J Biol Inorg Chem DOI 10.1007/s00775-014-1106-9

MINIREVIEW

Peroxygenase reactions catalyzed by cytochromes P450 Osami Shoji • Yoshihito Watanabe

Received: 1 October 2013 / Accepted: 7 January 2014 Ó SBIC 2014

Abstract Cytochromes P450 (P450s) catalyze monooxygenation of a wide range of less reactive organic molecules under mild conditions. By contrast with the general reductive oxygen activation pathway of P450s, an H2O2shunt pathway does not require any supply of electrons and protons for the generation of a highly reactive intermediate (compound I). Because the low cost of H2O2 allows us to use it in industrial-scale synthesis, the H2O2-shunt pathway is an attractive process for monooxygenation reactions. This review focuses on the P450-catalyzed monooxygenation of organic molecules using H2O2 as the oxidant. Keywords Cytochrome  Heme  X-ray crystallography  Site-directed mutagenesis  Dioxygen Introduction Cytochromes P450 (P450s) are a family of heme–thiolate enzymes that catalyze monooxygenation of drugs,

Responsible editors: Lucia Banci and Claudio Luchinat. O. Shoji and Y. Watanabe equally contributed to this work. O. Shoji (&) Department of Chemistry, Graduate School of Science, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8602, Japan e-mail: [email protected]

detoxification of xenobiotics, and biosynthesis of steroids [1, 2]. One of the most interesting features of P450s is that they catalyze monooxygenation of a wide range of less reactive organic molecules under mild conditions. Construction of biocatalysts based on P450s has therefore attracted much attention [1–11]. In general, P450s reductively activate molecular oxygen by using two electrons derived from a cofactor, NADH or NADPH, and two protons from outside the protein to generate an active species, ‘‘iron–oxo porphyrin p-cation radical,’’ which is called ‘‘compound I’’ after its discovery in peroxidases and catalases (Fig. 1) [12]. Hydrogen peroxide can also be used as an oxidant for the generation of the active species, through a ‘‘shunt reaction.’’ The use of H2O2 is an attractive option for monooxygenation reactions catalyzed by P450s, because electron transfer partners such as P450 reductase are not required for the generation of an active species. Furthermore, the low cost of H2O2 allows us to perform monooxygenations through the shunt path on an industrial scale. Thus, the H2O2-dependent P450 system has been considered as an important candidate for practical applications. This review focuses on the P450-catalyzed monooxygenation of organic molecules by using H2O2 as the oxidant. In the third section, artificial H2O2-dependent P450s created by point mutagenesis and directed evolution are described. In the final section, we summarize natural H2O2-dependent P450s and decoy molecule systems for peroxygenations of nonnative substrates.

H2O2-shunt reaction of P450s Y. Watanabe (&) Research Center for Materials Science, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8602, Japan e-mail: [email protected]

P450s are monooxygenases, and their catalytic cycle of monooxygenation begins with a reductive activation of molecular oxygen by using two electrons derived from a

123

J Biol Inorg Chem Fig. 1 a The crystal structure of P450 BM3 [Protein Data Bank (PDB) code 1FAG]. b The catalytic cycle of oxygen activation by cytochromes 450 (P450s) and its H2O2-shunt pathway

cofactor (NADH or NADPH) and two protons (Fig. 1). By contrast, P450s are known to precede an H2O2-shunt reaction where a supply of electrons and protons is not required for oxidation. In the H2O2-shunt reaction, the ferric heme iron reacts directly with H2O2 followed by a heterolytic cleavage of the O–O bond to give a highly reactive intermediate, so-called compound I that is the species responsible for the monooxygenation. The H2O2shunt reaction has been applied to a variety of oxidations that are commonly catalyzed by the usual P450 reaction system (Fig. 2). P450s examined include rabbit liver microsomal P450 [13, 14], rat liver microsomal P450 [15– 17], CYP119 [18], EpoK [19], CYP1A2 [20], CYP2S1 [21], CYP107AJ1 [22], CYP175A1 [23], P450eryF [24], P450st [25], CYP3A4 [26], and P450cam [27]. However, in general, the H2O2-shunt reaction is inefficient, and the catalytic activities of P450s are lower than those of reactions using cumene hydroperoxide or m-chloroperbenzoic acid [13, 28]. The crystal structures of P450s reported so far have revealed that amino acid residues that serve as general acid–base catalysts in peroxidases and catalase (histidine, aspartic acid, and glutamic acid) are not observed in the distal side of the heme (Fig. 3). General acid–base residues are crucial for the generation of compound I, which is responsible for H2O2-dependent oxidation (peroxidation). Indeed, most H2O2-dependent heme enzymes, such as horseradish peroxidase [29], chloroperoxidase [30, 31], aromatic peroxygenase [32–34], cytochrome c peroxidase [35], catalase-peroxidase [36], and catalase [37], have general acid–base residues close to the heme iron (Figs. 4, 5). The critical role of the general acid– base residues has been studied using a myoglobin

123

framework. When a general acid–base residue is introduced into an appropriate position of the distal side of myoglobin by point mutagenesis, the formation of myoglobin compound I is accelerated, and higher peroxidase and peroxygenase activities are observed [38–40]. The lack of any general acid–base residue in the active site of P450s and the high hydrophobicity of the heme pocket of P450s result in low efficiency for the formation of compound I and the subsequent peroxidation. The distal side of most P450s is composed of hydrophobic amino acids residues, except for the conserved threonine. The conserved threonine is crucial for the reductive activation of molecular oxygen, and thus mutations of the conserved threonine in P450cam to alanine or valine resulted in H2O2 production (uncoupling of P450) rather than oxidation of the substrate [41, 42]. Although the conserved hydroxyl group in the active site of P450 is crucial for activating molecular oxygen as a proton donor, this alcohol side chain is not expected to serve as an acid–base catalyst for the facile generation of compound I when H2O2 is using as an oxidant, because the pKa of the alcohol side chain is too high.

Artificial H2O2-dependent P450s To construct artificial H2O2-dependent P450s, a variety of mutants (Figs. 6, 7) have been prepared by site-directed mutagenesis and random mutagenesis (directed evolution). Pioneering work in the development of artificial H2O2dependent P450s was reported by Li et al. [43] in 2001. They prepared a series of Phe-87 mutants of P450 BM3 (heme domain) and succeeded in oxidizing myristic acid

J Biol Inorg Chem Fig. 2 Peroxygenase reactions catalyzed by P450s: a benzphetamine [13], b aniline [14], c androstenedione [15], d methoxyresorufin [19], e 2amino-3-methylimidazo[4,5f]quinolone [20], f styrene [18], g 7,12dimethylbenz[a]anthracene [21], h palmitoleic acid [23], and i 7-ethoxycoumarin [22]. CYP cytochrome P450, WT wild type

123

J Biol Inorg Chem Fig. 3 The active-site structures of a P450cam (PDB code 2CPP) [89], b CYP119 (PDB code 1F4T) [90], c CYP3A4 (PDB code 1TQN) [91], and d CYP175A1 (PDB code 1N97) [92]

Fig. 4 The active-site structures of horseradish peroxidase (PDB code1ATJ) [29], chloroperoxidase (PDB code 1CPO) [30, 31], aromatic peroxygenase (PDB code 2YP1) [32], cytochrome c peroxidase (PDB code 1CCA) [35], catalase-peroxidase (PDB code 1ITK) [36], and catalase (PDB code 4BLC) [37]

Fig. 5 General acid–base function of distal histidine in the formation of compound I using H2O2 as an oxidant

123

J Biol Inorg Chem

and p-nitrophenoxydodecanoic acid [44] using H2O2 as an oxidant. Among the mutants examined, an F87A mutant showed the highest catalytic activity, 162 nmol/min/nmol P450. However, its Michaelis–Menten constant (Km) for H2O2 was estimated to be 24 mM, which is very high compared with those of P450s that naturally use H2O2 as an oxidant; for example, 50–72 lM for P450SPa [45] and 21 lM for P450BSb, respectively [46]. They also demonstrated the transformation of indole to indigo using P450 BM3 mutants and H2O2 [47], in which the F87V mutant afforded the highest activity. Cirino and Arnold [48] also used the F87V mutant of P450 BM3 and developed further improved mutants by applying a directed evolution technique (random mutation) based on the F87A mutant. In the peroxygenation of p-nitrophenoxydodecanoic acid, the initial turnover rate of the resulting mutant containing a nine amino acid substitution in addition to F87A (I58V, H100R, F107L, A135S, M145V, N239H, S274T, K434E, and V446I), named 21B3, was more than 18 times higher than that of the F87A mutant [49]. Its Michaelis–Menten constant for H2O2 was also improved from 24 mM to 10 mM. The 21B3 mutant also catalyzed styrene epoxidation and transformation of several pesticides such as linuron (Fig. 6e) [50]. Furthermore, the thermostability of 21B3 was improved by random mutagenesis. The resulting mutant, named 5H6 (L52I, I58V, F87A, H100R, S106R, F107L, A135S, A184V, N239H, S274T, L324I, V340M, I366V, K434E, E442K, and V446I), resists thermal denaturation and retains peroxygenase activity [51]. The hydroxylation activity of P450 BM3 for p-nitrophenoxydodecanoic acid [44] was improved by preparing chimeras of CYP102A1 (P450 BM3) with F87A substitution and CYP102A2 with F88A substitution [52, 53]. The turnover rate of a chimera named 169–197 (the first and last residues of CYP102A2 were inserted into CYP102A1) reached 100.3 nmol/min/ nmol P450 [52]. To prepare authentic human metabolites of propranolol [54], the P450 BM3 mutant 9C1 with 13 amino acid substitutions (F87A, I58V, H100R, I102T, F107L, A135S, M145A, N239H, S274T, L324I, I366V, K434E, E442K, and V446I) was constructed, and further mutations of seven active-site residues (A74, L75, V78, F81, A82, F87, T88) were conducted on the basis of the crystal structure of P450 BM3 containing N-palmitoylglycine in its active site [55]. The resulting mutant, named 2C11 (A74V, A82L, F87G, I58V, H100R, I102T, F107L, A135S, M145A, N239H, S274T, L324I, I366V, K434E, E442K, and V446I) showed enhanced activity for the formation of the authentic human metabolites of propranolol. It is interesting to note that a double mutant of 21B3 (W96A/F405L) gave a stable variant showing a higher peroxidase activity than the 21B3 mutant [56]. Joo et al. [57] developed an H2O2-dependent P450 BM3 and H2O2-

dependent P450cam using directed evolution. A triple mutant of P450cam (E331K, R280L, and C242F) expressed in Escherichia coli showed enhanced naphthalene hydroxylation activity [57], whereas the purified mutant showed not much improved catalytic activity [58]. Kumar et al. [59] applied a directed evolution approach to screen P450 2B1 for an H2O2-supported reaction and developed a mutant (V183L, F202L, L209A, and S334P) with a more than 30-fold higher kcat (7.2 min-1) than the wild type (0.2 min-1) for O-deethylation of 7-ethoxy-4-trifluoromethylcoumarin. Many other mutants of P450 2B1 have been prepared, but kcat and Km (H2O2) of the resulting mutants such as L216W/F228I/T433S for the H2O2-supported reaction were not much improved [60]. Although many attempts have been devoted to improving the catalytic activities of P450 in the H2O2-shunt reaction by using mutants, there are not many mutants that show activities higher than those of the reductive oxygen activation process. It is noteworthy that any mutation introducing general acid–base residue(s) is not involved in artificial H2O2-dependent P450s. Mutations were sometimes introduced in places other than the active site. Therefore, no general acid–base residue in the active site was required for the facile generation of compound I in any artificial H2O2dependent P450s. These results suggest that water molecules exist in the active site and may serve as a general acid–base catalyst, even though further study is required to confirm the involvement of water molecules in the formation of compound I. If we could appropriately place an amino acid residue or residues that serve as a general acid–base catalyst in a suitable position in the active site of P450, we would obtain an artificial H2O2-dependent P450 with activity comparable to that of the natural P450s. It is interesting to note here that F87H, F87E, and F87D mutants of P450 BM3 were prepared to examine their effect on substrate selectivity and regioselectivity, but that the F87E and F87D mutants did not show any catalytic activity, whereas the F87H mutant showed a lower catalytic activity compared with the wild type. The CO difference spectra of the F87E and F87D mutants showed a peak at 420 nm indicating that these mutants are inactive forms; thus, they are unable to catalyze the monooxygenation through the general reductive oxygen activation pathway of P450s. We, however, presume that Phe-87 mutants would generate the active species through the H2O2-shunt pathway, because the location of Glu-87 on the distal side of P450 BM3 seems to be close enough to the heme iron to serve as a general acid–base catalyst. Similarly, we also presume that an A246E mutant of P450 BM3 would generate compound I by using H2O2, because the crystal structure of the heme domain of substrate-free A246E showed that the distance between an oxygen atom of the carboxylate group of Glu-246 and the heme iron is

123

J Biol Inorg Chem Fig. 6 Peroxygenase reactions catalyzed by P450 BM3 mutants: a pnitrophenoxydodecanoic acid [43, 48, 51, 52], b indole [47], c styrene [49], d propranolol [54], and e linuron [50]

˚ , whereas another carboxylate group of Glu-246 in 5.6 A another molecule of the asymmetric unit is coordinated with the heme iron [61, 62].

Natural H2O2-dependent P450s and the decoy molecule system for nonnative substrate peroxygenations Although most P450s use the reductive molecular oxygen activation process for their monooxygenations, P450SPa (CYP152B1) [45, 63–67] from Sphingomonas paucimobilis, P450BSb (CYP152A1) [46, 68–73] from Bacillus subtilis, and P450CLA (CYP152A2) [74] from Clostridium acetobutylicum use H2O2 as the oxidant and catalyze

123

hydroxylation of long-alkyl-chain fatty acids with high catalytic activities and high substrate specificity (Fig. 8). In addition, P450CLA catalyzes fatty acid hydroxylation through the reductive molecular oxygen activation pathway. Very recently, P450CLA and P450BSb were reported to perform peroxygenase reactions by the use of a light-driven system for in situ generation of H2O2 [75–78]. Among H2O2-dependent P450s, P450SPa reported by Matsunaga et al. [79] is the first P450 to be classified as an H2O2dependent P450. P450SPa catalyzes a-selective (100 %) hydroxylation of long-alkyl-chain fatty acids such as myristic acid and palmitic acid. In 2003, the crystal structure of P450BSb (CYP152A1) was reported at a reso˚ (Protein Data Bank code 1IZO) as the first lution of 2.1 A

J Biol Inorg Chem Fig. 7 Peroxygenase reactions catalyzed by P450 mutants: a naphthalene [57], b 7-ethoxy4-trifluoromethylcoumarin [59], and c 7-benzyloxyquinoline [60]

Fig. 8 Myristic acid hydroxylation catalyzed by H2O2-dependent P450s: a P450BSb [80], b P450SPa [80], and c P450CLA [74]

crystal structure of an H2O2-dependent P450 [69]. P450BSb has 42 % amino acid identity with P450SPa and oxidizes the a and b positions of fatty acids in a roughly 40:60 ratio. The crystal structure of a substrate-bound form of P450BSb revealed that P450BSb lacks general acid–base residues around the distal side of the heme, although general acid– base residues are highly conserved in peroxidases and peroxygenases, as described in ‘‘H2O2-shunt reaction of P450s.’’ Instead of the general acid–base residues, the terminal carboxylate group of the bound fatty acid interacts with the guanidinium group of Arg-242 located near the heme group (Fig. 9b). The distance between an oxygen atom of the carboxylate group of palmitic acid and the

˚ , which is very close to that of aromatic heme iron is 5.3 A ˚ ) [32] and chloroperoxidase (5.1 A ˚) peroxygenase (5.2 A [30, 31] (Fig. 4). Furthermore, the resulting salt bridge in the active site is very similar to that of aromatic peroxygenase [32]. The location of the carboxylate group of palmitic acid bound to P450BSb suggests that a general acid–base function for the facile formation of compound I is accomplished by the carboxylate group of the substrate (Fig. 9). Thus, the catalytic reaction begins with the fixation of a substrate by the interaction of the terminal carboxyl group of the fatty acid with Arg-242. The crystal structure of P450SPa con˚ (Protein Data taining palmitic acid at a resolution of 1.65 A Bank code 3AWM) was determined in 2011 [80]. The structure also revealed that the carboxylate group of the fatty acid interacts with Arg-241 located immediately above the heme. Two crystal structural studies confirm that the substrate-assisted activation mechanism is conserved in H2O2-dependent P450s. This unique reaction mechanism observed for P450BSb and P450SPa also contributes to their high substrate specificity. P450BSb and P450SPa do not oxidize substrates other than long-alkyl-chain fatty acids. For example, P450BSb never oxidizes substrates without a carboxyl group such as tetradecane, 1-tetradecanol, and tetradecanal. The fatty acid substrate itself serves both as an initiator and as an activator of the reaction, and thus P450SPa and P450BSb never oxidize substrates other than fatty acids.

123

J Biol Inorg Chem

Fig. 9 a A proposed catalytic hydroxylation mechanism for P450BSb and the roles of the substrate carboxylate–Arg-242 salt bridge. b The active-site structure of P450BSb containing palmitic acid (PDB code 1IZO). c The substrate misrecognition system in the oxidation of nonnative substrate (e.g., ethylbenzene) in the presence of heptanoic

acid as a decoy molecule. d The active-site structure of P450BSb cocrystallized with heptanoic acid (PDB code 2ZQX). An acetic acid molecule was placed in the electron density observed near the Arg242 residue

Recently, we found that P450BSb is able to oxidize a wide variety of nonnative substrates in the presence of a series of short-alkyl-chain carboxylic acids (C4–C10), which we call ‘‘decoy molecules’’ [81]. The peroxygenation of nonnative substrates never proceeds without decoy molecules, and thus these reactions are catalyzed by P450BSb with the help of these short-alkyl-chain carboxylic acids (Fig. 9c). Namely, P450BSb misrecognizes decoy molecules as its native substrates (long-alkyl-chain fatty acids) and the resulting salt bridge (Arg-242–carboxylate of fatty acid) serves as a general acid–base catalyst to allow the formation of an active species followed by the oxidation of nonnative substrates. The role of decoy molecules was confirmed on the basis of crystal structure analysis of a heptanoic acid bound form of P450BSb (Fig. 9d) [82]. A clear electron density assignable to the carboxylate group of heptanoic acid was observed in the active site, suggesting the participation of the carboxylate group of decoy molecules in the formation of compound I. Later, a clear structure of decoy-molecule-bound P450SPa was obtained, when (R)-ibuprofen was used as the decoy molecule (Fig. 10) [83]. By use of a simple substrate misrecognition

trick by using decoy molecules, nonnative substrate oxidations such as the one-electron oxidation of guaiacol [81], sulfoxidation of thioanisole [84], epoxidation of styrene [81], C–H bond hydroxylation of ethylbenzene [81], and aromatic C–H bond hydroxylation of 1-methoxynaphthalene [85] were catalyzed by P450BSb (Tables 1, 2). Interestingly, the catalytic activities are very much dependent on the alkyl-chain length of decoy molecules (Table 1). Furthermore, enantioselectivity in peroxygenation such as ethylbenzene hydroxylation is dependent on the structure of the decoy molecules used [81]. These results clearly show that the decoy molecule plays two important roles: (1) activation of P450BSb to form the active species compound I, and (2) regulation of the spatial orientation of the substrate. Therefore, the decoy molecule system can be applicable to synthetic chemistry. In the case of P450SPa, the stereoselectivity of styrene epoxidation was also controlled by the chirality of the decoy molecules (Fig. 10) [83]; whereas (R)-ibuprofen gave (S)-styrene oxide, (S)ibuprofen gave (R)-styrene oxide. More importantly, the enantioselectivity was predicted by docking simulation based on the structure of P450SPa containing (R)-ibuprofen

123

J Biol Inorg Chem Fig. 10 a Stereoselective epoxidation of styrene catalyzed by P450SPa in the presence of (R)-ibuprofen or S)-ibuprofen. b The active-site structure of the (R)-ibuprofen-bound form of P450SPa (PDB code 3VM4). ee% percent enantiomeric excess

Table 1 The effect of decoy molecules on the one-electron oxidation of guaiacol O

O OH

P450BSβ, H2O2

O

O O

Decoy molecules

O O

Decoy molecule

Turnover number/min

None

0

Acetic acid (2)

0

Propionic acid (3)

26

Butanoic acid (4)

230

Pentanoic acid (5)

1,900

Hexanoic acid (6)

2,420

Heptanoic acid (7)

3,750

Octanoic acid (8)

2,490

Nonanoic acid (9)

2,380

Decanoic acid (10)

2,360

Myristic acid (14)

14

Table 2 The reactions catalyzed by P450BSb in the presence of decoy molecules

The initial turnover rates in the presence and in the absence of carboxylic acids. The numbers in parentheses denote the number of carbon atoms.

[83]. The decoy molecule system does not require any substitution of amino acids to alter the substrate specificity of P450BSb or to alter the enantioselectivity of nonnative substrate oxidations. The decoy molecule system has also been applied to P450 BM3 in which a series of perfluorinated carboxylic acids serve as decoy molecules [86]. The ‘‘P450BM3-decoy molecule system’’ catalyzed gaseous alkane hydroxylation such as propane [86] and ethane [87] as well as benzene [88], whereas wild-type P450 BM3 never catalyzes these reactions.

Conclusions and future prospects In this review, we have described H2O2-dependent oxidations (peroxygenations) catalyzed by P450s. The H2O2shunt reaction of P450s, as practical biocatalysts, has been

considered as an attractive reaction system for the monooxygenation. The catalytic activities of most H2O2-shunt reactions, however, are lower than those of the natural reductive oxygen activation processes of P450s. The catalytic activities of H2O2-dependent reactions can be improved by mutagenesis. Poor activities of P450 mutants for H2O2-dependent reactions could be explained by the lack of general acid–base residues in the active site of P450s and their mutants. By contrast, for the natural H2O2dependent P450s, the general acid–base catalysts supplied by natural substrates (fatty acids) or decoy molecules greatly contribute to their high catalytic activities. These findings lead us to the conclusion that most P450s can be converted into efficient H2O2-dependent P450s, either by placing a general acid–base residue(s) at an appropriate position in the active site of P450s by mutagenesis or by

123

J Biol Inorg Chem

adding external molecules, such as decoy molecules, as an activator, or both. Acknowledgments This work was supported, in part, by a Grant-inAid for Scientific Research (S) to Y. W. (24225004), a Grant-in-Aid for Young Scientists (A) to O. S. (21685018), and a Grant-in-Aid for Scientific Research on Innovative Areas ‘‘Molecular Activation Directed Toward Straightforward Synthesis’’ to O.S. (25105724) from the Ministry of Education, Culture, Sports, Science, and Technology (Japan).

References 1. Ortiz de Montellano PR (ed) (2005) Cytochrome P450: structure, mechanism, and biochemistry. Plenum, New York 2. Sono M, Roach MP, Coulter ED, Dawson JH (1996) Chem Rev 96:2841–2888 3. Denisov IG, Makris TM, Sligar SG, Schlichting I (2005) Chem Rev 105:2253–2277 4. Whitehouse CJC, Bell SG, Wong LL (2012) Chem Soc Rev 41:1218–1260 5. Fasan R (2012) ACS Catal 2:647–666 6. Grogan G (2011) Curr Opin Chem Biol 15:241–248 7. Jung ST, Lauchli R, Arnold FH (2011) Curr Opin Biotechnol 22:809–817 8. O’Reilly E, Kohler V, Flitsch SL, Turner NJ (2011) Chem Commun 47:2490–2501 9. Schulz S, Girhard M, Urlacher VB (2012) ChemCatChem 4:1889–1895 10. Sakaki T (2012) Biol Pharm Bull 35:844–849 11. Urlacher VB, Girhard M (2012) Trends Biotechnol 30:26–36 12. Dunford HB, Stillman JS (1976) Coord Chem Rev 19:187–251 13. Nordblom GD, White RE, Coon MJ (1976) Arch Biochem Biophys 175:524–533 14. Renneberg R, Scheller F, Ruckpaul K, Pirrwitz J, Mohr P (1978) FEBS Lett 96:349–353 15. Hrycay EG, Gustafsson J-a, Ingelman-Sundberg M, Ernster L (1975) Biochem Biophys Res Commun 66:209–216 16. Renneberg R, Capdevila J, Chacos N, Estabrook RW, Prough RA (1981) Biochem Pharmacol 30:843–848 17. Holm KA, Engell RJ, Kupfer D (1985) Arch Biochem Biophys 237:477–489 18. Koo LS, Tschirret-Guth RA, Straub WE, Moenne-Loccoz P, Loehr TM, Ortiz de Montellano PR (2000) J Biol Chem 275:14112–14123 19. Ogura H, Nishida CR, Hoch UR, Perera R, Dawson JH, Ortiz de Montellano PR (2004) Biochemistry 43:14712–14721 20. Anari MR, Josephy PD, Henry T, O’Brien PJ (1997) Chem Res Toxicol 10:582–588 21. Bui PH, Hankinson O (2009) Mol Pharmacol 76:1031–1043 22. Niraula NP, Kanth BK, Sohng JK, Oh TJ (2011) Enzyme Microb Technol 48:181–186 23. Goyal S, Banerjee S, Mazumdar S (2012) Biochemistry 51:7880–7890 24. Khan KK, He YA, He YQ, Halpert JR (2002) Chem Res Toxicol 15:843–853 25. Matsumura H, Wakatabi M, Omi S, Ohtaki A, Nakamura N, Yohda M, Ohno H (2008) Biochemistry 47:4834–4842 26. Zhang Z, Li Y, Stearns RA, Ortiz De Montellano PR, Baillie TA, Tang W (2002) Biochemistry 41:2712–2718 27. Gelb MH, Heimbrook DC, Malkonen P, Sligar SG (1982) Biochemistry 21:370–377 28. Rabe KS, Kiko K, Niemeyer CM (2008) ChemBioChem 9:420–425

123

29. Gajhede M, Schuller DJ, Henriksen A, Smith AT, Poulos TL (1997) Nat Struct Biol 4:1032–1038 30. Sundaramoorthy M, Terner J, Poulos TL (1995) Structure 3:1367–1377 31. Sundaramoorthy M, Terner J, Poulos TL (1998) Chem Biol 5:461–473 32. Piontek K, Strittmatter E, Ullrich R, Grobe G, Pecyna MJ, Kluge M, Scheibner K, Hofrichter M, Plattner DA (2013) J Biol Chem 288:34767–34776 33. Wang XS, Peter S, Ullrich R, Hofrichter M, Groves JT (2013) Angew Chem Int Ed 52:9238–9241 34. Wang XS, Peter S, Kinne M, Hofrichter M, Groves JT (2012) J Am Chem Soc 134:12897–12900 35. Goodin DB, Mcree DE (1993) Biochemistry 32:3313–3324 36. Yamada Y, Fujiwara T, Sato T, Igarashi N, Tanaka N (2002) Nat Struct Biol 9:691–695 37. Ko T-P, Day J, Malkin AJ, McPherson A (1999) Acta Crystallogr Sect D 55:1383–1394 38. Watanabe Y, Ueno T (2003) Bull Chem Soc Jpn 76:1309–1322 39. Ozaki SI, Roach MP, Matsui T, Watanabe Y (2001) Acc Chem Res 34:818–825 40. Watanabe Y, Nakajima H, Ueno T (2007) Acc Chem Res 40:554–562 41. Imai M, Shimada H, Watanabe Y, Matsushima-Hibiya Y, Makino R, Koga H, Horiuchi T, Ishimura Y (1989) Proc Natl Acad Sci USA 86:7823–7827 42. Martinis SA, Atkins WM, Stayton PS, Sligar SG (1989) J Am Chem Soc 111:9252–9253 43. Li Q-S, Ogawa J, Shimizu S (2001) Biochem Biophys Res Commun 280:1258–1261 44. Schwaneberg U, Schmidt-Dannert C, Schmitt J, Schmid RD (1999) Anal Biochem 269:359–366 45. Matsunaga I, Sumimoto T, Ueda A, Kusunose E, Ichihara K (2000) Lipids 35:365–371 46. Matsunaga I, Ueda A, Sumimoto T, Ichihara K, Ayata M, Ogura H (2001) Arch Biochem Biophys 394:45–53 47. Li QS, Ogawa J, Schmid RD, Shimizu S (2005) Biosci Biotechnol Biochem 69:293–300 48. Cirino PC, Arnold FH (2002) Adv Synth Catal 344:932–937 49. Cirino PC, Arnold FH (2003) Angew Chem Int Ed 42:3299–3301 50. Sanchez-Sanchez L, Roman R, Vazquez-Duhalt R (2012) Pestic Biochem Physiol 102:169–174 51. Salazar O, Cirino PC, Arnold FH (2003) ChemBioChem 4:891–893 52. Otey CR, Silberg JJ, Voigt CA, Endelman JB, Bandara G, Arnold FH (2004) Chem Biol 11:309–318 53. Otey CR, Landwehr M, Endelman JB, Hiraga K, Bloom JD, Arnold FH (2006) PLoS Biol 4:789–798 54. Otey CR, Bandara G, Lalonde J, Takahashi K, Arnold FH (2006) Biotechnol Bioeng 93:494–499 55. Haines DC, Tomchick DR, Machius M, Peterson JA (2001) Biochemistry 40:13456–13465 56. Vidal-Limon A, Aguila S, Ayala M, Batista CV, Vazquez-Duhalt R (2013) J Inorg Biochem 122:18–26 57. Joo H, Lin ZL, Arnold FH (1999) Nature 399:670–673 58. Matsuura K, Tosha T, Yoshioka S, Takahashi S, Ishimori K, Morishima I (2004) Biochem Biophys Res Commun 323:1209–1215 59. Kumar S, Chen CS, Waxman DJ, Halpert JR (2005) J Biol Chem 280:19569–19575 60. Kumar S, Liu H, Halpert JR (2006) Drug Metab Dispos 34:1958–1965 61. Joyce MG, Girvan HM, Munro AW, Leys D (2004) J Biol Chem 279:23287–23293 62. Girvan HM, Marshall KR, Lawson RJ, Leys D, Joyce MG, Clarkson J, Smith WE, Cheesman MR, Munro AW (2004) J Biol Chem 279:23274–23286

J Biol Inorg Chem 63. Imai Y, Matsunaga I, Kusunose E, Ichihara K (2000) J Biochem 128:189–194 64. Matsunaga I, Yamada M, Kusunose E, Miki T, Ichihara K (1998) J Biochem 124:105–110 65. Matsunaga I, Sumimoto T, Kusunose E, Ichihara K (1998) Lipids 33:1213–1216 66. Matsunaga I, Yokotani N, Gotoh O, Kusunose E, Yamada M, Ichihara K (1997) J Biol Chem 272:23592–23596 67. Matsunaga I, Yamada M, Kusunose E, Nishiuchi Y, Yano I, Ichihara K (1996) FEBS Lett 386:252–254 68. Matsunaga I, Shiro Y (2004) Curr Opin Chem Biol 8:127–132 69. Lee DS, Yamada A, Sugimoto H, Matsunaga I, Ogura H, Ichihara K, Adachi S, Park SY, Shiro Y (2003) J Biol Chem 278:9761–9767 70. Matsunaga I, Yamada A, Lee DS, Obayashi E, Fujiwara N, Kobayashi K, Ogura H, Shiro Y (2002) Biochemistry 41:1886–1892 71. Matsunaga I, Sumimoto T, Ayata M, Ogura H (2002) FEBS Lett 528:90–94 72. Lee D-S, Yamada A, Matsunaga I, Ichihara K, Adachi S-i, Park S-Y, Shiro Y (2002) Acta Crystallogr Sect D 58:687–689 73. Matsunaga I, Ueda A, Fujiwara N, Sumimoto T, Ichihara K (1999) Lipids 34:841–846 74. Girhard M, Schuster S, Dietrich M, Durre P, Urlacher VB (2007) Biochem Biophys Res Commun 362:114–119 75. Girhard M, Kunigk E, Tihovsky S, Shumyantseva VV, Urlacher VB (2013) Biotechnol Appl Biochem 60:111–118 76. Ipe BI, Niemeyer CM (2006) Angew Chem Int Ed 45:504–507 77. Gandubert VJ, Torres E, Niemeyer CM (2008) J Mater Chem 18:3824–3830 78. Rajendran V, Konig A, Rabe KS, Niemeyer CM (2010) Small 6:2035–2040

79. Matsunaga I, Kusunose E, Yano I, Ichihara K (1994) Biochem Biophys Res Commun 201:1554–1560 80. Fujishiro T, Shoji O, Nagano S, Sugimoto H, Shiro Y, Watanabe Y (2011) J Biol Chem 286:29941–29950 81. Shoji O, Fujishiro T, Nakajima H, Kim M, Nagano S, Shiro Y, Watanabe Y (2007) Angew Chem Int Ed 46:3656–3659 82. Shoji O, Fujishiro T, Nagano S, Tanaka S, Hirose T, Shiro Y, Watanabe Y (2010) J Biol Inorg Chem 15:1331–1339 83. Fujishiro T, Shoji O, Kawakami N, Watanabe T, Sugimoto H, Shiro Y, Watanabe Y (2012) Chem Asian J 7:2286–2293 84. Fujishiro T, Shoji O, Watanabe Y (2010) Tetrahedron Lett 52:395–397 85. Shoji O, Wiese C, Fujishiro T, Shirataki C, Wu¨nsch B, Watanabe Y (2010) J Biol Inorg Chem 15:1109–1115 86. Kawakami N, Shoji O, Watanabe Y (2011) Angew Chem Int Ed 50:5315–5318 87. Kawakami N, Shoji O, Watanabe Y (2013) Chem Sci 4:2344–2348 88. Shoji O, Kunimatsu T, Kawakami N, Watanabe Y (2013) Angew Chem Int Ed 52:6606–6610 89. Poulos TL, Finzel BC, Howard AJ (1987) J Mol Biol 195:687–700 90. Yano JK, Koo LS, Schuller DJ, Li HY, de Montellano PRO, Poulos TL (2000) J Biol Chem 275:31086–31092 91. Yano JK, Wester MR, Schoch GA, Griffin KJ, Stout CD, Johnson EF (2004) J Biol Chem 279:38091–38094 92. Yano JK, Blasco F, Li HY, Schmid RD, Henne A, Poulos TL (2003) J Biol Chem 278:608–616

123

Peroxygenase reactions catalyzed by cytochromes P450.

Cytochromes P450 (P450s) catalyze monooxygenation of a wide range of less reactive organic molecules under mild conditions. By contrast with the gener...
3MB Sizes 3 Downloads 0 Views