Enzymatic Cascades for the Regio- and

0 downloads 0 Views 1MB Size Report
important alternatives to more traditional chemical catalysts [1, 2]. ... an (R)-selective ω-TA which has replaced a rhodium-catalyzed asymmetric enamine ...
Enzymatic Cascades for the Regio- and Stereoselective Synthesis of Chiral Amines School of Chemistry, University of Manchester, Manchester Institute of Biotechnology, 131 Princess Street, Manchester, M1 7DN, UK Elaine O’Reillyǂ & Nicholas J. Turner E-Mail: [email protected] [email protected]

Abstract Significant advancements in protein engineering and DNA technology have seen biocatalytic transformations take the place of traditional chemical manipulations in both academia and industry for the preparation of active pharmaceutical ingredients (APIs) and other medicinally relevant compounds. However, despite the large repertoire of commercially available biocatalysts which are readily accessible, enzymes which mediate the formation of C-C bonds and those which enable convergent synthesis remain largely undeveloped. In order to expand the scope of biocatalytic retrosynthesis and enable it to complement traditional chemical retrosynthesis it is essential to develop a ‘toolbox’ of biocatalysts which build molecular complexity. Of particular interest is the development of one-pot enzymatic cascades for the synthesis of functionalised, chiral building blocks without the need for protecting group manipulations or harsh reaction conditions. Highly regio- and stereoselective chemoenzymatic cascades have been developed for the synthesis of a range of chiral amines employing ω-transaminases and monoamine oxidase variants.

Introduction Nature boasts an extraordinary variety of biocatalysts which have the ability to mediate synthetically challenging transformations with exceptional levels of regio- and stereoselectivity. These properties, coupled with advances in protein engineering and DNA technologies mean that enzymes represent important alternatives to more traditional chemical catalysts [1, 2]. Directed evolution has allowed the substrate range, stability and kinetic properties of enzymes to be tailored and means that we no longer rely solely on their natural activity [3]. Addition of enzymes to the catalytic “toolbox”, including a diverse range of commercially available biocatalysts (figure 1), has enabled chemists to redesign synthetic routes to important target structures including bioactive natural products and active pharmaceutical ingredients (APIs), providing more efficient and cost-effective syntheses. A number of chemoenzymatic routes to the blockbuster API, Lipitor, have been developed including ketoreductase-, nitrilase-, halohydrin dehalogenase- and aldolase-mediated transformations and these diverse approaches are testament to the broad synthetic utility of biocatalysts (figure 2) [1, 4]. Chiral amines are prevalent motifs in drug molecules and the development of efficient synthetic routes to optically pure derivatives remains a fundamental goal for the pharmaceutical industry. ωtransaminases (ω-TAs) and variants of monoamine oxidase from Aspergillus niger (MAO-N) represent two important families of biocatalysts for the production of optically pure chiral amines. TAs can mediate the reductive amination of pro-chiral ketones and allow access to the corresponding (S)- or (R)-amines in high optical purity [5-9]. Recognising their potential, Merck and Codexis have engineered an (R)-selective ω-TA which has replaced a rhodium-catalyzed asymmetric enamine ǂ

Current address – School of Science & the Environment, Manchester Metropolitan University, John Dalton East, Oxford Road, Manchester, M1 5GD, UK

hydrogenation in the industrial production of the oral antihyperglycemic drug, Sitagliptin (figure 3) [7]. MAO-N catalyses the oxygen-dependent conversion of amines to imines [10-17] and has been extensively engineered to accommodate a diverse range of architectures, including the API, boceprevir [15] (figure 3) and related motifs. The hydrogen peroxide produced following oxidation has contributed towards the successful evolution of MAO-N as it has enabled the development of a highly effective high-throughput screen which is substrate independent (figure 4).

Figure 1. Examples of commercially available enzymes from the ‘biocatalytic toolbox’.

Figure 2. Chemoenzymatic routes for the synthesis of the side-chain of Atorvastatin; KRED = ketoreductase; HHDH = halohydrin dehalogenase.

Figure 3. APIs Sitagliptin and Boceprevir accessed via a chemoenzymatic route employing and engineered ωTA and MAO-N variant.

Figure 4. A high-throughput colorimetric screen for monoamine oxidase activity employing horseradish peroxidase (HRP) and a HRP substrate.

Chemo-enzymatic cascade for the regio- and stereoselective synthesis of 2,5disubstituted pyrrolidines Enzymatic and chemo-enzymatic cascades involving the combination of several enzymatic transformations in a concurrent one-pot process allows complex molecular architectures to be accessed with high optical purity, under ambient conditions and without the need for costly intermediate purification steps or protecting group manipulations [18-20]. Elegant cascades employing both ω-TAs [18-21] and MAO-N variants [10-17] have been reported for the asymmetric synthesis of a range of important amine scaffolds as well as for effective high-throughput screens. However, until recently, a cascade involving a combination of these two important amine synthesis enzymes had not been explored. Considerable efforts have focused on the development of asymmetric routes to both cis- and trans2,5-disubstituted pyrrolidines as these structures represent important scaffolds in pharmaceutical drugs and natural products [22-29]. Accessing the trans-diastereomers via reduction of the corresponding imine in high de is not straightforward and led us to develop a tandam biocatalytic approach to their synthesis. Our strategy (figure 5) encompasses a highly regio- and stereoselective TA mediated reductive amination and spontaneous cyclisation followed by a diastereoselective chemo-enzymatic conversion to the corresponding pyrrolidine using a combination of an (S)-selective MAO-N variant and the non-selective reducing agent NH3.BH3 [30].

Figure 5. Chemoenzymatic approach to the synthesis of chiral 2,5-disubstituted pyrrolidines [30]

We initially examined the reductive amination of non-symmetrical diketone 1a with the commercially available (S)-selective ω-TA, ATA113, employing the lactate dehydrogenase (LDH)/glucose dehydrogenase (GDH) system. Following reductive amination exclusively on the sterically less demanding methyl ketone, spontaneous cyclisation provided the corresponding 2,5-pyrroline 3a in excellent yield and high optical purity (91% yield, > 99% ee) (figure 6a). A novel transaminase (pfATA) from Pseudogulbenkiania ferrooxidans was also explored which shares 95% sequence identity with the transaminase from Chromobacterium violaceum (cv-ATA) (ATCC 12472). pf-ATA also mediated the reductive amination of diketone 1a, however, the 2,5-pyrroline 3a was obtained with a modest ee (75%). Commercially available (R)-selective ATA117 catalysed the regio- and stereoselective mono-amination of 1a, affording pyrroline (R)-3a in 65% yield and >99% ee (figure 6b). ATA113 and pf-ATA were then tested for their ability to catalyse the reductive amination of a series of structurally related diketones 1b-d (table 1). ATA113 provided methyl-substituted pyrrolines 3b-g as the sole regioisomers in excellent conversion and high ee, while substitution of the small methyl for a larger ethyl group resulted in a slightly reduced ee (entry 11 & 13). Pf-ATA provided pyrrolines 3b-g in modest to poor ee and surprisingly, replacement of the methyl substituent by ethyl resulted in a switch in stereoselectivity, affording (R)-3f-g as the predominant enantiomers (entries 12 & 14).

Figure 6. One-pot reductive amination of diketone 1a employing a) ATA113 and b) ATA117

Table 1. ω-TA mediated reductive amination of 3a-g

Entry

Substrate

1

1a

2 3 4 5 6

1a 1b 1b 1c 1c

ω-TA

Conv. [%]

ee [%]

ATA113

> 99

> 99 (S)

PfATA

> 99 > 99 > 99 > 99 > 99

75 (S) > 99 (S) > 78 (S) > 99 (S) 68 (S)

ATA113 PfATA

ATA113 PfATA

7 8 9 10 11 12 13 14

1d 1d 1e 1e 1f 1f 1g 1g

ATA113 PfATA

ATA113 PfATA

ATA113 PfATA

ATA113 PfATA

> 99 > 99 > 99 > 99 60 > 99 > 99 75

> 99 (S) 76 (S) > 99 (S) 78 (S) 96 (S) 76 (R) 94 (S) 46 (R)

Kroutil and co-workers recently developed a chemo-enzymatic route to related 2,6-disubstituted piperidines employing an ω-transaminase followed by diastereoselective hydrogenation using Pd/C [31]. The same strategy is not effective for the synthesis of the five-membered derivatives due to poor diastereoselectivity in the reduction step and so we sought to employ MAO-N for a chemo-enzymatic deracemisation. Two MAO-N variants (D5 and D9) were selected which have previously been evolved to accept structurally related amine frameworks, including pyrrolidines and piperidines. (S)and (R)-3a were exposed to NH3.BH3 in the presence of (S)-selective MAO-N, resulting in the accumulation of (2S,5R)- or (2R,5R)-pyrrolidine 4a in excellent de (>99%) following successive rounds of non-selective reduction/(S)-selective oxidation (figure 7). The approach exploits the exquisite regio- and stereoselectivity displayed by MAO-N which avoids stereorandomisation of the C-2-stereocentre generated by the (S)-selective ω-TA. Using the combination of NH3.BH3 /MAO-N D9 allowed access to 4b-g with excellent de (96-99%) through successive rounds of non-selective reduction/selective oxidation (table 2).

Figure 7. MAO-N mediated synthesis of (2S,5R)-4a and (2R,5R)-4a

Table 2. MAO/NH3.BH3 mediated asymmetric synthesis of (2S,5R)-4a-g

Entry

Substrate

1 2 3 4 5 6 7 8 9 10 11 12 13 14

3a 3a 3b 3b 3c 3c 3d 3d 3e 3e 3f 3f 3g 3g

MAO-N variant D5 D9 D5 D9 D5 D9 D5 D9 D5 D9 D5 D9 D5 D9

de [%] (2S,5R) > 99 96 88 98 > 99 > 99 > 99 > 99 68 > 99 64 > 99 56 96

The compatibility of the ω-TA and MAO-N under the same reaction conditions has also allowed the development of a one-pot cascade (figure 8) for the synthesis of target (2S,5R)-4a in 82% isolated yield and >99% de. We have also extended the one-pot TA/MAO-N approach for the preparation of (2S,5R)-4b, (2S,5R)-4d and (2S,5R)-4e in > 99% conversion and > 99% de starting from the corresponding diketones (table 3), demonstrating the generality of this one-pot approach.

Figure 8. One-pot asymmetric synthesis of (2S,5R)-4a on preparative-scale

Table 3. ATA113/MAO-N one-pot cascade for the synthesis of (2S,5R)-4a-4b, and (2S,5R)-4d-e Ketone

ω-TA

MAO-N

Conv. [%]

de [%]

1aǂ

ATA113

D5

> 99

> 99 (2S,5R)-4a

1bǂǂ 1dǂǂ 1eǂǂ

ATA113 ATA113 ATA113

D9 D9 D9

> 99 > 99 > 99

> 99 (2S,5R)-4b > 99 (2S,5R)-4d > 99 (2S,5R)-4e

ǂ

25mM substrate concentration; ǂǂ5mM substrate concentration

Monoamine oxidase/ω-Transaminase Cascade for the Deracemisation and Dealkylation of Amines While the deracemisation of cyclic amines using MAO-N variants has proven to be an effective means of accessing optically pure derivatives, employing this approach for the deracemisation of acyclic derivatives is limited due to competing hydrolysis of the imine in aqueous conditions. We have exploited a MAO-N/ω-TA cascade which relies on this competing hydrolysis for the effective deracemisation of a panel of acyclic amines 5a-h (figure 9, table 4) [32].

Figure 9. MAO-N/ω-TA cascade for the overall dealkylation of amines 5a-h.

Optically pure MBA analogues have been employed as chiral auxiliaries [33], for chiral derivatisation [34] and for the synthesis of peptides [35, 36]. The effective kinetic resolution of MBA-derivatives using ω-TAs has been reported, although employing this method for accessing optically pure amines results in a maximum 50% yield. A two-step approach has been successfully used for the deracemisation of amines by exploiting enantio-complementary TAs to achieve an initial resolution followed by an asymmetric synthesis [37]. Our approach takes advantage of a MAO-N mediated enantioselective oxidation of the (S)-amine to the corresponding ketone followed by reductive amination employing an (R)-selective TA. Exploiting this method, optically pure amines 5a-d can be accessed in high conversion and >99% ee. Further optimisation is ongoing with the aim of reducing the quantity of expensive alanine donor employed including the possibility of recycling the methyl-, ethyl- and isopropylamine by-products formed during the dealkylation as an alternative amine donor source. The same method can also be used for the regioselective dealkylation of amines. Oxidation of N-alkyl amines 6a-c using MAO-N provided the corresponding aldehyde which was quantitatively converted to 8 by the action of a commercially available (R)-selective TA (ATA117) (figure 10).

Table 4. Deracemisation of racemic amines 1a-h using a MAO-N/ω-TA cascade process Rac-1

Conversion[a]

5a 5b 5c 5d 5e 5f 5g 5h

99% 87% 91% 90% 99% 90% 90% 81%

ee (R)-enantiomer >99% >99% >99% >99% >99% >99% >99% >99%

[a] Conversion is based on conversion of the ketone

Figure 10. MAO-N/ω-TA cascade for the overall dealkylation of amines 6a-c

Conclusions The first examples of the use of transaminases in combination with monoamine oxidase variants have been developed. The compatibility of these important amine synthesis biocatalysts enables the regioand stereoselective synthesis of chiral 2,5-disubstituted pyrrolidine building blocks in excellent conversion (>99 %) and de (96-99 %). The same combination also enables the deracemisation and dealkylation of a range of methylbenzylamine derivatives. This technology in combination with the expanding ‘toolbox’ of transaminase and monoamine oxidase variants means that this approach will have broad utility for the construction of chiral amine frameworks.

References [1]

Reetz, M. T. (2013) Biocatalysis in organic chemistry and biotechnology: past, present, and future. J. Am. Chem. Soc. 135:12480 – 12496.

[2]

Turner, N. J. and O’ Reilly, E. (2013) Biocatalytic Retrosynthesis. Nat. Chem. Biol. 9: 285-288.

[3]

Turner, N. J. (2009) Directed evolution drives the next generation of biocatalysts. Nat. Chem. Biol. 5:567 - 573.

[4]

Ma, S.K. et. al. (2010) A green-by-design biocatalytic process for atorvastatin intermediate. Green Chem. 12:81 – 86.

[5]

Mathew, S. and Yun, H. (2012) ω-Transaminases for the production of optically pure amines and unnatural amino acids. ACS Catal. 2:993 – 1001.

[6]

Höhne, M. and Bornscheuer, U. (2009) Biocatalytic routes to optically active amines. ChemCatChem. 1:42 – 51.

[7]

Savile, C. K., Janey, J. M., Mundorff, E. C., Moore, J. C., Tam, S., Jarvis, W. R., Colbeck, J. C., Krebber, A., Fleitz, F. J., Brands, J., Devine, P. N., Huisman, G. W. and Hughes, G. J. (2010) Biocatalytic asymmetric synthesis of chiral amines from ketones applied to Sitagliptin manufacture. Science 329:305 –309.

[8]

Koszelewski, D., Tauber, K., Faber, K. and Kroutil, W. (2010) ω-Transaminases for the synthesis of non-racemic α-chiral primary amines. Trends Biotechnol. 28:324 – 332.

[9]

Frodsham, L., Golden, M., Hard, S., Kenworthy, M. N., Klauber, D. J., Leslie, K., Macleod, C., Meadows, R. E., Mulholland, K. R., Reilly, J., Squire, C., Tomasi, S.,Watt, D. and Wells, A. S. (2013) Use of ω-Transaminase enzyme chemistry in the synthesis of a JAK2 kinase inhibitor. Org. Process Res. Dev. 17:1123 -1130.

[10]

Alexeeva, M., Enright, A., Dawson, M. J., Mahmoudian, M. and Turner, N. J. (2002) Deracemization of alpha-methylbenzylamine using an enzyme obtained by in vitro evolution. Angew. Chem. Int. Ed. 41:3177 – 3180.

[11]

Carr, R., Alexeeva, M., Enright, A., Eve, T. S. C., Dawson, M. J. and Turner, N. J. Directed evolution of an amine oxidase possessing both broad substrate specificity and high enantioselectivity. Angew. Chem. Int. Ed. 42:4807 – 4810.

[12]

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

[13]

Rowles, I., Malone, K. J., Etchells, L. L., Willies, S. C. and Turner, N. J. (2012) Directed evolution of the enzyme monoamine oxidase (MAO-N): Highly efficient chemo-enzymatic deracemisation of the alkaloid (±)-Crispine A. ChemCatChem. 4:1259 – 1261.

[14]

Znabet, A., Polak, M. M., Janssen, E., de Kanter, F. J. J., Turner, N. J., Orru, R. V. A. and Ruijter, E. (2010) A highly efficient synthesis of telaprevir by strategic use of biocatalysis and multicomponent reactions. Chem. Commun. 46: 7918 – 7920.

[15]

Li, T., Liang, J., Ambrogelly, A., Brennan, T., Gloor, G., Huisman, G., Lalonde, J., Lekhal, A., Mijts, B., Muley, S., Newman, L., Tobin, M., Wong, G., Zaks, A. and Zhang, X. (2012) Efficient, chemoenzymatic process for manufacture of the Boceprevir bicyclic [3.1.0]proline intermediate based on amine oxidase-catalyzed desymmetrization. J. m. Chem. Soc. 134:6467 – 6472.

[16]

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

[17]

Ghislieri, D., Houghton, D., Green, A. P., Willies, S. C. and Turner, N. J. (2013) Monoamine oxidase (MAO-N) catalyzed deracemization of tetrahydro-β-carbolines: Substrate dependent switch in enantioselectivity. ACS Catal. 3:2869 – 2872.

[18]

Schrittwieser, J. H., Sattler, J., Resch, V., Mutti, F. G., Kroutil, W. (2011) Recent biocatalytic oxidation–reduction cascades. Curr. Opin. Chem. Biol. 15:249-256.

[19]

Oroz-Guinea, I. and García-Junceda, E. (2013) Enzyme catalysed tandem reactions. Curr. Opin. Chem. Biol. 17:236 -249.

[20]

Ricca, E., Brucher, B. and Schrittwieser, J. H. (2011) Multi-enzymatic cascade reactions: Overview and perspectives. Adv. Synth. Catal. 353:2239-2262.

[21]

Sattler, J. H., Fuchs, M., Tauber, K., Mutti, F. G., Faber, K., Pfeffer, J., Haas, T. and Kroutil, W. (2012) Redox Self-Sufficient Biocatalyst Network for the Amination of Primary Alcohols. Angew. Chem. Int. Ed. 51:9156 – 9159.

[22]

Brenneman, J. B., Martin, S. F. (2004) Application of Intramolecular Enyne metathesis to the synthesis of Aza[4.2.1]bicyclics:  Enantiospecific total synthesis of (+)-Anatoxin-a. Org. Lett. 6:1329-1331.

[23]

Xu, F. et. al.(2013) Asymmetric synthesis of cis-2,5-disubstituted pyrrolidine, the core scaffold of β3-AR agonists. Org. Lett. 15:1342-1345.

[24]

Hanessian, S., Bayrakdarian, M. and Luo, X. (2002) Total Synthesis of A-315675:  A potent inhibitor of influenza neuraminidase. J. Am. Chem. Soc. 124:4716 – 4721.

[25]

Aggarwal, V. K., Astle, C. J. and Rogers-Evans, M. (2004) A concise asymmetric route to the bridged bicyclic tropane alkaloid ferruginine using enyne ring-closing metathesis. Org. Lett. 6:1469-1471.

[26]

Zhang, S., Xu, L., Miao, L., Shu, H. and Trudell, M. L. (2007) General strategy for the construction of enantiopure pyrrolidine-based alkaloids. Total synthesis of (−)monomorine. J. Org. Chem. 72:3133 – 3136.

[27]

Goti, A., Cicchi, S., Mannucci, V., Cardona, F., Guarna, F., Marino, P. and Tejero, T. (2003) Iterative organometallic addition to chiral hydroxylated cyclic nitrones:  Highly stereoselective syntheses of α,α‘- and α,α-substituted hydroxypyrrolidines Org. Lett. 5:4235 - 4238.

[28]

Trost, B. M., Horne, D. B. and Woltering, M. J. (2003) Palladium-catalyzed DYKAT of vinyl epoxides: Enantioselective total synthesis and assignment of the configuration of (+)Broussonetine G. Angew. Chem. Int. Ed. 42:5987 – 5990.

[29]

Severino, E. A. and Correia, C. R. D. (2000) Heck arylation of endocyclic enecarbamates with diazonium salts. Improvements and a concise enantioselective synthesis of (−)-Codonopsinine. Org. Lett. 2:3039 - 3042.

[30]

O'Reilly, E., Iglesias, C., Ghislieri, D., Hopwood, J., Galman, J. L., Lloyd, R. C. and Turner, N. J. (2014) A regio‐and stereoselective ω‐transaminase/monoamine oxidase cascade for the synthesis of chiral 2, 5‐disubstituted pyrrolidines. Angew. Chem. Int. Ed. 53: 2447-2450

[31]

Simon, R. C., Grischek, B., Zepeck, F., Steinreiber, A., Belaj, F. and Kroutil. W (2012) Regio- and stereoselective monoamination of diketones without protecting groups. Angew. Chem. Int. Ed. 51:6713-6716.

[32]

O’Reilly, E., Iglesias, C. and Turner, N.J. (2014) Monoamine oxidase–ω- transaminase cascade for the deracemisation and dealkylation of amines ChemCatChem, DOI: 10.1002/cctc.201300990.

[33]

O'Reilly, E., Balducci, D. and Paradisi, F. (2010) A stereoselective synthesis of αdeuterium-labelled (S)-α-amino acids. Amino Acids 39:849-858.

[34]

Chaudhari, S. R., and Suryaprakash, N. (2012) Simple and efficient methods for discrimination of chiral diacids and chiral alpha-methyl amines. Org. Biomol. Chem. 10:6410-6419.

[35]

O'Reilly, E., Lestini, E., Balducci, D. and Paradisi, F. (2009) One-step diketopiperazine synthesis using phase transfer catalysis. Tetrahedron Lett. 50:1748-50.

[36]

O'Reilly, E., Pes, L. and Paradisi, F. (2010) From amines to diketopiperazines: a one-pot approach Tetrahedron Lett. 51, 1696-1697.

[37]

Koszelewski, D., Clay, D., Rozzell, D. and Kroutil W. (2009) Deracemisation of αchiral primary amines by a one-pot, two-step cascade reaction catalysed by ωtransaminases. Eur. J. Org. Chem. 2289-2292.