieske business: Structure-function of Rieske non-heme oxygenases. Biochem. Biophys. Res. Commun. 338, 175–190.

Fuchs, G., Boll, M., and Heider, J. (2011). Microbial degradation of aromatic compounds- From one strategy to four. Nat. Rev. Microbiol. 9, 803–816.

Fuentes, S., Méndez, V., Aguila, P., and Seeger, M. (2014). Bioremediation of petroleum hydrocarbons: Catabolic genes, microbial communities, and applications. Appl. Microbiol. Biotechnol.

Furukawa, K., Suenaga, H., and Goto, M. (2004). Biphenyl dioxygenases: Functional versatilities and directed evolution. J. Bacteriol.

Ghosal, D., Ghosh, S., Dutta, T.K., and Ahn, Y. (2016). Current State of Knowledge in Microbial Degradation of Polycyclic Aromatic Hydrocarbons (PAHs): A Review. 7.

Gibson, D.T., and Parales, R.E. (2000). Aromatic hydrocarbon dioxygenases in environmental biotechnology. Curr. Opin. Biotechnol. 11, 236–243.

Grifoll, M., Selifonov, S.A., and Chapman, P.J. (1995). Transformation of substituted fluorenes and fluorene analogs by Pseudomonas sp. strain F274. Appl. Environ. Microbiol.

Hajdu, J., Neutze, R., Sjögren, T., Edman, K., Szöke, A., Wilmouth, R.C., and Wilmot, C.M. (2000). Analyzing protein functions in four dimensions. Nat. Struct. Biol. 7, 1006–1012.

Hegg, E.L., and Que, L. (1997). The 2-His-1-carboxylate facial triad - An emerging structural motif in mononuclear non-heme iron(II) enzymes. Eur. J. Biochem. 250, 625–629.

Hubel, A., and Skubitz, A.P.N. (2017). Principles of cryopreservation. Biobanking Hum. Biospecimens Princ. Pract. 368, 1–21.

Hudlicky, T., Gonzalez, D., and Gibson, D.T. (1999). Enzymatic dihydroxylation of aromatics in enantioselective synthesis: Expanding asymmetric methodology. Aldrichimica Acta 32, 35–62.

Inoue, K., Habe, H., Yamane, H., Omori, T., and Nojiri, H. (2005). Diversity of carbazole-degrading bacteria having the car gene cluster: Isolation of a novel gram-positive carbazole-degrading bacterium. FEMS Microbiol. Lett. 245, 145–153.

Inoue, K., Habe, H., Yamane, H., and Nojiri, H. (2006). Characterization of novel carbazole catabolism genes from gram-positive carbazole degrader Nocardioides aromaticivorans IC177. Appl. Environ. Microbiol. 72, 3321–3329.

Inoue, K., Ashikawa, Y., Umeda, T., Abo, M., Katsuki, J., Usami, Y., Noguchi, H., Fujimoto, Z., Terada, T., Yamane, H., et al. (2009). Specific Interactions between the Ferredoxin and Terminal Oxygenase Components of a Class IIB Rieske Nonheme Iron Oxygenase, Carbazole 1,9a-Dioxygenase. J. Mol. Biol. 392, 436–451.

INOUE, K., WIDADA, J., NAKAI, S., ENDOH, T., URATA, M., ASHIKAWA, Y., SHINTANI, M., SAIKI, Y., YOSHIDA, T., HABE, H., et al. (2004). Divergent Structures of Carbazole Degradative car Operons Isolated from Gram-negative Bacteria. Biosci. Biotechnol. Biochem. 68, 1467–1480.

Jeoung, J.-H., Bommer, M., Lin, T.-Y., and Dobbek, H. (2013). Visualizing the substrate-, superoxo-, alkylperoxo-, and product-bound states at the nonheme Fe(II) site of homogentisate dioxygenase. Proc. Natl. Acad. Sci. 110, 12625–12630.

Jha, A.M., and Bharti, M.K. (2002). Mutagenic profiles of carbazole in the male germ cells of Swiss albino mice. Mutat. Res. - Fundam. Mol. Mech. Mutagen.

Kal, S., and Que, L. (2017). Dioxygen activation by nonheme iron enzymes with the 2-His-1-carboxylate facial triad that generate high-valent oxoiron oxidants. J. Biol. Inorg. Chem. 22, 339–365.

Karlsson, A., Parales, J. V., Parales, R.E., Gibson, D.T., Eklund, H., and Ramaswamy, S. (2003). Crystal structure of naphthalene dioxygenase: Side-on binding of dioxygen to iron. Science (80-. ). 299, 1039–1042. Kauppi, B., Lee, K., Carredano, E., Parales, R.E., Gibson, D.T., Eklund, H., and Ramaswamy, S. (1998). Structure of an aromatic-ring-hydroxylating dioxygenase – naphthalene 1,2-dioxygenase. Structure 6, 571–586. Kawasaki, Y., and Freire, E. (2011). Finding a better path to drug selectivity. Drug Discov. Today.

Kawasaki, Y., Chufan, E.E., Lafont, V., Hidaka, K., Kiso, Y., Mario Amzel, L., and Freire, E. (2010). How much binding affinity can be gained by filling a cavity? Chem. Biol. Drug Des.

Knoot, C.J., Purpero, V.M., and Lipscomb, J.D. (2015). Crystal structures of alkylperoxo and anhydride intermediates in an intradiol ring-cleaving dioxygenase. Proc. Natl. Acad. Sci. 112, 388–393.

Koehntop, K.D., Emerson, J.P., and Que, L. (2005). The 2-His-1-carboxylate facial triad: A versatile platform for dioxygen activation by mononuclear non-heme iron(II) enzymes. J. Biol. Inorg. Chem. 10, 87–93.

Kovaleva, E.G., and Lipscomb, J.D. (2007). Crystal structures of Fe2+dioxygenase superoxo, alkylperoxo, and bound product intermediates. Science (80-. ). 316, 453–457.

Kovaleva, E.G., and Lipscomb, J.D. (2008). Versatility of biological non-heme Fe(II) centers in oxygen activation reactions. Nat. Chem. Biol. 4, 186–193.

Krebs, C., Fujimori, D.G., Walsh, C.T., and Bollinger, J.M. (2007). Non-Heme Fe ( IV )– Oxo Intermediates. Acc. Chem. Res. 40, 484–492.

Ladbury, J.E., Klebe, G., and Freire, E. (2010). Adding calorimetric data to decision making in lead discovery: A hot tip. Nat. Rev. Drug Discov.

Lau, P.C., and Lorenzo, V. (1999). Peer reviewed: genetic engineering: the frontier of bioremediation. Environ. Sci. Technol. 33, 124A–8A.

Leavitt, S., and Freire, E. (2001). Direct measurement of protein binding energetics by isothermal titration calorimetry. Curr. Opin. Struct. Biol.

Lebedev, A.A., Young, P., Isupov, M.N., Moroz, O. V., Vagin, A.A., and Murshudov, G.N. (2012). JLigand: A graphical tool for the CCP4 template-restraint library. Acta Crystallogr. Sect. D Biol. Crystallogr. 68, 431–440. Lee, Y.H., Ikegami, T., Standley, D.M., Sakurai, K., Hase, T., and Goto, Y. (2011). Binding energetics of ferredoxin-NADP+reductase with ferredoxin and its relation to function. ChemBioChem.

Li, F., Meier, K.K., Cranswick, M.A., Chakrabarti, M., Van Heuvelen, K.M., Münck, E., and Que, L. (2011). Characterization of a high-spin non-heme FeIII-OOH intermediate and its quantitative conversion to an FeIV=O complex. J. Am. Chem. Soc.

Lipscomb, J.D. (2008). Mechanism of extradiol aromatic ring-cleaving dioxygenases. Curr. Opin. Struct. Biol. 18, 644–649.

Lobastova, T.G., Sukhodolskaya, G. V., Nikolayeva, V.M., Baskunov, B.P., Turchin, K.F., and Donova, M. V. (2004). Hydroxylation of carbazoles by Aspergillus flavus VKM F-1024. FEMS Microbiol. Lett.

Martins, B.M., Svetlitchnaia, T., and Dobbek, H. (2005). 2-Oxoquinoline 8-monooxygenase oxygenase component: Active site modulation by Rieske-[2Fe-2S] center oxidation/reduction. Structure 13, 817–824.

Mason, J. (1992). The Electron-Transport Proteins of Hydroxylating Bacterial Dioxygenases. Annu. Rev. Microbiol. 46, 277–305.

Matsuzawa, J., Umeda, T., Aikawa, H., Suzuki, C., Fujimoto, Z., Okada, K., Yamane, H., and Nojiri, H. (2013). Crystallization and preliminary X-ray diffraction studies of the reduced form of the terminal oxygenase component of the Rieske nonhaem iron oxygenase system carbazole 1,9a-dioxygenase. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 69, 1284–1287.

Menzie, C.A., Potocki, B.B., and Joseph, S. (1992). Exposure to Carcinogenic PAHs in The Environment. Environ. Sci. Technol. 26, 1278–1284.

Mumbo, J., Henkelmann, B., Abdelaziz, A., Pfister, G., Nguyen, N., Schroll, R., Munch, J.C., and Schramm, K.W. (2014). Persistence and dioxin-like toxicity of carbazole and chlorocarbazoles in soil. Environ. Sci. Pollut. Res.

Murshudov, G.N., Skubák, P., Lebedev, A.A., Pannu, N.S., Steiner, R.A., Nicholls, R.A., Winn, M.D., Long, F., and Vagin, A.A. (2011). REFMAC5 for the refinement of macromolecular crystal structures. Acta Crystallogr. Sect. D Biol. Crystallogr. 67, 355–367.

Nam, J.W., Nojiri, H., Noguchi, H., Uchimura, H., Yoshida, T., Habe, H., Yamane, H., and Omori, T. (2002). Purification and characterization of carbazole 1,9a-dioxygenase, a three-component dioxygenase system of Pseudomonas resinovorans strain CA10. Appl. Environ. Microbiol. 68, 5882–5890.

Nam, J.W., Noguchi, H., Fujimoto, Z., Mizuno, H., Ashikawa, Y., Abo, M., Fushinobu, S., Kobashi, N., Wakagi, T., Iwata, K., et al. (2005). Crystal structure of the ferredoxin component of carbazole 1,9a-dioxygenase of Pseudomonas resinovorans strain CA10, a novel Rieske non-heme iron oxygenase system. Proteins Struct. Funct. Genet. 58, 779–789.

Neibergall, M.B., Stubna, A., Mekmouche, Y., Münck, E., and Lipscomb, J.D. (2007). Hydrogen peroxide dependent cis-dihydroxylation of benzoate by fully oxidized benzoate 1,2-dioxygenase. Biochemistry.

Nestler, F.H.M. (1974). Characterization of wood-preserving coal-tar creosote by gas-liquid chromatography. Anal. Chem.

Nojiri, H., Nam, J.W., Kosaka, M., Morii, K.I., Takemura, T., Furihata, K., Yamane, H., and Omori, T. (1999). Diverse oxygenations catalyzed by carbazole 1,9a-dioxygenase from pseudomonas sp. strain ca10. J. Bacteriol. 181, 3105–3113.

Nojiri, H., Ashikawa, Y., Noguchi, H., Nam, J.W., Urata, M., Fujimoto, Z., Uchimura, H., Terada, T., Nakamura, S., Shimizu, K., et al. (2005). Structure of the terminal oxygenase component of angular dioxygenase, carbazole 1,9a-dioxygenase. J. Mol. Biol. 351, 355–370.

NOJIRI, H. (2012). Structural and Molecular Genetic Analyses of the Bacterial Carbazole Degradation System. Biosci. Biotechnol. Biochem.

NOJIRI, H., and OMORI, T. (2002). Molecular Bases of Aerobic Bacterial Degradation of Dioxins: Involvement of Angular Dioxygenation. Biosci. Biotechnol. Biochem. 66, 2001–2016.

Otwinowski, Z., Minor, W., and Mode, O. (1997). Processing of X-Ray Diffraction Data Collected in Oscillation Mode BT - Methods in Enzymology. Methods Enzymol. 276, 307–326.

Ouchiyama, N., Zhang, Y., Omori, T., and Kodama, T. (1993). Biodegradation of Carbazole by Pseudomonas spp. CA06 and CA10. Biosci. Biotechnol. Biochem.

Parales, J. V., Parales, R.E., Resnick, S.M., and Gibson, D.T. (1998a). Enzyme specificity of 2-nitrotoluene 2,3-dioxygenase from Pseudomonas sp. strain JS42 is determined by the C-terminal region of the ?? subunit of the oxygenase component. J. Bacteriol. 180, 1194–1199.

Parales, R.E., Emig, M.D., Lynch, N.A., and Gibson, D.T. (1998b). Substrate specificities of hybrid naphthalene and 2,4-dinitrotoluene dioxygenase enzyme systems. J. Bacteriol. 180, 2337–2344.

Parales, R.E., Parales, J. V., and Gibson, D.T. (1999). Aspartate 205 in the catalytic domain of naphthalene dioxygenase is essential for activity. J. Bacteriol. 181, 1831–1837.

Peddinghaus, S., Brinkmann, M., Bluhm, K., Sagner, A., Hinger, G., Braunbeck, T., Eisenträger, A., Tiehm, A., Hollert, H., and Keiter, S.H. (2012). Quantitative assessment of the embryotoxic potential of NSO-heterocyclic compounds using zebrafish (Danio rerio). Reprod. Toxicol.

Penfield, J.S., Worrall, L.J., Strynadka, N.C., and Eltis, L.D. (2014). Substrate specificities and conformational flexibility of 3-ketosteroid 9??-hydroxylases. J. Biol. Chem.

Peters, W.B., Frasca, V., and Brown, R.K. (2009). Recent developments in isothermal titration calorimetry label free screening. Comb. Chem. High Throughput Screen.

Prigge, S.T., Eipper, B.A., Mains, R.E., and Amzel, L.M. (2004). Dioxygen Binds End-On to Mononuclear Copper in a Precatalytic Enzyme Complex. Science (80-. ). 304, 864–867.

Que, L. (2000). One motif - Many different reactions. Nat. Struct. Biol. 7, 182–184.

Rivard, B.S., Rogers, M.S., Marell, D.J., Neibergall, M.B., Chakrabarty, S., Cramer, C.J., and Lipscomb, J.D. (2015). Rate-Determining Attack on Substrate Precedes Rieske Cluster Oxidation during Cis-Dihydroxylation by Benzoate Dioxygenase. Biochemistry 54, 4652–4664.

Roelfes, G., Vrajmasu, V., Chen, K., Ho, R.Y.N., Rohde, J.U., Zondervan, C., La Crois, R.M., Schudde, E.P., Lutz, M., Spek, A.L., et al. (2003). End-on and side-on peroxo derivatives of non-heme iron complexes with pentadentate ligands: Models for putative intermediates in biological iron/dioxygen chemistry. In Inorganic Chemistry, pp. 2639–2653.

Salam, L.B., Ilori, M.O., and Amund, O.O. (2017). Properties, environmental fate and biodegradation of carbazole. 3 Biotech.

Sambrook, J., and W Russell, D. (2001). Molecular Cloning: A Laboratory Manual. Cold Spring Harb. Lab. Press. Cold Spring Harb. NY.

Sato, S.I., Nam, J.W., Kasuga, K., Nojiri, H., Yamane, H., and Omori, T. (1997). Identification and characterization of genes encoding carbazole 1,9a-dioxygenase in Pseudomonas sp. strain CA10. J. Bacteriol. 179, 4850–4858.

Seo, J., Keum, Y., and Li, Q.X. (2009). Bacterial Degradation of Aromatic Compounds. Int. J. Environ. Res. Public Health 6, 278–309.

Shintani, M., Urata, M., Inoue, K., Eto, K., Habe, H., Omori, T., Yamane, H., and Nojiri, H. (2007). The Sphingomonas plasmid pCAR3 is involved in complete mineralization of carbazole. J. Bacteriol. 189, 2007–2020.

Solomon, E.I., Brunold, T.C., Davis, M.I., Kemsley, J.N., Lee, S.-K., Lehnert, N., Neese, F., Skulan, A.J., Yang, Y.-S., and Zhou, J. (2000). Geometric and Electronic Structure/Function Correlations in Non-Heme Iron Enzymes. Chem. Rev. 100, 235–350.

Tamanaha, E., Zhang, B., Guo, Y., Chang, W.C., Barr, E.W., Xing, G., St Clair, J., Ye, S., Neese, F., Bollinger, J.M., et al. (2016). Spectroscopic Evidence for the Two C-H-Cleaving Intermediates of Aspergillus nidulans Isopenicillin N Synthase. J. Am. Chem. Soc.

Tarasev, M., and Ballou, D.P. (2005). Chemistry of the catalytic conversion of phthalate into its cis-dihydrodiol during the reaction of oxygen with the reduced form of phthalate dioxygenase. Biochemistry 44, 6197–6207. Timmis, K.N., and Pieper, D.H. (1999). Bacteria designed for bioremediation. Trends Biotechnol. 17, 201–204. Tsuda, H., Hagiwara, A., Shibata, M., Ohshima, M., and Ito, N. (1982). Carcinogenic effect of carbazole in the liver of (C57BL/6N x C3H/HeN)F1 mice. J. Natl. Cancer Inst.

Urata, M., Uchimura, H., Noguchi, H., Sakaguchi, T., Takemura, T., Eto, K., Habe, H., Omori, T., Yamane, H., and Nojiri, H. (2006). Plasmid pCAR3 contains multiple gene sets involved in the conversion of carbazole to anthranilate. Appl. Environ. Microbiol. 72, 3198–3205.

Vagin, A., and Teplyakov, A. (2010). Molecular replacement with MOLREP. Acta Crystallogr. Sect. D Biol. Crystallogr. 66, 22–25.

Varjani, S.J. (2017). Microbial degradation of petroleum hydrocarbons. Bioresour. Technol.

Vejarano, F., Suzuki-Minakuchi, C., Ohtsubo, Y., Tsuda, M., Okada, K., and Nojiri, H. (2018). Complete Genome Sequence of the Marine Carbazole-Degrading Bacterium Erythrobacter sp. Strain KY5. Microbiol. Resour. Announc.

Wang, Y., Li, J., and Liu, A. (2017). Oxygen activation by mononuclear nonheme iron dioxygenases involved in the degradation of aromatics. J. Biol. Inorg. Chem. 22, 395–405.

Winn, M.D., Ballard, C.C., Cowtan, K.D., Dodson, E.J., Emsley, P., Evans, P.R., Keegan, R.M., Krissinel, E.B., Leslie, A.G.W., McCoy, A., et al. (2011). Overview of the CCP4 suite and current developments. Acta Crystallogr. Sect. D Biol. Crystallogr. 67, 235–242.

Wiseman, T., Williston, S., Brandts, J.F., and Lin, L.N. (1989). Rapid measurement of binding constants and heats of binding using a new titration calorimeter. Anal. Biochem.

Wolfe, M.D., and Lipscomb, J.D. (2003). Hydrogen peroxide-coupled cis-diol formation catalyzed by naphthalene 1,2-dioxygenase. J. Biol. Chem.

Wolfe, M.D., Parales, J. V., Gibson, D.T., and Lipscomb, J.D. (2001). Single turnover chemistry and regulation of O2activation by the oxygenase component of naphthalene 1,2-dioxygenase. J. Biol. Chem. 276, 1945–1953. Wolfe, M.D., Altier, D.J., Stubna, A., Popescu, C. V., Münck, E., and Lipscomb, J.D. (2002). Benzoate 1,2-dioxygenase from Pseudomonas putida: Single turnover kinetics and regulation of a two-component rieske dioxygenase. Biochemistry 41, 9611–9626.

Zhang, Y., and Skolnick, J. (2005). TM-align: A protein structure alignment algorithm based on the TM-score. Nucleic Acids Res. 33, 2302–2309.

Zidek, L., Novotny, M. V., and Stone, M.J. (1999). Increased protein backbone conformational entropy upon hydrophobic ligand binding. Nat. Struct. Biol. Matsuzawa, Master thesis, 2013. Nam, Doctoral thesis, 2000.

参考文献をもっと見る