Takayoshi
Awakawa
*ab,
Takahiro
Mori
abc,
Richiro
Ushimaru
abd and
Ikuro
Abe
*ab
aGraduate School of Pharmaceutical Sciences, the University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan. E-mail: awakawa@mol.f.u-tokyo.ac.jp; abei@mol.f.u-tokyo.ac.jp
bCollaborative Research Institute for Innovative Microbiology, the University of Tokyo, Yayoi 1-1-1, Bunkyo-ku, Tokyo 113-8657, Japan
cPRESTO, Japan Science and Technology Agency, Kawaguchi, Saitama, Japan
dACT-X, Japan Science and Technology Agency, Kawaguchi, Saitama, Japan
First published on 1st June 2022
Non-heme iron- and α-ketoglutarate-dependent oxygenases (αKG OXs) are key enzymes that play a major role in diversifying the structure of fungal meroterpenoids. They activate a specific C–H bond of the substrate to first generate radical species, which is usually followed by oxygen rebound to produce cannonical hydroxylated products. However, in some cases remarkable chemistry induces dramatic structural changes in the molecular scaffolds, depending on the stereoelectronic characters of the substrate/intermediates and the resulting conformational changes/movements of the active site of the enzyme. Their molecular bases have been extensively investigated by crystallographic structural analyses and structure-based mutagenesis, which revealed intimate structural details of the enzyme reactions. This information facilitates the manipulation of the enzyme reactions to create unnatural, novel molecules for drug discovery. This review summarizes recent progress in the structure-based engineering of αKG OX enzymes, involved in the biosynthesis of polyketide-derived fungal meroterpenoids. The literature published from 2016 through February 2022 is reviewed.
The biosynthesis of polyketide-derived meroterpenoids is achieved by polyketide synthases (PKSs), prenyltransferases (PTs), flavin-dependent monooxygenases (FMOs), terpene cyclases (CYCs),6 and tailoring enzymes, including isomerases, transferases, and oxidases, such as cytochrome P450 oxygenases (P450s)7 or non-heme iron- and α-ketoglutarate-dependent oxygenases (αKG OXs).2,8–11 PKSs construct polyketide cores from acetate-derived building blocks, PTs install prenyl side-chains, FMOs oxidize the olefin of the prenyl group to form epoxides, and CYC catalyzes protonation-initiated cyclization of the terpene moiety, and a tailoring enzyme decorates the resultant scaffold to generate complex meroterpenoid structures. Among the tailoring enzymes, αKG OXs often play a crucial role in amplifying the structural diversity and complexity of the molecules.12–17 They employ mononuclear non-heme Fe(II) and the co-substrate αKG, and generate a highly reactive ferryl–oxo intermediate (Fe(IV)O) with the concomitant oxidative decarboxylation of αKG to succinate.18,19 The Fe(IV)O first abstracts a hydrogen atom from a specific unactivated C–H bond in the substrate (Fig. 1). The generated intermediate radical species facilitate a wide range of reactions, including hydroxylation, desaturation, epoxidation, C–X bond formation, and C–C bond reconstruction.2,3,5,8–10,12–21 As a result, diverse meroterpenoid structures are produced, as exemplified by the biosyntheses of farnesyl-3,5-dimethylorsellinic acid (DMOA) 1 derived terretonin 2, austinol 3, emeridone F 4, anditomin 5, paraherquonin 6, and novofumigatonin 7 (Fig. 2).2 The structural basis of the αKG OX enzyme reactions has recently been extensively studied by X-ray crystallography and site-directed mutagenesis experiments. In this review, recent progress in the structure-based engineering of αKG OX enzymes to create unnatural novel molecules for drug discovery is summarized.
The comparison of the AusE and PrhA structures revealed that V150, A232, and M241 in PrhA are replaced with L150, S232, and V241 in AusE (Fig. 4c). To evaluate the effects of these substitutions, these three amino acid pairs were replaced with their counterpart residues. Remarkably, the AusE S232A variant produced 12 and 11 in comparable amounts, and AusE L150V/S232A produced 12 as a major product. As anticipated, the PrhA V150L/A232S variant produced 11 as a major product. The kcat and Km values of AusE L150V/S232A and PrhA V150L/A232S were very similar to those of PrhA and AusE, respectively, indicating that they efficiently catalyze the non-native reactions. Thus, the structure-based engineering successfully achieved functional interconversion of the two enzymes. The crystal structures of PrhA-V150L/A232S and PrhA-V150L/A232S/M241V complexed with each substrate revealed that the position of the D-ring of the substrate does not change, because it is tightly fixed by loop A and hairpin B, while significant conformational changes shift the orientation of the A/B rings toward the iron center (Fig. 4d). The wild-type PrhA abstracts H-5 of 8, leading to the production of 10, while PrhA V150L/A232S abstracts H-2 to produce 9, as in the case of wild-type AusE (Fig. 3). Based on the altered regioselectivities, it is likely that the A-ring of 8 undergoes a boat-to-chair flip and the C-2 of the substrate becomes closer to the iron in PrhA V150L/A232S than that in the wild-type enzyme (4.2 Å vs. 4.4 Å). This example shows how the enzyme active site dictates the conformation of the terpene moiety of the substrate, leading to the two different regioselectivities of the reactions. In PrhA V150L/A232S complexed with 9, the distance between H-5 and Fe(II) is the shortest, which is consistent with the proposed mechanism of the reaction initiation by the abstraction of H-5. After the formation of 9, the reconstruction of the spirocycle ring system is proposed as follows. The C-5 radical first reacts with the double bond between C-1 and C-2, resulting in the cyclopropylcarbinyl radical intermediate, and the C-2 radical then reacts with an electron of the bond between C-1 and C-10, followed by the dehydrogenation of H-9 and the formation of the C-9/C-10 double bond, to produce 11 (Fig. 5a).
Interestingly, the PrhA V150L/A232S variant oxidizes C-13 on 11 to produce unnatural novel products, 13 and 14, and PrhA V150L/A232S/M241V further oxidizes 13 into 16, and 14 into 15 (Fig. 5b). The crystal structure of PrhA V150L/A232A with 11 revealed that the C-10/C-13 bond is rotated clockwise by 30°, thus reducing the distance from C-13 to Fe(II).
It is remarkable that the double amino acid substitutions V150L and A232S in PrhA increase the number of oxidations from two (8 to 11) to five times (8 to 16). Because the structure of the D-ring of the substrate does not change during the reaction, the enzyme can recognize and accommodate the oxidized products 9, 11, 13, and 14, even though its A/B ring structures are altered significantly. The prediction and evaluation of the conformational changes of substrates in the active site of the enzyme are crucial to rationally design the structures of molecules, as illustrated in the in silico-designed targeted protein engineering of class I terpene cyclases.25,26
In the isomerization reaction, which starts with the abstraction of H-12 of 18 which is bound to an enzyme via O-1, -3, -4, and O-5 like 17, the two distances from H-12 to O-2/αKG (3.5 Å) and to w4 (4.9 Å) support that the structure is reasonable for H-12 abstraction (Fig. 6 and 7c). To further investigate the dynamics of the isomerization, the DFT calculation was performed.28 After the H-12 abstraction of 18 to generate 18a, four reaction steps can be proposed (Fig. 8): (i) C8–O2 bond cleavage (18a to 18b), (ii) C12–C5′ bond formation (18b to 18c), (iii) C8–C2′ bond formation (18c to 18d), and (iv) quench of the C-7′ radical by ascorbic acid (18d to 19). In accordance with the prediction that the energetically stable tertiary carbon radical 18b is preferred, the calculated energy for the C–O bond cleavage is only 6.6 kcal mol−1. In another pathway, anti-Baldwin 5-end-trig cyclization proceeds to yield the α-oxy carbon radical intermediate 18b′ with an activation energy of 23.4 kcal mol−1, indicating that the pathway to 18b is more likely than this pathway. During the conversion from 18c to 18d, several conformational changes with minute activation barriers are expected to occur, and the C-8 radical approaches to the Michael acceptor carbon center C-2′. The radical at C-7′ in 18d is finally quenched by a reducing agent, resulting in the generation of 19.
To examine the roles of the amino acids in the active site of AndA, N121, R239, and Y272′ were substituted with alanine or apolar residues of similar bulkiness. The AndA N121A variant exhibited reduced activity, and the AndA R239A, R239M, R239V, and Y272A variants only yielded 18 and failed to produce 19. Because R239 and Y272′ interact with the lid-like loop region, these substitutions likely interrupted the sequential conformational changes required for the isomerization reaction. These observations indicated the importance of the lid region to maintain the reaction.
The combination of structural and computational studies nicely illustrates the mechanism of the isomerization enzyme reaction to produce the unique bicyclo[2.2.2]octane system. Like AusE and PrhA,24 AndA recognizes the A/B rings of 17 and 18 by hydrogen bonding with N121 and hydrophobic interactions, but not tight interactions as seen in those with the D ring, thus allowing the dynamic conformational changes of the intermediates during the enzyme reaction. Computational analyses of the enzyme-bound reaction intermediates will provide important information regarding how the enzyme manages the conformational changes of the substrate/intermediates. For example, the underlying reasons for the requirement of ascorbate in the in vitro enzyme reaction, the native reductant, and the mechanism of the reduction from 18d to 19 remain unknown. Similar reductive radical quenching is also observed in several unique αKG OX enzyme reactions, including FtmOx1 (endoperoxidase),30–32 CarC (epimerase),33 and SnoN (epimerase),34 but the mechanisms of their reductions remain obscure, even though they are essential for the functional conversions of αKG OXs from oxygenases to isomerases.
Notably, the distance from C-7′ of 20 to Fe(II) (4.2 Å) is shorter than the distance from C-13 (6.5 Å) in the complex structure (Fig. 10c), suggesting that this binding model is an artifact or represents the reaction of a C-7′ oxidation reaction to produce the derailment product 22 (Fig. 11). The docking simulation of the NvfI-αKG binary complex with 20 indicated that the enzyme accommodates 20 by adjusting its conformation so that C-13 is located closer to Fe(II) than C-7′. Furthermore, the comparison of the NvfI-αKG and NvfI-αKG/20 complex structures suggested that the movement of the loop around E208 increases the volume of the enzyme active site around the A-ring of 20, and the radical formation at C-13 and the following conformational changes move the substrate location deeper inside the tunnel (Fig. 10b).
Fig. 11 Proposed reaction mechanism of NvfI to synthesize 21 from 20. The structures of the derailment products 22 and the product 23 from the H138A variant reaction are also depicted. |
The substitution of the F127 on loop A and W199 on loop B lead to a loss of activity, which shows that these residues are essential for the endoperoxide forming activity. Their roles were proposed as gate-keepers that retain the correct conformation of 20. In other endoperoxide-forming enzymes; i.e., heme-dependent prostaglandin H synthase (PGHS, also known as cyclooxygenase, COX)39 and αKG OX fumitremorgin B endoperoxidase (FtmOx1),30–32 the active site tyrosine residue close to the Fe(IV)O was proposed to play a critical role to relay a radical from the reaction intermediate (Fig. 12). However, the substitution of the Y residues in the NvfI active site to F did not affect the reactivity. These data indicated that NvfI employs a distinct mechanism from those of PGHS and FtmOx1. Interestingly, NvfI H138A reduced the endoperoxide-forming activity by 20%, and formed newly generated 23 (Fig. 11), which indicated that the hydrogen bonding between H138 and the C-3 ester group is important to adjust the location of the substrate (Fig. 10c). Consequently, the reaction mechanism from 20 to 21 was proposed, as shown in Fig. 11. The Fe(IV)O species abstracts the hydrogen atom from C-13 to generate the primary radical A, which reacts with O2 to form the peroxide radical B. The generated radical B reacts with C-2′ to yield radical C, which then reacts with the Fe(III)-hydroxyl species to produce 21 (path a). It is also possible that electron transfer from the radical C to the ferric iron species generates the cationic intermediate D, and the successive attack by water yields 21 (path b). In path a, the oxygen ligands in the Fe(IV) and Fe(III) states should be exchanged with the solvent water during the catalysis as reported in the reaction of other αKG OXs.40 NvfI thus consumes only one molecule of αKG to produce one molecule of 21, as hypothesized in the reaction mechanism described above. The absolute configuration of OH-3′ is restricted to the R-configuration, indicating that the enzyme controls the orientation of the hydroxyl-rebound or the addition of water.
The structural analysis of NvfI revealed how the monomeric αKG OX accepts the substrate with the conformational changes of loop I and II. More importantly, this study proposed a new mechanism of the rare enzymatic endoperoxide formation reaction without the help of the tyrosine residue for the radical relay, although the possibility that other residues may work in the relay cannot be excluded. NvfI may suppress the direct oxygen rebound between the radical A and Fe(III)–OH by facilitating the large conformational changes of the radical A. The prolonged lifetime of the radical intermediates may lead to the reaction with molecular oxygen, although the mechanism of molecular oxygen retention in the enzyme remains enigmatic. This hypothesis of oxygen rebound prevention could be applicable to other non-hydroxylating αKG OX enzyme reactions.12–16 This study provided useful information for future enzyme engineering to introduce unique chemical scaffolds into the products.
Fig. 14 Reactions of monomer TlxJ and heterodimer TlxI-J to synthesize talaromyolides 24a, 24b, 24d, 24e, and 25, and unnatural analogs 28 and 29. |
The reaction pathway from 27 to 24a, 24d, and 25 was proposed, as described below. First, TlxI-J abstracts H-5 to give the C-5a hydroxylated trans-drimane intermediate (Fig. 14a). The deprotonation of the OH-5a leads to the retro-aldol reaction which cleaves the C4–C5 bond, as proposed in yaminterritrem B biosynthesis,44 and the resultant intermediate undergoes successive ketal formation to produce 24d. The hydroxylation at C-4 of the retro-aldol reaction product generates 24a. It is intriguing that 25 is derived from the cis-drimane intermediate, while its yield is lower than those of the other products. After radical formation at C-5, the trans-drimane is likely transformed into cis-drimane before the oxygen rebound.
Interestingly, the unnatural meroterpenoids 27 and 28 were also observed in the TlxI-J reaction (Fig. 14b). They are expected to be generated via repeat demethylations of C-14 and C-15 through sequential oxygenation from the methyl group to the carboxylic acid followed by decarboxylation, as in cholesterol biosynthesis45 and terpenoid secondary metabolite biosynthesis.46,47 The successive C-3 demethylation possibly proceeds via oxidative decarboxylation, as in αKG18,19 and 4-hydroxyphenylpyruvate decarboxylation.48,49
The gel filtration analyses of TlxI-J and TlxA-C revealed that they form unprecedented αKG OX enzyme heterodimers.12,13 The amounts of TlxJ and TlxA purified from an Escherichia coli expression system are low when they are solely expressed, but they are efficiently produced as a monomer or heterodimer when co-expressed with TlxI or TlxC. Together, these data indicate that TlxI and TlxC serve as chaperone-like proteins to increase the stability of TlxJ and TlxA, respectively. Interestingly, TlxC cannot be substituted for TlxI, and vice versa, indicating the rigid specificity for the heterodimer formation.
To investigate the molecular basis of the unprecedented αKG OX heterodimer, the X-ray crystal structure of TlxI-J complexed with Fe/N-oxalylglycine (NOG)50 was analyzed.43 TlxJ in the heterodimer adopts the conserved DSBH fold, while TlxI does not possess a complete DSBH fold. In fact, loop A′ in TlxI is much shorter than the corresponding region of TlxJ (Fig. 15a). These observations suggest that the reaction chamber cannot be constructed in the TlxI monomer. Thus, in analogy with the other homodimeric αKG OXs such as PrhA and AndA,11,16 a funnel-like substrate binding site is expected to be constructed with the DSBH fold (TlxJ) as a chamber, and loop A (TlxJ) and loop B′ (TlxI) as a lid. The metal binding site (H136, D138, and H213) and αKG binding site (R224, N133, and T173) are conserved in the TlxJ structure, but the residue corresponding to R224 of TlxJ is substituted with S211 in TlxI (Fig. 15b and c), indicating that TlxI cannot bind αKG.
The substitution of the amino acid residues oriented toward the active site of TlxJ, including R120 and L239 located on the DSBH core and I259′ located on loop B′, significantly reduced the reactivity, while these variants maintained the heterodimer structure as confirmed by gel filtration, suggesting that these residues are required for substrate binding. Interestingly, TlxJL239A-I newly accumulated the C-1 hydroxylated 27, indicating that the substitutions in the hydrophobic reaction chamber are effective to alter the reaction profile, as observed in PrhA/AusE and AndA.24,28
To investigate the differences between the heterodimer and homodimer structures, the amino acid residues located at the interface of TlxI-J were substituted. There are several amino acid pairs responsible for the hydrogen bonding, including R63′ and S276, T224′ and F236 (main chain), R135′ and D152, and W262′ and P238 (Fig. 16a). As expected, the TlxI R63′A, T224′A, R135′A, and W262′A variants no longer form a heterodimer with TlxJ, as evidenced by the pull-down assays, indicating that they are important for the interaction between monomers. Comparison of the structure of TlxI-J with those of PrhA and AndA demonstrated that the method of heterodimer formation is quite different from that of homodimer formation (Fig. 16b and c).
The reason why TlxJ requires TlxI to form a functional heterodimer remains to be elucidated. Since TlxJ exists as a monomer when it is produced in E. coli, it is likely that TlxJ cannot form a homodimer by itself. The TlxJ monomer can oxidize 26, although this reaction is much slower than that of TlxI-J. These results indicate that the TlxJ monomer can also form a premature reaction cavity, but it requires TlxI to form a reaction cavity that can accept 27. The CRISPR-based gene inactivation of TlxI prevented the production of talaromyolides and accumulated 27, implying that TlxI-J also forms a heterodimer in vivo. This study presented the characterization of the first heterodimeric αKG OXs. The interaction of each monomer in the heterodimer is quite different from that in the homodimer. Given that TlxI-J can catalyze a high number of oxygenations (possibly 11 times to synthesize 28 and 29), the heterodimer structure might decrease the rigidity of substrate recognition. If two αKG OXs are artificially complexed as a heterodimer, a multifunctional heterodimer like TlxI-J could be created. The simulation showing the creation of the cis-drimane intermediate in the enzyme after the C-5 radical intermediate would also be intriguing.
Fig. 17 Synthetic reactions from andiconin D 31. The syntheses of 34 and 35 by SptF WT, the synthesis of 46 by SptF I63A, and the synthesis of 47 by SptF F133A. |
Fig. 18 Synthesis of unnatural products by SptF from andilesin C 36 (a), preandiloid B 17 (b), terretonin J 39 and A 40 (c), and terretonin C 41 and terretonin 2 (d). |
Inspired by its catalytic versatility, the substrate scope of SptF was further investigated using six fungal meroterpenoids as unnatural substrates, including preandiloid B 17, preandiloid C 18, terretonin J 39, terretonin A 40, terretonin C 41, and terretonin 2 (Fig. 18b–d). Except for 18, all of the tested substrates were accepted by SptF, indicating that SptF has relaxed substrate specificity. SptF catalyzed the C-1 hydroxylation of 17 (Fig. 18b), and reacted with 39 and 40 to generate 42 and 43 possessing the 5/3/6/6/6 scaffold, and the reaction with 41 and 2 to yield 44 and 45 possessing the 5/3/5/5/6/6 scaffold, respectively (Fig. 18c and d). The consumption rates of 40 and 2 by SptF were lower than 10%, indicating that the bulky functional group at the D-ring hampers the reaction. SptF also oxidizes various kinds of steroids, including androsterone, testosterone, and progesterone, with moderate catalytic efficiencies (34–53%).
To determine the molecular basis of the exceptional promiscuity of SptF,51,52 its X-ray crystallographic structure was analyzed. SptF alone exhibits the conserved DSBH fold (Fig. 19a). The complex structures with 31 and 36 showed that these two substrates are located at almost identical positions, indicating that the A-ring of the substrates does not alter the binding manner. The distances from the iron to the initial reaction site C-11 of 31 and 36 are 4.2 and 4.3 Å, respectively (Fig. 19b and c), and are reasonable distances for the hydrogen abstraction. The hydrophobic surface, consisting of I63, F133, and I231, mainly supports the substrate in the enzyme cavity. The lid-like region interacts with the A-rings of 31 and 36via hydrogen bonding with only N65, while their E rings are supported by S114, the main chain of L199, and T148. The loose interaction with the lid-like loop explains the broad substrate specificity toward structurally diverse compounds. Interestingly, the unnatural substrate terretonin C 41 was also accepted by SptF in a different ligand binding mode, with the direct hydrogen bond interaction between the D-ring and N150 and the indirect hydrogen bond network via water molecules supported by T148, D130, and L199 (Fig. 19d). Remarkably, the lid-like loop region was not observed in the complex structure with 41, while the conformation of the other active site regions are almost identical. In this structure, the substrate is mainly supported by hydrophobic interactions with F133 and I231, which are close to the C-8, C-10, and C-13 methyl groups. The C-6 enol of 41 is closest to the iron (3.7 Å), suggesting that the formation of the cyclopropane rings of 42, 43, 44, and 45 are triggered by the hydrogen bond abstraction from this enol.
To investigate the role of I63, F133, and I231, each was substituted with alanine. As a result, SptF I63A prevented the production of 34 and 35, but produced the novel product 46 (Fig. 17). The structure of 46 was determined to be an E-ring opened dicarboxylic acid, which is generated through two rounds of oxidation at the C-1′ of 31. In contrast, the SptF I63A and F133A variants almost completely lost their ability to produce 34 and 35, and accumulated 32 and 33, respectively. The SptF F133A variant also converted 31 into the C-11 hydroxylated 47 (Fig. 17). These data indicated the crucial role of hydrophobic interactions for substrate binding and the potential to engineer the enzyme reaction by site-directed substitutions of the hydrophobic residues.
The importance of the hydrophilic residues, including N65, S114, T148, and N150, was also investigated in detail. N65A significantly reduced the activity by 80% and accumulated 33. The SptF T148A, S114A, and N150A variants decreased the production efficiencies of 34 and 35 by 7, 17, and 48%, and changed the production ratios of 34:35 to 7:3 for the S114A and T148S variants and 1:9 for the N150A variant, as compared with 4:6 for SptF WT. These observations indicate that the three hydrophilic residues are important to define the orientation of 33 within the reaction chamber. When N65 was substituted with threonine, which is the corresponding residue in the closely related AndF, the variant produced 35 as a major product from 31, with a smaller amount of 34. In addition to the hydrophobic and hydrophilic interactions, the functions of I63 and N65 in the lid-like loop region were investigated. Remarkably, the SptF I63A and N65A variants accepted 41 with similar efficiencies to the wild-type enzyme, while the reactivity on 31 was 14 times lower, suggesting the different substrate binding modes between the two substrates. To evaluate the importance of the flexible lid-like loop region, the loop was truncated by 6 residues (Δ6, H61-V66), 9 residues (Δ9, K58-V66), 16 residues (Δ16, N54-K69), and 19 residues (W53-K71). The Δ6, Δ9, and Δ16 truncated SptFs were reacted with 31, 32, and 33. Interestingly, these variants lost the activities of 31 and 32, and generated several minor products, but no longer produced 33–35. In contrast, these variants showed 26–52% activity toward 33, and generated small amounts of 34 and 35. This observation suggested the importance of the lid-like region to control the reaction. While the unnatural meroterpenoid substrates were accepted by the variants to a lower extent, the steroidal substrates were no longer accepted.
Based on the investigations of the structure–function relationship, the reaction mechanisms for the SptF-catalyzed oxidation reactions were proposed (Fig. 20). Hydrogen abstraction at C-11 of 31 by the Fe(IV)O species firstly generates a radical, which undergoes the cleavage of the C-4′–C-3′ bond and the construction of the C-4′–C-11 bond. The following H-9′ abstraction initiates the formation of the exo-methylene 33, and the successive epoxidation produces 32. In pathway a, the second epoxidation of C-1–C-2 and the H-9 abstraction, followed by the rebound of the hydroxyl group, produced 34 as a product. In pathway b, H-7′ is abstracted to induce the cleavage of the C-8–C-2′ bond and generate the double bond between C-7 and C-8. The hydroxyl rebound between the radical at C-2′ and the Fe(III)-hydroxyl species yields IM2. The 2′-OH reacts with the 4′-carbonyl and the subsequent electron migration results in the protonation of C-7′ to generate IM3. The H-7′ in IM3 is abstracted by the Fe(IV)O species in the next round of oxidation, and the generated allylic radical receives a hydroxyl group at C-5′ from the Fe(III)-OH species to produce 35.
SptF accepts a remarkably broad range of meroterpenoid and steroid substrates. The solved complex structures revealed the different substrate binding mode for each substrate. Compound 41 can bind to the enzyme in the inverse manner, as compared with those for 31 and 36. SptF also utilizes the loop region as a lid for the active site cavity, but it does not appear to bind the substrate tightly. The relaxed binding mode of the loop region and the hydrophobic reaction chamber would allow the promiscuous substrate binding.
To obtain more versatile and useful αKG OX enzymes, further search for new enzymes that catalyze novel reactions or exhibit unique substrate specificities is still needed. The genome mining method would be helpful in searching for such enzymes.53,54 In the study of the fungal αKG OX AsqJ in the quinolone alkaloid biosynthesis, various research techniques have been applied to elucidate its stepwise desaturation and epoxidation reactions (e.g. X-ray crystallography, pre-steady-state enzyme kinetics, Mössbauer spectroscopy, assays with unnatural substrates, DFT calculations, and molecular dynamic simulations).29,55–60 Likewise, studies on structural biology, bioorganic chemistry, and computational chemistry will provide important information for enzyme engineering of αKG OXs. Future development of computational calculations of catalytic reactions is expected to enable highly accurate simulations of enzyme dynamics, which will provide useful insight for more rigorous control of enzyme reactions and the creation of biocatalysts by de novo design. In addition, a recent engineering study to convert an αKG-dependent amino acid hydroxylase into a halogenase suggested some strategies to evolve the αKG OXs towards enzymes with different catalytic functions.61 During the writing of this manuscript, two studies based on the saturated random mutagenesis to evolve αKG OXs in meroterpenoid biosynthesis were published.62,63 Such random mutagenesis-based studies complement the rational enzyme engineering. Finally, the engineered enzymes can be utilized in combinatorial biosyntheses in heterologous expression hosts such as Aspergillus oryzae, to produce pharmaceutically useful molecules, as demonstrated by the recent successful productions of unnatural novel diterpene fungal meroterpenoids.64–66 Therefore, the discovery, engineering, and reconstitution of fungal meroterpenoid oxidation pathway will continue to provide important insight into natural product chemistry and enzymology, and will lead to the production of beneficial drugs.
This journal is © The Royal Society of Chemistry 2023 |