Skip to main content

Recent progress in directed evolution of stereoselective monoamine oxidases

Abstract

Monoamine oxidases (MAOs) use molecular dioxygen as oxidant to catalyze the oxidation of amines to imines. This type of enzyme can be employed for the synthesis of primary, secondary, and tertiary amines by an appropriate deracemization protocol. Consequently, MAOs are an attractive class of enzymes in biocatalysis. However, they also have limitations in enzyme-catalyzed processes due to the often-observed narrow substrate scope, low activity, or poor/wrong stereoselectivity. Therefore, directed evolution was introduced to eliminate these obstacles, which is the subject of this review. The main focus is on recent efforts concerning the directed evolution of four MAOs: monoamine oxidase (MAO-N), cyclohexylamine oxidase (CHAO), D-amino acid oxidase (pkDAO), and 6-hydroxy-D-nicotine oxidase (6-HDNO).

Introduction

When performing catalytic stereoselective transformations, organic chemists can choose between chiral chemical catalysts (Noyori 2002; Sharpless 2002; Walsh and Kozlowski 2009; Zhou 2011; Blaser and Schmidt 2004; MacMillan 2008; List 2010; Allen and MacMillan 2012; Atodiresei et al. 2015; Wang and Tan 2018; Vetica et al. 2017) and enzymes (Drauz et al. 2012; Tao et al. 2009; Gotor et al. 2008; Li et al. 2018b; Reetz 2016a). The well-documented advantages of using enzymes are their exquisite regioselectivity and stereoselectivity at ambient conditions in many chemical reactions (Drauz et al. 2012; Tao et al. 2009; Gotor et al. 2008; Li et al. 2018b; Reetz 2016a). Thus, the use of enzymes as catalysts in synthetic organic chemistry has experienced rapid growth during the last 4 decades. However, the truly broad application of enzymes still suffered from several long-standing limitations for non-natural substrates, e.g., limited substrate scope, poor selectivity, insufficient stability, and sometimes substrate or product inhibition (Reetz 2016b; Ni et al. 2014; Sheldon and Pereira 2017). Nowadays, all of these long-standing limitations of enzymes can be generally addressed by directed evolution (Drauz et al. 2012; Tao et al. 2009; Gotor et al. 2008; Li et al. 2018b; Reetz 2016a).

Directed evolution involves repeated cycles of gene mutagenesis, expression, and screening of mutant enzyme libraries, simulating natural evolution (Reetz 2011; Quin and Schmidt-Dannert 2011; Brustad and Arnold 2011; Bommarius et al. 2011; Siloto and Weselake 2012; Bornscheuer et al. 2012; Porter et al. 2016; Zeymer and Hilvert 2018; Arnold 2018; Denard et al. 2015). When focusing on stereo- and/or regioselectivity, screening is the labor-intensive step (bottleneck of directed evolution) (Acevedo-Rocha et al. 2014; Reymond 2006). As the first step in each cycle, gene mutagenesis is crucial for the success of directed evolution, and therefore, considerable efforts have been invested in the exploration of gene mutagenesis techniques. The most commonly employed gene mutagenesis techniques are error-prone polymerase chain reaction (epPCR) (Leung 1989; Cadwell and Joyce 1994; Chen and Arnold 1993; Reetz et al. 1997), saturation mutagenesis (Estell et al. 1985; Kirsch and Joly 1998; Vandeyar et al. 1988; Zheng et al. 2004), and DNA shuffling (Stemmer 1994). The three mutagenesis techniques have been widely applied in directed evolution to address the limitations of enzymes. However, efficient approaches for delivering small and “smart” libraries needed in optimizing and inverting stereoselectivity of enzymes had to be explored. Along this line, Reetz et al. introduced Combinatorial Active Site Saturation Test (CAST) (Clouthier et al. 2006; Reetz et al. 2005, 2006a), according to which sites around the binding pocket are chosen for randomization. When the initial CAST libraries do not harbor fully optimized mutants, then iterative saturation mutagenesis (ISM) can be applied (Reetz and Carballeira 2007; Reetz et al. 2006b). The combination CAST/ISM has proven to be an efficient strategy in directed evolution, routinely applied by numerous groups (Acevedo-Rocha et al. 2018; Li et al. 2018a; Zhang et al. 2019a, b; Chen et al. 2018a, b).

Enantiomerically pure chiral amines are valuable synthetic intermediates for the preparation of pharmaceuticals; approximately one-third of the chiral pharmaceuticals on the market contain chiral amine functional groups (Blacker and Headley 2010; Harvey 2008). Traditionally, chiral amines have been obtained by resolution-based methods, for instance, crystallization of a diastereomer using a chiral acid to form a salt (Bálint et al. 2001) or kinetic resolution of a racemate with an enzyme (Guranda et al. 2001; Lee 1999; Van Langen et al. 2000; Skalden et al. 2016). Unfortunately, the maximal theoretical yield of a given enantiomer based on these methods is 50%, severely limiting the efficiency of such a kinetic resolution-mediated process. As a result, the development of efficient and widely applicable biocatalysts for the synthesis of chiral primary, secondary, and tertiary amines has received significant attention (Turner and Truppo 2010). For example, Bäckvall et al. developed a lipase-catalyzed process based on dynamic kinetic resolution (DKR) of primary amines, which requires a transition metal complex for the racemization step (Thalén et al. 2009). In 2002, Turner et al. reported a novel deracemization method for the preparation of optically active primary chiral amines, which involves the stereoinversion of one enantiomer to the other by repeated cycles of enantioselective enzymatic oxidation to the imine using an appropriate amine oxidase, followed by non-selective reduction to the racemic starting amines (Scheme 1) (Alexeeva et al. 2002). In the Turner‘s deracemization method, the key issue at the outset was to identify a highly enantioselective and active amine oxidase, which at that time was challenging. Turner et al. initiated directed evolution of MAO-N in conjunction with a colorimetric plate-based high-throughput screening assay (Alexeeva et al. 2002). After cycles of directed evolution based on random mutagenesis using a mutator strain, a “toolbox” of MAO-N variants was developed in the Turner group, which has been successfully applied to the deracemization of primary (Carr et al. 2003), secondary (Alexeeva et al. 2002), and tertiary amines (Dunsmore et al. 2006). The process was also applied to the synthesis of chiral heterocyclic amines such as substituted pyrrolidines (Köhler et al. 2010) and tetrahydro-isoquinolines (Ghislieri et al. 2013; Rowles et al. 2012). Prompted by these pioneering studies, more monoamine oxidases were identified and subjected to directed evolution for chiral amine synthesis; for instance, CHAO (Leisch et al. 2011; Li et al. 2014a), pkDAO (Yasukawa et al. 2014, 2018), and 6-HDNO (Heath et al. 2014).

Scheme 1
scheme 1

Process of the deracemization of chiral amines by employing recursive cycles of enantioselective oxidation using an amine oxidase coupled with a non-selective reducing agent

The present review summarizes recent efforts regarding the directed evolution of monoamine oxidases for the synthesis of chiral amines. It is not meant to be comprehensive. Rather, a select number of representative studies are featured and analyzed, illustrating the viability of directed evolution for manipulating the catalytic properties of monoamine oxidases.

Directed evolution of monoamine oxidase from Aspergillus niger (MAO-N)

The most frequently reported monoamine oxidase is MAO-N, which displays high activity towards aliphatic amines, e.g., amylamine and butylamine, low but measurable activity to benzylamine and α-methylbenzylamine (Carr et al. 2003). Turner et al. chose this enzyme as a viable starting point for improving both the catalytic activity and stereoselectivity by directed evolution. At the time of the initial study, the X-ray data were not available; therefore, the method of random mutagenesis based on E. coli XL1-Red mutator strain was employed for generating a library of MAO-N variants. Facing the huge library size of variants created by this gene mutagenesis technique, the initial priority was the development of an effective high-throughput screening assay based upon capture of the generated hydrogen peroxide by a peroxidase in the presence of 3,3′-diaminobenzidine, giving a dark pink, insoluble product (Fig. 1) (Carr et al. 2003). With this plate-based colorimetric assay in hand, a subset of the library (about 150,000 clones) was screened, finally identifying an improved mutant (Asn336Ser). Relative to wild-type (WT) enzyme, the activity and stereoselectivity of Asn336Ser variant increased 47-fold and 5.8-fold, respectively. Without further mutation, the Asn336Ser variant was tested for a panel of amine substrates with broad structural features. Variant Asn336Ser showed significantly wider substrate profile relative to WT MAO-N, particularly for primary amines, but also for a number of secondary amines. The new variant was used in the deracemization of rac-1-methyltetrahydroisoquinoline, producing high yield and ee value. To extend the chemoenzymatic deracemization method to encompass more types of amines, the MAO-N were then subjected to several rounds of directed evolution following a similar procedure as described above, identifying two property improved variants MAO-N-D3 (Asn336Ser/Met348Lys/Ile246Met) (Carr et al. 2005) and MAO-N-D5 (Asn336Ser/Met348Lys/Ile246Met/Thr384Asn/Asp385Ser) (Dunsmore et al. 2006). This D5 variant displayed good activity toward a wide range of tertiary amines (Fig. 2), particularly for pyrrolidine derivatives that are flanked by bulky aryl groups. Preparative scale deracemization of Rac-N-methyl-2-phenylpyrrolidine (3) was performed at 25 mM concentration using the D5 variant as oxidant, giving (R)-3 in 75% isolated yield and 99% ee within 24 h (Fig. 3) (Dunsmore et al. 2006). As an illustration of the potential of variant D5 in organic synthetic chemistry, it was also applied for the desymmetrization of the bicyclic pyrrolidine 11 (Fig. 4), a building block for the synthesis of hepatitis C virus protease inhibitor telaprevir (14) (Köhler et al. 2010). The reaction reached more than 98% conversion within 7 h, giving 77% isolated yield and 94% ee, and recrystallization of the trimer of 13 improved the ee to at least 98% (Köhler et al. 2010).

Fig. 1
figure 1

Colorimetric plate-based screening assay for amine oxidase activity by capture of the hydrogen peroxide produced using 3,3′-diaminobenzidine with peroxidase. a Colonies with D-α-methylbenzylamine as screening substrate. b Identical clones with the L-enantiomer as substrate

Fig. 2
figure 2

Structurally diverse amines were subjected to the Turner’s deracemization method using variant MAO-N-5

Fig. 3
figure 3

Deracemization of racemic N-methyl-2-phenylpyrrolidine 3 to (R)-N-methyl-2-phenylpyrrolidine by employing a sequence of enantioselective oxidations using variant MAO-N-5 coupled with the non-selective reducing agent NH3BH3

Fig. 4
figure 4

Conversion of the symmetrical bicyclic pyrrolidine 11 to an l-proline analog 13 via initial MAO-N catalyzed desymmetrization followed by diastereoselective addition of cyanide and subsequent hydrolysis

Although impressive advances have been achieved in the directed evolution of MAO-N, the present “toolbox” of MAO-N did not accept bulky amines such as 4-chlorobenzhydrylamine (15) and 1-phenyltetrahydroisoquinoline (16) (Fig. 5), motifs that are present in the commercially available drugs Levocetirizine (17) and Solifenacin (18) (Fig. 5), respectively (Hermanns et al. 2002; Nguyen et al. 2011). To improve the “toolbox” of MAO-N further to catalyze these bulky amines, the D5 variant of MAO-N was subjected to further directed evolution. Based on the results of docking α-methylbenzylamine into the MAO-N D5 active center, it became clear that increasing the volume of binding pocket would allow the accommodation of the bulky substrates with two aryl substituents. To reach this goal, two residues (Ala429 and Trp430) were targeted for saturation mutagenesis (Ghislieri et al. 2013). By screening the two saturation mutagenesis libraries, variant MAO-N D10 with a new single amino acid substitution Trp430Gly was identified, which was active toward amine 15 for the first time, although only to a moderate extent. In an attempt to improve the activity further, attention was then turned to the active site channel, (previous) mutations in this region resulting in a panel of mutants MAO-N D9 (A-D) (Table 1) which showed significantly enhanced activity to the bulky amine crispine A relative to MAO-N D5. A series of mutants MAO-N D11 (A-D) (Table 1) were created by the combination of essential mutations in MAO-N D9 (A-D) with MAO-N D10, and their activity and stereoselectivity towards amine 15 were measured by monitoring the conversion of the corresponding enantiomer of 15 using HPLC, which proved MAO-N D11C to be the best mutant. Consistent with the activity tests, the computed volume of the binding pocket in D5, D10, and D11C clearly increase from 140 Å3 (D5) to 195 Å3 (D10), and further to 464 Å3 (D11C). The D11C variant was applied for the deracemization of rac-15 on a 0.5 g scale, and (R)-15 with 97% ee was obtained within 48 h with 45% isolated yield. The compromised yield was shown to be due to partial hydrolysis of the unstable intermediate imine. A similar procedure was also applied for the deracemization of rac-16 on a 1 g scale, 98% ee, and 90% isolated yield being obtained within 24 h (Ghislieri et al. 2013).

Table 1 MAO-N variants obtained by directed evolution and respective oxidation rate towards (S)-15 and (R)-15
Fig. 5
figure 5

Substituted benzhydrylamines 15 and 16 as important structural motifs in commercial pharmaceuticals Levocetirizine 17 and Solifenacin 18, respectively

Relevant to this work is a study of the Reetz group involving the simultaneous manipulation of activity and stereoselectivity of MAO-N by focusing simultaneously on residues lining the entrance tunnel and the binding pocket (Li et al. 2017). In this case, 23 residues located in the entrance tunnel and the binding pocket of MAO-N were chosen and split into six groups for saturation mutagenesis using NDT codon degeneracy (Phe, Leu, Ile, Val, Tyr, His, Asn, Asp, Cys, Arg, Ser, and Gly) as building blocks (Fig. 6). The corresponding saturation mutagenesis libraries were then screened toward three model amines for identifying positive hits and critical (“hot”) positions. Subsequently, further ISM was carried out based on the information obtained in the first round of directed evolution, leading to high active and stereoselective variants LG-I-D11 (W230R/W430C/C214L) and LG-J-B4 (W230I/T354S/W430R/M242R/Y365 V). The mutations induced in these variants lead to reversal of enantioselectivity of Turner-type deracemization in the synthesis of amines (S)- 1-methyl-1,2,3,4-tetrahydroisoquinoline (S-19), (S)-1-phenyl-1,2,3,4-tetrahydroisoquinoline (S-20), (S)-1-ethyl-1,2,3,4-tetrahydroisoquinoline (S-21), and (S)-1-isopropyl-1,2,3,4-tetrahydroisoquinoline (S-22) (Fig. 7). This is a significant result, because only the corresponding (R)-19 were previously accessible using the earlier MAO-N variants. Molecular Docking and Molecular Dynamics Simulations indicate that it is the increased hydrophobicity of the entrance tunnel acting in concert with the altered shape of the binding pocket that results in the altered catalytic profile.

Fig. 6
figure 6

MAO-N residues chosen for saturation mutagenesis, marked in the homology model, which was built using the crystal structure of MAO-N-D3 (PDB: 2VVL). A: Active site mutation sites (yellow), selected on the basis of induced fit docking of amine 19 (green). B: Residues surrounding the substrate access tunnel (red) likewise chosen for mutagenesis (shown in green)

Fig. 7
figure 7

Deracemization of racemates 19, 20, 21, and 22 by employing a sequence of enantioselective oxidations with MAO-N mutants and non-selective reduction with NH3BH3

In another significant case, directed evolution for simultaneously improving activity and thermostability of MAO-N was achieved in a Merck/Codexis collaboration, which developed a chemoenzymatic method for the synthesis of a bicyclic[3.1.0]proline moiety “P2” (Fig. 8), the key structural feature in Boceprevir (Li et al. 2012). Since the crystal structure of MAO-N was not available at that time, error-prone PCR and homology model-guided mutagenesis were employed for the first round of directed evolution, and variant MAO-N 156 with two single mutations A289V and K348Q was obtained, which showed 2.4-fold higher activity in comparison with WT. Thereafter, two parallel strategies were chosen in the second round evolution: recombination of positive mutations with MAO-N 156; family shuffling of MAO-N 156 with the homolog from A. oryzae, leading to two positive hits (MAO-N 274 and MAO-N 291) with respective threefold and sixfold improvement in activity. Two further rounds of directed evolution were conducted based on family shuffling and active site-targeted mutation. This finally provided the super variant MAO-N 401, which was successfully applied for large-scale production of (1R,2S,5S)-6,6-dimethyl-3-azabicyclo[3.1.0]hexane-2-carbonitrile (26) with correct absolute configuration and > 99% ee value in a chemoenzymatic process, thereby significantly decreasing the cost, while improving the sustainability for the production of Boceprevir (Li et al. 2012).

Fig. 8
figure 8

The Merck/Codexis process for manufacturing a Boceprevir intermediate based on amine oxidase-catalyzed desymmetrization

Directed evolution of monoamine oxidase from Brevibacterium oxydans IH-35A (CHAO)

The monoamine oxidase CHAO, a 50 kDa flavoprotein responsible for the oxidation of cyclohexylamine to cyclohexanone in Brevibacterium oxydans IH-35A, displays high activity towards a wide range of structurally different primary amines (Leisch et al. 2011). However, deracemization of secondary amines via WT CHAO cannot be achieved due to low or no activity towards this kind of amines. To explore the potential of CHAO for the synthesis of optically pure secondary amines, for example 1,2,3,4-tetrahydroquinoline (THQ) derivatives, directed evolution was carried out for the first time in Dunming Zhu’s group (Li et al. 2014b). In this attempt, 11 amino acid residues (F88, T198, L199, M226, Q233, Y321, F351, L353, F368, P422, and Y459) located in the active center of CHAO were targeted for saturation mutagenesis. Four single mutants (T198F, L199T, M226F, and Y459T), displaying more than 20-fold activity towards 2-methyl-THQ relative to WT CHAO, were identified. Subsequently, ISM were performed on the four positive “hot” positions (T198, L199, M226, and Y459) for improving the activity further. This led to two new hits, namely T198F/L199S and T198F/L199S/M226F. The triple variant T198F/L199S/M226F was successfully applied for the deracemization of 2-methyl-1,2,3,4-tetrahydroquinoline on a preparative scale, giving 76% isolated yield and 98% ee of the (R)-enantiomer (Fig. 9).

Fig. 9
figure 9

Deracemization of 2-methyl-1,2,3,4-tetrahydroquinoline using triple variant T198F/L199S/M226F and chemical reductant

To further expand the biocatalytic repertoire of CHAO, Yao et al. created a new library containing diverse mutants, and then assayed them towards a panel of 2-substituted THQs (Fig. 10). Several highly (S)-selective mutants with notably enhanced activity were identified. Significantly, variant L225A showed reversed selectivity (R-selectivity) in reactions of 1,2,3,4-tetrahydro-2-methylquinoline (31), 1,2,3,4-tetrahydro-2-isopropylquinoline (32), 2-cyclopropyl-1,2,3,4-tetrahydroquinoline (33), and 2-(2-Benzo[1,3]dioxol-5-yl-ethyl)-1,2,3,4-tetrahydroquinoline (42). Molecular dynamic simulations were conducted based on the models of variant L225A harboring (R)- and (S)- cyclopropyl-THQ, and computational results revealed the mechanism of enantioselectivity reversal (Yao et al. 2018).

Fig. 10
figure 10

2-Substituted THQs were tested as substrates in deracemization using CHAO mutants

Directed evolution of R-selective monoamine oxidases

Directed evolution of MAO-N and CHAO has provided diversified biocatalysts for the synthesis of chiral amines. However, until 2014, no enantio-complementary (R)-amine oxidases had been reported for general application. In this respect, Asano et al. evolved a flavin-dependent porcine kidney D-amino acid oxidase (pkDAO) into an R-stereoselective amine oxidase (Yasukawa et al. 2014). Based on the crystal structure of pkDAO, two residues Tyr228 and Arg283 at the active center were subjected to saturation mutagenesis. Screening the respective libraries towards the model amine rac-α-methylbenzylamine identified three hits, namely R283G, R283A, and R283C. These were then used as templates for ISM at residue Tyr 228, leading to three improved mutants (Y228L/R283G, Y228L/R283A and Y228L/R283C). Deracemization of rac-α-methylbenzylamine on a preparative scale was conducted using the best variant Y228L/R283G, producing 99% (S)-α-methylbenzylamine with 65% isolated yield. Crystal structure analysis revealed that the mutations introduced in the directed evolution induced critical changes at the active center, which in turn render the α-hydrogen atom of (R)-α-methylbenzylamine to point towards the N5 atom of FAD, while in the case of (S)-α-methylbenzylamine, the N5 atom of FAD remains remote, therefore, leading to high R-stereoselectivity (Yasukawa et al. 2014). For further expanding the substrate specificity of pkDAO towards bulky amines, such as amine 15, variants R283G and Y228L/R283G were chosen as templates for saturation mutagenesis at residues Leu51, Ile215, and Ile230, which could play a critical role in accommodating the bulky substrate (Yasukawa et al. 2018). After high-throughput screening with colorimetric assay, four positive mutants, namely I230A/R283G, I230C/R283G, I230F/R283G, and Y228L/I230C/R283G, were identified. Among them, variant I230A/R283G showed highest catalytic efficiency towards (S)-15, and the deracemization of rac-15 using this variant was also achieved successfully within 1 h, producing (R)-15 in 98% ee. As expected, the crystal structure of variant I230A/R283G indicated that the mutations introduced in the variant provided extra space to accommodate the 4-Cl-phenyl ring of amine 15 (Yasukawa et al. 2018).

Another R-selective amine oxidase was developed based on 6-hydroxy-D-nicotine (6-HDNO) in Turner’s group. Wild-type 6-HDNO displayed relative narrow substrate scope, and hence, directed evolution again based on CASTing was performed with the aim to broaden its substrate spectrum (Heath et al. 2014). Two saturation mutagenesis libraries A (Leu373/Leu375) and B (Glu350/Glu352) were constructed using NNK code degeneracy, which were then screened with the colorimetric solid-phase assay. Although no positive variants was found in library A, three positive hits, namely E350L/E352D, E350 V/E352D, and E350L, were identified in library B. The E350L/E352D variant, showing significant improvement in substrate specificity, was used for the deracemization of several amines. As expected, in all cases, the variant exhibited R-stereoselectivity, the opposite enantioselectivity observed when using MAO-N (Heath et al. 2014).

Conclusion

Monoamine oxidases use molecular dioxygen as the stoichiometric oxidant to catalyze the irreversible oxidation of amines to imines. This feature avoids the problem of controlling the reaction equilibrium position, making them an attractive class of enzyme for chiral amine synthesis (Turner 2011). Relative to the reported methods for chiral amine preparation, such as classical resolution of the corresponding racemate (Siedlecka 2013), lipase-catalyzed kinetic resolution (Poulhès et al. 2011; Verho and Bäckvall 2015; Oláh et al. 2016; Gustafson et al. 2014), and asymmetric hydrogenation of imines (Liu and Du 2013; Ghattas et al. 2012; de Vries and Mršić 2011; Li et al. 2016; Echeverria et al. 2016; Lautens and Larin 2018), monoamine oxidases, once subjected to protein engineering, display widespread substrate specificity in Turner-type deracemization for the industrial synthesis of enantiomerically pure primary, secondary, and tertiary amines as well as chiral heterocyclic amines. In this deracemization process, the chemical reducing agent (typically the ammonia borane complex NH3BH3) is added in excess amount for reducing the imine non-selectively back to the amine. To avoid an excess of NH3BH3, imine reductase as an excellent alternative of chemical reducing agent was successfully applied for the synthesis of chiral amines in a new cascade reaction based on the combination of imine reductase with amine oxidase (Gand et al. 2014; Leipold et al. 2013; Mangas-Sanchez et al. 2017; Mitsukura et al. 2010, 2013). These developments open the door for further industrial preparations of chiral amines by deracemization processes (Heath et al. 2016).

Although original monoamine oxidases have suffered traditionally from distinct limitations; for example, low activity, narrow substrate scope, wrong stereoselectivity, and insufficient thermostability, with the help of directed evolution as summarized here, most problems have been solved, thereby creating a versatile “toolbox” of monoamine oxidases. CAST/ISM has proven to be a particularly viable mutagenesis strategy. With the advance of new protein engineering techniques, such as rational design (Choi et al. 2015; Otten et al. 2010) and machine learning (Li et al. 2019), it can be hoped that new and useful mutants of various monoamine oxidases will be generated even more rapidly and efficiently than in the past. In some cases, the insolubility of certain substrates in aqueous medium causes practical problems, which need to be solved in the future with the aid of bioprocess engineering. The pharmaceutical industry and other fine chemical companies will certainly profit from such potential advances.

Availability of data and materials

All data generated or analyzed during this study are included in this published article.

Abbreviations

MAOs:

monoamine oxidases

MAO-N:

monoamine oxidase from Aspergillus niger

CHAO:

cyclohexylamine oxidase from Brevibacterium oxydans IH-35A

pkDAO:

D-amino acid oxidase porcine kidney

6-HDNO:

6-hydroxy-D-nicotine oxidase

epPCR:

error-prone polymerase chain reaction

CAST:

Combinatorial Active Site Saturation Test

ISM:

iterative saturation mutagenesis

DKR:

dynamic kinetic resolution

WT:

wild type

THQ:

1,2,3,4-tetrahydroquinoline

References

  • Acevedo-Rocha CG, Agudo R, Reetz MT (2014) Directed evolution of stereoselective enzymes based on genetic selection as opposed to screening systems. J Biotechnol 191:3–10

    Article  CAS  PubMed  Google Scholar 

  • Acevedo-Rocha CG, Gamble CG, Lonsdale R, Li A, Nett N, Hoebenreich S, Lingnau JB, Wirtz C, Fares C, Hinrichs H (2018) P450-catalyzed regio-and diastereoselective steroid hydroxylation: efficient directed evolution enabled by mutability landscaping. ACS Catal 8(4):3395–3410

    Article  CAS  Google Scholar 

  • Alexeeva M, Enright A, Dawson MJ, Mahmoudian M, Turner NJ (2002) Deracemization of αα-methylbenzylamine using an enzyme obtained by in vitro evolution. Angew Chem Int Ed 41(17):3177–3180

    Article  CAS  Google Scholar 

  • Allen AE, MacMillan DW (2012) Synergistic catalysis: a powerful synthetic strategy for new reaction development. Chem Sci 3(3):633–658

    Article  CAS  Google Scholar 

  • Arnold FH (2018) Directed evolution: bringing new chemistry to life. Angew Chem Int Ed 57(16):4143–4148

    Article  CAS  Google Scholar 

  • Atodiresei I, Vila C, Rueping M (2015) Asymmetric organocatalysis in continuous flow: opportunities for impacting industrial catalysis. ACS Catal 5(3):1972–1985

    Article  CAS  Google Scholar 

  • Bálint J, Egri G, Czugler M, Schindler J, Kiss V, Juvancz Z, Fogassy E (2001) Resolution of α-phenylethylamine by its acidic derivatives. Tetrahedron Asymmetry 12(10):1511–1518

    Article  Google Scholar 

  • Blacker J, Headley CE (2010) Dynamic resolution of chiral amine pharmaceuticals: turning waste isomers into useful product. In: Dunn PJ, Wells A, Williams MT (eds) Green chemistry in the pharmaceutical industry. Wiley, Weinheim, pp 269–288

    Chapter  Google Scholar 

  • Blaser HU, Schmidt E (2004) Asymmetric catalysis on industrial scale. Wiley, Hoboken

    Google Scholar 

  • Bommarius AS, Blum JK, Abrahamson MJ (2011) Status of protein engineering for biocatalysts: how to design an industrially useful biocatalyst. Curr Opin Chem Biol 15(2):194–200

    Article  CAS  PubMed  Google Scholar 

  • Bornscheuer U, Huisman G, Kazlauskas RJ, Lutz S, Moore J, Robins K (2012) Engineering the third wave of biocatalysis. Nature 485(7397):185–194

    Article  CAS  PubMed  Google Scholar 

  • Brustad EM, Arnold FH (2011) Optimizing non-natural protein function with directed evolution. Curr Opin Chem Biol 15(2):201–210

    Article  CAS  PubMed  Google Scholar 

  • Cadwell RC, Joyce GF (1994) Mutagenic PCR. PCR Methods Appl 3(6):S136–S140

    Article  CAS  PubMed  Google Scholar 

  • Carr R, Alexeeva M, Enright A, Eve TS, Dawson MJ, Turner NJ (2003) Directed evolution of an amine oxidase possessing both broad substrate specificity and high enantioselectivity. Angew Chem Int Ed 42(39):4807–4810

    Article  CAS  Google Scholar 

  • Carr R, Alexeeva M, Dawson MJ, Gotor-Fernández V, Humphrey CE, Turner NJ (2005) Directed evolution of an amine oxidase for the preparative deracemisation of cyclic secondary amines. ChemBioChem 6(4):637–639

    Article  CAS  PubMed  Google Scholar 

  • Chen K, Arnold FH (1993) Tuning the activity of an enzyme for unusual environments: sequential random mutagenesis of subtilisin E for catalysis in dimethylformamide. Proc Natl Acad Sci USA 90(12):5618–5622

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Chen K, Huang X, Kan SJ, Zhang RK, Arnold FH (2018a) Enzymatic construction of highly strained carbocycles. Science 360(6384):71–75

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Chen K, Zhang S-Q, Brandenberg OF, Hong X, Arnold FH (2018b) Alternate heme ligation steers activity and selectivity in engineered cytochrome P450-catalyzed carbene-transfer reactions. J Am Chem Soc 140(48):16402–16407

    Article  CAS  PubMed  Google Scholar 

  • Choi J-M, Han S-S, Kim H-S (2015) Industrial applications of enzyme biocatalysis: current status and future aspects. Biotechnol Adv 33(7):1443–1454

    Article  CAS  PubMed  Google Scholar 

  • Clouthier CM, Kayser MM, Reetz MT (2006) Designing new Baeyer-Villiger monooxygenases using restricted CASTing. J Org Chem 71(22):8431–8437

    Article  CAS  PubMed  Google Scholar 

  • de Vries JG, Mršić N (2011) Organocatalytic asymmetric transfer hydrogenation of imines. Catal Sci Technol 1(5):727–735

    Article  CAS  Google Scholar 

  • Denard CA, Ren H, Zhao H (2015) Improving and repurposing biocatalysts via directed evolution. Curr Opin Chem Biol 25:55–64

    Article  CAS  PubMed  Google Scholar 

  • Drauz K, Gröger H, May O (2012) Enzyme catalysis in organic synthesis, vol 1. Wiley, Hoboken

    Book  Google Scholar 

  • Dunsmore CJ, Carr R, Fleming T, Turner NJ (2006) A chemo-enzymatic route to enantiomerically pure cyclic tertiary amines. J Am Chem Soc 128(7):2224–2225

    Article  CAS  PubMed  Google Scholar 

  • Echeverria P-G, Ayad T, Phansavath P, Ratovelomanana-Vidal V (2016) Recent developments in asymmetric hydrogenation and transfer hydrogenation of ketones and imines through dynamic kinetic resolution. Synthesis 48(16):2523–2539

    Article  CAS  Google Scholar 

  • Estell DA, Graycar TP, Wells JA (1985) Engineering an enzyme by site-directed mutagenesis to be resistant to chemical oxidation. J Biol Chem 260(11):6518–6521

    CAS  PubMed  Google Scholar 

  • Gand M, Müller H, Wardenga R, Höhne M (2014) Characterization of three novel enzymes with imine reductase activity. J Mol Catal B Enzym 110:126–132

    Article  CAS  Google Scholar 

  • Ghattas G, Chen D, Pan F, Klankermayer J (2012) Asymmetric hydrogenation of imines with a recyclable chiral frustrated Lewis pair catalyst. Dalton Trans 41(30):9026–9028

    Article  CAS  PubMed  Google Scholar 

  • Ghislieri D, Green AP, Pontini M, Willies SC, Rowles I, Frank A, Grogan G, Turner NJ (2013) Engineering an enantioselective amine oxidase for the synthesis of pharmaceutical building blocks and alkaloid natural products. J Am Chem Soc 135(29):10863–10869

    Article  CAS  PubMed  Google Scholar 

  • Gotor V, Alfonso I, García-Urdiales E (2008) Asymmetric organic synthesis with enzymes. Wiley, Hoboken

    Book  Google Scholar 

  • Guranda DT, van Langen LM, van Rantwijk F, Sheldon RA, Švedas VK (2001) Highly efficient and enantioselective enzymatic acylation of amines in aqueous medium. Tetrahedron Asymmetry 12(11):1645–1650

    Article  CAS  Google Scholar 

  • Gustafson KP, Lihammar R, Verho O, Engstrom K, Bäckvall J-E (2014) Chemoenzymatic dynamic kinetic resolution of primary amines using a recyclable palladium nanoparticle catalyst together with lipases. J Org Chem 79(9):3747–3751

    Article  CAS  PubMed  Google Scholar 

  • Harvey AL (2008) Natural products in drug discovery. Drug Discov Today 13(19–20):894–901

    Article  CAS  PubMed  Google Scholar 

  • Heath RS, Pontini M, Bechi B, Turner NJ (2014) Development of an R-selective amine oxidase with broad substrate specificity and high enantioselectivity. ChemCatChem 6(4):996–1002

    Article  CAS  Google Scholar 

  • Heath RS, Pontini M, Hussain S, Turner NJ (2016) Combined imine reductase and amine oxidase catalyzed deracemization of nitrogen heterocycles. ChemCatChem 8(1):117–120

    Article  CAS  Google Scholar 

  • Hermanns N, Dahmen S, Bolm C, Bräse S (2002) Asymmetric, catalytic phenyl transfer to imines: highly enantioselective synthesis of diarylmethylamines. Angew Chem Int Ed 41(19):3692–3694

    Article  CAS  Google Scholar 

  • Kirsch RD, Joly E (1998) An improved PCR-mutagenesis strategy for two-site mutagenesis or sequence swapping between related genes. Nucleic Acids Res 26(7):681–683

    Article  Google Scholar 

  • Köhler V, Bailey KR, Znabet A, Raftery J, Helliwell M, Turner NJ (2010) Enantioselective biocatalytic oxidative desymmetrization of substituted pyrrolidines. Angew Chem Int Ed 49(12):2182–2184

    Article  CAS  Google Scholar 

  • Lautens M, Larin EM (2018) Pd/Zn-catalyzed asymmetric transfer hydrogenation of imines with alcohols. Synfacts 14(07):0730

    Article  Google Scholar 

  • Lee S (1999) Designing enzymatic resolution of amines. Chem Commun 2:127–128

    Google Scholar 

  • Leipold F, Hussain S, Ghislieri D, Turner NJ (2013) Asymmetric reduction of cyclic imines catalyzed by a whole-cell biocatalyst containing an (S)-imine reductase. ChemCatChem 5(12):3505–3508

    Article  CAS  Google Scholar 

  • Leisch H, Grosse S, Iwaki H, Hasegawa Y, Lau PC (2011) Cyclohexylamine oxidase as a useful biocatalyst for the kinetic resolution and dereacemization of amines. Can J Chem 90(1):39–45

    Article  CAS  Google Scholar 

  • Leung DW (1989) A method for random mutagenesis of a defined DNA segment using a modified polymerase chain reaction. Technique 1:11–15

    Google Scholar 

  • Li T, Liang J, Ambrogelly A, Brennan T, Gloor G, Huisman G, Lalonde J, Lekhal A, Mijts B, Muley S (2012) Efficient, chemoenzymatic process for manufacture of the Boceprevir bicyclic [3.1. 0] proline intermediate based on amine oxidase-catalyzed desymmetrization. J Am Chem Soc 134(14):6467–6472

    Article  CAS  PubMed  Google Scholar 

  • Li G, Ren J, Iwaki H, Zhang D, Hasegawa Y, Wu Q, Feng J, Lau PC, Zhu D (2014a) Substrate profiling of cyclohexylamine oxidase and its mutants reveals new biocatalytic potential in deracemization of racemic amines. Appl Microbiol Biotechnol 98(4):1681–1689

    Article  CAS  PubMed  Google Scholar 

  • Li G, Ren J, Yao P, Duan Y, Zhang H, Wu Q, Feng J, Lau PC, Zhu D (2014b) Deracemization of 2-methyl-1,2,3,4-tetrahydroquinoline using mutant cyclohexylamine oxidase obtained by iterative saturation mutagenesis. ACS Catal 4(3):903–908

    Article  CAS  Google Scholar 

  • Li S, Li G, Meng W, Du H (2016) A frustrated Lewis pair catalyzed asymmetric transfer hydrogenation of imines using ammonia borane. J Am Chem Soc 138(39):12956–12962

    Article  CAS  PubMed  Google Scholar 

  • Li G, Yao P, Gong R, Li J, Liu P, Lonsdale R, Wu Q, Lin J, Zhu D, Reetz MT (2017) Simultaneous engineering of an enzyme’s entrance tunnel and active site: the case of monoamine oxidase MAO-N. Chem Sci 8(5):4093–4099

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Li G, Garcia-Borràs M, Fürst MJ, Ilie A, Fraaije MW, Houk KN, Reetz MT (2018a) Overriding traditional electronic effects in biocatalytic Baeyer–Villiger reactions by directed evolution. J Am Chem Soc 140(33):10464–10472

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Li G, Wang J-b, Reetz MT (2018b) Biocatalysts for the pharmaceutical industry created by structure-guided directed evolution of stereoselective enzymes. Biorg Med Chem 26(7):1241–1251

    Article  CAS  Google Scholar 

  • Li G, Dong Y, Reetz MT (2019) Can machine learning revolutionize directed evolution of selective enzymes? Adv Synth Catal 361(11):2377–2386

    CAS  Google Scholar 

  • List B (2010) Emil Knoevenagel and the roots of aminocatalysis. Angew Chem Int Ed 49(10):1730–1734

    Article  CAS  Google Scholar 

  • Liu Y, Du H (2013) Chiral dienes as “ligands” for borane-catalyzed metal-free asymmetric hydrogenation of imines. J Am Chem Soc 135(18):6810–6813

    Article  CAS  PubMed  Google Scholar 

  • MacMillan DW (2008) The advent and development of organocatalysis. Nature 455(7211):304–308

    Article  CAS  PubMed  Google Scholar 

  • Mangas-Sanchez J, France SP, Montgomery SL, Aleku GA, Man H, Sharma M, Ramsden JI, Grogan G, Turner NJ (2017) Imine reductases (IREDs). Curr Opin Chem Biol 37:19–25

    Article  CAS  PubMed  Google Scholar 

  • Mitsukura K, Suzuki M, Tada K, Yoshida T, Nagasawa T (2010) Asymmetric synthesis of chiral cyclic amine from cyclic imine by bacterial whole-cell catalyst of enantioselective imine reductase. Org Biomol Chem 8(20):4533–4535

    Article  CAS  PubMed  Google Scholar 

  • Mitsukura K, Kuramoto T, Yoshida T, Kimoto N, Yamamoto H, Nagasawa T (2013) A NADPH-dependent (S)-imine reductase (SIR) from Streptomyces sp. GF3546 for asymmetric synthesis of optically active amines: purification, characterization, gene cloning, and expression. Appl Microbiol Biotechnol 97(18):8079–8086

    Article  CAS  PubMed  Google Scholar 

  • Nguyen TB, Wang Q, Guéritte F (2011) Chiral phosphoric acid catalyzed enantioselective transfer hydrogenation of ortho-hydroxybenzophenone N–H ketimines and applications. Chem Eur J 17(35):9576–9580

    Article  CAS  PubMed  Google Scholar 

  • Ni Y, Holtmann D, Hollmann F (2014) How green is biocatalysis? To calculate is to know. ChemCatChem 6(4):930–943

    Article  CAS  Google Scholar 

  • Noyori R (2002) Asymmetric catalysis: science and opportunities (Nobel lecture). Angew Chem Int Ed 41(12):2008–2022

    Article  CAS  Google Scholar 

  • Oláh M, Boros Z, Hornyánszky G, Poppe L (2016) Isopropyl 2-ethoxyacetate—an efficient acylating agent for lipase-catalyzed kinetic resolution of amines in batch and continuous-flow modes. Tetrahedron 72(46):7249–7255

    Article  CAS  Google Scholar 

  • Otten LG, Hollmann F, Arends IW (2010) Enzyme engineering for enantioselectivity: from trial-and-error to rational design? Trends Biotechnol 28(1):46–54

    Article  CAS  PubMed  Google Scholar 

  • Porter JL, Rusli RA, Ollis DL (2016) Directed evolution of enzymes for industrial biocatalysis. ChemBioChem 17(3):197–203

    Article  CAS  PubMed  Google Scholar 

  • Poulhès F, Vanthuyne N, MlP Bertrand, Sp Gastaldi, Gr Gil (2011) Chemoenzymatic dynamic kinetic resolution of primary amines catalyzed by CAL-B at 38–40 °C. J Org Chem 76(17):7281–7286

    Article  PubMed  CAS  Google Scholar 

  • Quin MB, Schmidt-Dannert C (2011) Engineering of biocatalysts: from evolution to creation. ACS Catal 1(9):1017–1021

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Reetz MT (2011) Laboratory evolution of stereoselective enzymes: a prolific source of catalysts for asymmetric reactions. Angew Chem Int Ed 50(1):138–174

    Article  CAS  Google Scholar 

  • Reetz MT (2016a) Directed evolution of selective enzymes: catalysts for organic chemistry and biotechnology. Wiley, Hoboken

    Book  Google Scholar 

  • Reetz MT (2016b) What are the limitations of enzymes in synthetic organic chemistry? The Chemical Record 16(6):2449–2459

    Article  CAS  PubMed  Google Scholar 

  • Reetz MT, Carballeira JD (2007) Iterative saturation mutagenesis (ISM) for rapid directed evolution of functional enzymes. Nat Protoc 2(4):891–903

    Article  CAS  PubMed  Google Scholar 

  • Reetz MT, Zonta A, Schimossek K, Jaeger KE, Liebeton K (1997) Creation of enantioselective biocatalysts for organic chemistry by in vitro evolution. Angew Chem Int Ed Engl 36(24):2830–2832

    Article  CAS  Google Scholar 

  • Reetz MT, Bocola M, Carballeira JD, Zha D, Vogel A (2005) Expanding the range of substrate acceptance of enzymes: combinatorial active-site saturation test. Angew Chem Int Ed 44(27):4192–4196

    Article  CAS  Google Scholar 

  • Reetz MT, Carballeira JD, Peyralans J, Höbenreich H, Maichele A, Vogel A (2006a) Expanding the substrate scope of enzymes: combining mutations obtained by CASTing. Chem Eur J 12(23):6031–6038

    Article  CAS  PubMed  Google Scholar 

  • Reetz MT, Carballeira JD, Vogel A (2006b) Iterative saturation mutagenesis on the basis of B factors as a strategy for increasing protein thermostability. Angew Chem Int Ed 45(46):7745–7751

    Article  CAS  Google Scholar 

  • Reymond J-L (2006) Enzyme assays: high-throughput screening, genetic selection and fingerprinting. Wiley, Hoboken

    Google Scholar 

  • Rowles I, Malone KJ, Etchells LL, Willies SC, Turner NJ (2012) Directed evolution of the enzyme monoamine oxidase (MAO-N): highly efficient chemo-enzymatic deracemisation of the Alkaloid (±)-crispine A. ChemCatChem 4(9):1259–1261

    Article  CAS  Google Scholar 

  • Sharpless KB (2002) Searching for new reactivity (Nobel lecture). Angew Chem Int Ed 41(12):2024–2032

    Article  CAS  Google Scholar 

  • Sheldon RA, Pereira PC (2017) Biocatalysis engineering: the big picture. Chem Soc Rev 46(10):2678–2691

    Article  CAS  PubMed  Google Scholar 

  • Siedlecka R (2013) Recent developments in optical resolution. Tetrahedron 69(31):6331–6363

    Article  CAS  Google Scholar 

  • Siloto RM, Weselake RJ (2012) Site saturation mutagenesis: methods and applications in protein engineering. Biocatal Agric Biotechnol 1(3):181–189

    Article  CAS  Google Scholar 

  • Skalden L, Peters C, Ratz L, Bornscheuer UT (2016) Synthesis of (1R,3R)-1-amino-3-methylcyclohexane by an enzyme cascade reaction. Tetrahedron 72(46):7207–7211

    Article  CAS  Google Scholar 

  • Stemmer WP (1994) Rapid evolution of a protein in vitro by DNA shuffling. Nature 370(6488):389–391

    Article  CAS  PubMed  Google Scholar 

  • Tao JA, Lin G-Q, Liese A (2009) Biocatalysis for the pharmaceutical industry: discovery, development, and manufacturing. Wiley, Hoboken

    Book  Google Scholar 

  • Thalén LK, Zhao D, Sortais JB, Paetzold J, Hoben C, Baeckvall JE (2009) A chemoenzymatic approach to enantiomerically pure amines using dynamic kinetic resolution: application to the synthesis of norsertraline. Chem Eur J 15(14):3403–3410

    Article  PubMed  CAS  Google Scholar 

  • Turner NJ (2011) Enantioselective oxidation of C-O and C–N bonds using oxidases. Chem Rev 111(7):4073–4087

    Article  CAS  PubMed  Google Scholar 

  • Turner NJ, Truppo MD (2010) Biocatalytic routes to nonracemic chiral amines. In: Chiral amine synthesis. pp 431–459

    Chapter  Google Scholar 

  • Van Langen L, Oosthoek N, Guranda D, Van Rantwijk F, Švedas V, Sheldon R (2000) Penicillin acylase-catalyzed resolution of amines in aqueous organic solvents. Tetrahedron Asymmetry 11(22):4593–4600

    Article  Google Scholar 

  • Vandeyar MA, Weiner MP, Hutton CJ, Batt CA (1988) A simple and rapid method for the selection of oligodeoxynucleotide-directed mutants. Gene 65(1):129–133

    Article  CAS  PubMed  Google Scholar 

  • Verho O, Bäckvall J-E (2015) Chemoenzymatic dynamic kinetic resolution: a powerful tool for the preparation of enantiomerically pure alcohols and amines. J Am Chem Soc 137(12):3996–4009

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Vetica F, Chauhan P, Dochain S, Enders D (2017) Asymmetric organocatalytic methods for the synthesis of tetrahydropyrans and their application in total synthesis. Chem Soc Rev 46(6):1661–1674

    Article  CAS  PubMed  Google Scholar 

  • Walsh PJ, Kozlowski MC (2009) Fundamentals of asymmetric catalysis. University Science Books, Sausalito

    Google Scholar 

  • Wang Y-B, Tan B (2018) Construction of axially chiral compounds via asymmetric organocatalysis. Acc Chem Res 51(2):534–547

    Article  CAS  PubMed  Google Scholar 

  • Yao P, Cong P, Gong R, Li J, Li G, Ren J, Feng J, Lin J, Lau PC, Wu Q (2018) Biocatalytic route to chiral 2-substituted-1,2,3,4-tetrahydroquinolines using cyclohexylamine oxidase muteins. ACS Catal 8(3):1648–1652

    Article  CAS  Google Scholar 

  • Yasukawa K, Nakano S, Asano Y (2014) Tailoring d-amino acid oxidase from the pig kidney to R-stereoselective amine oxidase and its use in the deracemization of αα-Methylbenzylamine. Angew Chem Int Ed 53(17):4428–4431

    Article  CAS  Google Scholar 

  • Yasukawa K, Motojima F, Ono A, Asano Y (2018) Expansion of the substrate specificity of porcine kidney d-amino acid oxidase for S-stereoselective oxidation of 4-Cl-benzhydrylamine. ChemCatChem 10(16):3500–3505

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Zeymer C, Hilvert D (2018) Directed evolution of protein catalysts. Annu Rev Biochem 87:131–157

    Article  CAS  PubMed  Google Scholar 

  • Zhang RK, Chen K, Huang X, Wohlschlager L, Renata H, Arnold FH (2019a) Enzymatic assembly of carbon-carbon bonds via iron-catalysed sp3 C-H functionalization. Nature 565(7737):67–72

    Article  CAS  PubMed  Google Scholar 

  • Zhang RK, Huang X, Arnold FH (2019b) Selective C-H bond functionalization with engineered heme proteins: new tools to generate complexity. Curr Opin Chem Biol 49:67–75

    Article  CAS  PubMed  Google Scholar 

  • Zheng L, Baumann U, Reymond J-L (2004) An efficient one-step site-directed and site-saturation mutagenesis protocol. Nucleic Acids Res 32(14):e115–e115

    Article  PubMed  PubMed Central  Google Scholar 

  • Zhou Q-L (2011) Privileged chiral ligands and catalysts. Wiley, Hoboken

    Book  Google Scholar 

Download references

Acknowledgements

GYL thanks the National Natural Science Foundation of China (Grant No. 21807111), the fund of Elite Youth Program of CAAS, and Agricultural Science and Technology Innovation Program of CAAS (CAAS-XTCX2019024). We also thank Professor Reetz for critical reading and modification of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 21807111), the fund of Elite Youth Program of CAAS, and Agricultural Science and Technology Innovation Program of CAAS(CAAS-XTCX2019024).

Author information

Authors and Affiliations

Authors

Contributions

GYL and JQD conceived and wrote this paper. BBL, YCQ, YJD, and JR were involved collecting related material and critical reading of this paper. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Guangyue Li.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Duan, J., Li, B., Qin, Y. et al. Recent progress in directed evolution of stereoselective monoamine oxidases. Bioresour. Bioprocess. 6, 37 (2019). https://doi.org/10.1186/s40643-019-0272-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40643-019-0272-6

Keywords