Engineering of Cyclohexanone Monooxygenase for the

Feb 27, 2019 - One-Pot Synthesis of Phenylglyoxylic Acid from Racemic Mandelic Acids via Cascade Biocatalysis. Journal of Agricultural and Food Chemis...
8 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF TEXAS DALLAS

Article

Engineering of cyclohexanone monooxygenase for the enantioselective synthesis of S-omeprazole Yan Zhang, Yin-Qi Wu, Na Xu, Qian Zhao, Hui-Lei Yu, and Jian-He Xu ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.9b00224 • Publication Date (Web): 27 Feb 2019 Downloaded from http://pubs.acs.org on March 3, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Manuscript ID: sc-2019-00224k

Engineering of cyclohexanone monooxygenase for the enantioselective synthesis of S-omeprazole Yan Zhang,a Yin-Qi Wu,a Na Xu,a Qian Zhao,b Hui-Lei Yu,a* Jian-He Xua aState

Key Laboratory of Bioreactor Engineering and Shanghai Collaborative

Innovation Center for Biomanufacturing, East China University of Science and Technology, 130 Meilong Road, Shanghai 200237, China; bJiangsu Key Laboratory of Chiral Drug Development, Jiangsu Aosaikang Pharmaceutical Co., Ltd., 766 Kening Road, Nanjing 211112, China;

*Corresponding

author: E mail: [email protected]

State Key Laboratory of Bioreactor Engineering, East China University of Science and Technology, 130 Meilong Road, Shanghai 200237, China, Tel. +86-21-6425 2498; Fax. +86-21-6425 0840.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT: Enzymatic asymmetric sulfoxidation using molecular oxygen as oxidant is a promising green chemistry approach to chiral sulfoxide production. Despite the broad substrate spectrum of cyclohexanone monooxygenases (CHMOs), some unnatural substrates with bulky functional groups, such as the pharmaceutically relevant omeprazole sulfide, cannot be effectively accepted by CHMOs. Herein, we describe a set of variants derived from an Acinetobacter calcoaceticus CHMO (AcCHMO), whose active sites adjacent to the substrate tunnel were altered to shift the substrate specificity from cyclohexanone monooxygenation toward omeprazole sulfide sulfoxidation. We performed homologous modeling and molecular docking to identify key residues that might affect the substrate specificity. Two libraries of residues lining the active center of AcCHMO were then constructed and screened by an effective halo-based selection method using the solubility difference between the substrate (omeprazole sulfide) and product (esomeprazole). Functional evaluation of the resultant variants showed that the substrate specificity of AcCHMO was markedly altered from the small natural substrate (cyclohexanone) toward the desired bulky substrate (omeprazole sulfide) despite the extremely poor activity detected even for the best variant, M2 (0.61 U/gprot). The crystal structure of M2 complexed with a flavin adenine dinucleotide (FAD) prosthetic group was determined, which provided insight into the altered substrate specificity. To improve the activity of enzyme M2 toward pharmaceutical precursor omeprazole sulfide, we performed both local and global protein engineering among the two CASTing libraries surrounding FAD+ and NADP+ prosthetic groups, and an error-prone PCR library of the full-length

ACS Paragon Plus Environment

Page 2 of 36

Page 3 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

AcCHMO. As a result, variant M6 was obtained, giving a 50-fold higher activity than M2. This structure-guided protein engineering of AcCHMO provided a promising candidate for converting omeprazole sulfide into (S)-omeprazole using a green biocatalytic method.

KEYWORDS: Biocatalysis; Cyclohexanone monooxygenase; Substrate specificity; Protein engineering; Esomeprazole sulfoxide; Halo-based selection method

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 36

INTRODUCTION Chiral sulfoxides have extensive and important applications, not only as chiral auxiliary reagents and chiral ligands, but also in pharmaceuticals, such as prazoles, which are proton-pump inhibitors (PPIs, for gastroesophageal reflux treatment; peptic ulcer disease treatment; even Zllinger-Ellison syndrome treatment).1 To date, the synthesis of chiral sulfoxides has mainly relied on the transition-metal-catalyzed asymmetric oxidation of sulfides,2 such as the synthesis of esomeprazole, the S-enantiomer of omeprazole, which requires large amounts of titanium catalyst and diethyl tartrate.3 Furthermore, the downstream removal of sulfones as over-oxidation byproducts requires tedious operations and has a large environmental impact. Enzymatic asymmetric sulfoxidation using molecular oxygen as the oxidant is considered a promising green chemistry approach to obtaining chiral sulfoxides. Type-I Baeyer–Villiger monooxygenases (BVMOs), which utilize O2 as the electron acceptor, FAD (flavin adenine dinucleotide) as the cofactor, and NADPH (nicotinamide adenine dinucleotide phosphate) as the electron donor, can catalyze the Baeyer–Villiger

oxidation

reaction.

Among

BVMOs,

cyclohexanone

monooxygenases (CHMOs), which are class B flavoprotein monooxygenases, are named according to their natural substrate, cyclohexanone.4,5 Apart from their role in nature, CHMOs can also oxidize over 100 non-natural substrates with excellent regio-, chemo-, or stereoselectivity, including the sulfoxidation and oxidation of nitrogen, boron, and selenium compounds, and epoxidation of alkenes.6–12 However, some non-natural substrates cannot be accepted by CHMOs, such as omeprazole

ACS Paragon Plus Environment

Page 5 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

sulfide, which is an important precursor in esomeprazole production (Scheme 1). The biooxidation of prazole sulfides has been explored by several groups, including whole-cell oxidation of rabeprazole and omeprazole sulfides using Cunninghamella echinulata MK40 and Lysinibacillus sp. B71.13,14 Where rabeprazole and omeprazole were accumulated up to 2.5 g/L and 0.115 g/L, respectively, with high enantiomeric excesses (ee) (>99%, (S)-enantiomer). Recently, we discovered two novel

Type-I

BVMOs,

BoBVMO

and

AmBVMO,

from

Bradyrhizobium

oligotrophicum and Aeromicrobium marinum, respectively, that catalyze sulfides oxidation. Notably, these two BVMOs catalyzed the sulfoxidations of bulky prazole sulfides,

including

omeprazole,

ilaprazole,

rabeprazole,

pantoprazole,

and

lansoprazole sulfides.15 However, the opposite product configuration to that desired was obtained when omeprazole sulfide was used as substrate. Furthermore, the poor thermostability and soluble expression levels limited their further application. In addition to discovering new strains or enzymes, structure-guided protein engineering is a powerful tool for extending the scope of substrate acceptance, and enhancing the thermostability and stereoselectivity of BVMOs. To gain insight into the catalytic mechanism and structure–function relationships, and provide guidance for protein engineering, many scientists have committed to solving the crystal structures of BVMOs. However, very few crystal structures of BVMOs are available at present. The first solved BVMO crystal structure was that of phenylacetone monooxygenase (PAMO), which originates from moderately thermophilic bacterium Thermobifida fusca, combined with a FAD prosthetic group in the structure.16 In

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

contrast to CHMOs, PAMO is a thermally robust enzyme with a relatively narrow range of substrate acceptance.17 Therefore, the solved crystal structure of PAMO has been exploited for the structure-inspired enzyme redesign of PAMO. Through fragment assembly of PAMO and CHMONCIMB9871, the two enzymes were successfully switched into phenylcyclohexanone monooxygenase (PCHMO), which resulted in broadened substrate acceptance while maintaining the thermostability.18 Subsequently, the first crystal structure of a CHMO from Rhodococcus sp. strain HI-31 (RmCHMO) was published in 2009, which provided mechanistic insight into the CHMO-catalyzed Baeyer–Villiger oxidation.19,20 As the sequence identity of RmCHMO and CHMONCIMB9871 was 57%, the solved crystal structure of RmCHMO provided an important template for the homology modeling of CHMONCIMB9871. Accordingly, other protein engineering approaches have been used to alter the substrate specificity, increase or invert the enantioselectivity, and improve the thermostability of BVMOs.21–35 These reports have inspired us to evolve BVMOs as catalysts for the oxidation of pharmaceutically relevant bulky sulfides through structure-guided protein engineering. Meanwhile, a high-throughput screening method is necessary to improve the screening efficiency of mutant libraries. The common method used to determine BVMO activity involves measuring NAD(P)H consumption, which does not suit the very low activity of BVMOs. The other methods for screening BVMO mutant libraries depend on either HPLC/GC analyses or coupling with other reactions indirectly.36–38 The engineered variants of CHMONCIMB9871 disclosed in a Codexis patent

ACS Paragon Plus Environment

Page 6 of 36

Page 7 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

showed the desired enantioselectivity for oxidizing prochiral omeprazole sulfide into (S)-omeprazole.39,40 However, its structure–function relationships are unknown. Herein, we altered the substrate specificity and improved the catalytic activity of a CHMO from Acinetobacter calcoaceticus (AcCHMO, with 70% identity to CHMONCIMB9871) through semi-rational protein engineering. We divided the enzyme structure into two regions, namely, the core region, including the FAD binding sites, NADPH binding sites, and substrate tunnel, and the surface region (Fig. 1). Different mutagenesis methods were adopted for these two regions. Furthermore, the crystal structure of variant M2 was solved by X-ray diffraction at a resolution of 2.2 Å, showing that the structural changes might be responsible for the dramatically varied substrate specificity. Furthermore, we invented an effective plate assay method based on the solubility difference between the substrate (omeprazole sulfide) and product (esomeprazole sulfoxide), where the mutant activity was directly correlated to the sizes of transparent halos on Luria–Bertani (LB) agar plates.

EXPERIMENTAL SECTION Chemicals and Materials. All prazole sulfides, sulfoxides, and sulfones were provided by Aosaikang Pharmaceutical Co., Nanjing, China. All other commercial chemicals were purchased from TCI, Macklin, Aladdin, or Sigma-Aldrich. PrimerSTAR HS and rTaq polymerase, restriction enzymes (Dpn I, Nde I, and Hind III), and T4 DNA ligase were purchased from TaKaRa Biotechnology Co., Dalian, China. Primers were synthesized by Generay Biotech Co., Shanghai, China.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 36

Homology Modelling and Molecular Docking. Homology modelling of AcCHMO was performed using Modeller 9.15 software based on the crystal structures of FAD- and NADPH-dependent TmCHMO from Thermocrispum municipale (protein databank (PDB) ID: 5M10) or RmCHMO from Rhodococcus sp. strain HI-31 (PDB ID: 3UCL) with a sequence identity of 58%. The stereochemistry of the model was evaluated using a Ramachandran plot, and the percentage of residues in the allowed region was 97.4%. Molecular docking was performed with AutoDock Vina using the default program parameters. The center coordinates of the grid box were calculated by visual molecular dynamics (VMD), and the size of the grid box was set as 24 Å in each dimension. The docking results were selected according to their binding affinities and molecule conformations.41–43 CAVER 3.0 was used to identify the tunnels existing in AcCHMO (http://www.caver.cz).44 Mutant

Library

Construction.

Site-directed

mutagenesis

(SDM)

and

combinatorial active site saturation testing (CASTing) were introduced by PCR into the pET28a-AcCHMO template DNA (either WT or variants) using the QuikChange site-directed mutagenesis protocol.45 The PCR mixture (50 L) contained plasmid (Template, 50 ng), 1.25 U PrimerSTAR HS, and both the forward and reverse primers. The reaction mixture was preheated at 98 °C for 10 s, annealed at 55 °C for 30 s, and elongated at 72 °C for 7 min, which was repeated for 30 cycles. The PCR product was digested with 0.5 U Dpn I at 37 °C for 1 h. Plasmids containing the mutated gene were transformed into E. coli BL21(DE3) host cells and then plated on an LB agar plate with 50 g/mL kanamycin. The sequence of each variant was

ACS Paragon Plus Environment

Page 9 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

confirmed by DNA sequencing at Saiyin Biotechnology Co., Shanghai, China. Error-prone PCR (epPCR) was conducted using plasmid pET28a-M5 as the template and 50 M Mn2+ was selected for the desired mutagenesis rate (1–3 mutation sites per gene based on sequence analysis of 20 samples). The PCR product was digested with Nde I and Hind III restriction enzymes. The digested product was then ligated into the corresponding sites of pET28a. The recombinant plasmid was then transformed into E. coli BL21 (DE3) cells. The primers used in this study are listed in Table S1. Mutant Library Screening. A stepwise three-round screening strategy was adopted. Primary screening was conducted by inoculation onto LB agar plates supplemented with kanamycin (50 g/mL), isopropyl--ᴅ-thiogalactopyranoside (IPTG, 0.1 mM), omeprazole sulfide (2 mM), and cosolvent DMSO (1%, v/v). After incubation at 30 °C for 12 h, the potentially positive clones with visible halos (larger than the control) produced by enzymatic monooxygenation of the less water-soluble omeprazole sulfide were selected and further cultured in LB liquid medium (4 mL) for secondary screening. The harvested cells were resuspended in 100 mM potassium phosphate buffer (KPB, pH 9.0) containing 1 mM omeprazole sulfide and 1 mM NADPH (500 L total volume). After the reacting at 30 °C for 1 h, the conversion and ee were determined by HPLC, as described previously.15 The specific activity of positive variants with enhanced conversion rates was further measured using purified enzymes as the third round of screening.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Protein Expression and Purification. The E. coli cells expressing wild-type and variants of AcCHMO were cultivated in LB medium at 37 °C and 180 rpm. When the optical density at 600 nm (OD600) reached 0.6–0.8, IPTG was added to a final concentration of 0.2 mM and the cultivation temperature was decreased to 16 °C for protein overexpression. After cultivation for 12 h, the cells were harvested by centrifuging at 5,000 g for 10 min, and the pellets were washed twice with ice-cooled 100 mM KPB (pH 9.0). The resultant cells were resuspended in the same buffer and disrupted by ultrasonication. The cell lysates were centrifuged at 12,000 g at 4 °C for 30 min to remove the particulate fraction. The supernatant was loaded onto a HisTrap Ni-NTA FF column (5 mL, GE Healthcare) pre-equilibrated with buffer A (50 mM KPB, 500 mM NaCl, 10 mM imidazole, pH 8.0). The protein was eluted using an increasing gradient of imidazole, from 10 to 150 mM, at a flow rate of 5 mL/min. The resulting pure protein was collected and concentrated by ultrafiltration. The purified proteins from Ni2+ affinity chromatography were further polished through gel-filtration chromatography (SuperoseTM 12 10/300 GL, GE Healthcare) (Fig. S1). The fractions were determined by SDS-PAGE (Fig. S2). The peak fractions were concentrated to 14 mg/mL and used for crystallization. Crystallization and Diffraction Data Collection. Crystals were grown at 18 °C using the sitting-drop vapor diffusion method 46 by mixing protein solution (2 L, 14 mg/mL) with reservoir solution (2 L) containing 25% (w/v) PEG 3350, 0.1 M Tris-HCl (pH 8.5), and 0.2 M NaCl. Crystals were cryoprotected with reservoir solution plus 15% glycerol. X-ray diffraction data were collected using beamline

ACS Paragon Plus Environment

Page 10 of 36

Page 11 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

BL19U2

47

at the National Center for Protein Science (Shanghai, China). Native and

derivative datasets were processed using XDS

48

and HKL3000 software

49,

respectively. Structure Determination and Refinement. The initial diffraction phases were determined by a molecular replacement method using the crystal structure of RmCHMO from Rhodococcus sp. HI-31 (PDB ID: 3UCL) as a search model. Phaser software 50 was used for single-wavelength anomalous diffraction (SAD) phasing and PHENIX. Autobuild

51

was used to build the initial model. The initial model and

phase information were then transferred to the native data set. Other parts of the structure were manually built in COOT

52

and iteratively refined using

PHENIX.Refine.53 In the final model, more than 97.0% of residues fell in the favored region in the Ramachandran plot and the final Rwork/Rfree values were 0.1848/0.2290, as shown in Table S4. The PDB coordinate was deposited with entry ID 6A37. Specific Activity and Enantioselectivity Assays. The specific activity and stereoselectivity of AcCHMO variants toward cyclohexanone (1a), thioanisole (1b), 5-methoxy-2-methylthio-1H-benzimidazole (1c), and omeprazole sulfide (1d) was measured using purified enzymes. In a 500-L reaction system, 2 mM substrate 1a or 1b (0.2 mM for 1c or 1d), 2 mM NADPH for 1a or 1b (0.2 mM NADPH for 1c or 1d), and diluted enzyme were mixed in KPB (100 mM, pH 9.0). After the reaction was performed at 30 °C for 10 min with mixing at 1000 rpm in a Thermomixer (Eppendorf, Germany), the reaction mixture was extracted with an equal volume of ethyl acetate. The specific activity and enantioselectivity were determined by GC or

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

HPLC, as described elsewhere.15 Enzymatic Sulfoxidation of Omeprazole Sulfide to Producing Esomeprazole. The 10-mL reaction mixture was composed of 3 g/L omeprazole sulfide as substrate (with 10% (v/v) DMSO), ᴅ-glucose (1.5 equiv.), purified enzyme M5 or M6 (1 g/L), glucose dehydrogenase (GDH, 2 g/L), NADP+ (0.2 mM), and potassium phosphate buffer (100 mM, pH 9.0), and was shaken at 180 rpm and 25 °C. Samples were intermittently removed and extracted to analyze the conversion rate by HPLC.

RESULTS AND DISCUSSION Construction of a Halo-based Plate Assay Method. To improve the efficiency of mutant library screening, we designed and constructed an effective halo-based selection method, according to our previous experience with an esterase assay.54 The medium components for the plate assay were first optimized in terms of substrate concentration, cosolvent ratio, and IPTG addition. As a preliminary experimental result, when 2 mM omeprazole sulfide and 1% (v/v) DMSO were added, the transparent LB agar plate turned milky white, as shown in Fig. S3. When no IPTG was added, the colonies grew well, but without any transparent halos on the plate (Fig. S3A). In contrast, when 0.1 mM IPTG was added, obvious transparent halos appeared, but the colonies grew slower (Fig. S3B). The real mutant library screening plate is shown in Fig. S3C. Focused Mutagenesis in the Core Region of AcCHMO. The residues lining the substrate binding pocket and substrate tunnel were more likely to have a profound

ACS Paragon Plus Environment

Page 12 of 36

Page 13 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

effect on the enzyme activity and substrate specificity. Based on the homology modelling, molecular docking results, and hot sites in the literature, two regions of substrate binding sites, one adjacent to the substrate tunnel (Library A) and the other buried inside the protein (Library B), were selected (Fig. 2A). Two other regions were also considered (Fig. 2B), covering amino acid residues surrounding the FAD prosthetic group (Library C) and NADP+ prosthetic group (Library D). Accounting for the primary sequence and three-dimensional structure simultaneously, we attempted protein engineering based on combinatorial active-site saturation testing (CASTing) and iterative saturation mutagenesis (ISM), as proposed by Reetz et al.55– 57

Nineteen residues in libraries A and B were divided into eleven groups, considering

the synergistic effect between them. While twenty-two residues in libraries C and D were also divided into eleven groups, using the same approach. The group details are shown in Table S2, and the degeneracy NNK or NDT codons (N = A, G, C, T; K = G, T;

D

=

A,

G,

T;

NDT

involves

12

codons

for

12

amino

acids

(N/S/I/D/G/V/Y/C/F/H/R/L) and NNK involves 32 codons for 20 amino acids) were used to construct the locally focused mutagenesis libraries of AcCHMO. The mutant libraries were rapidly screened using the halo-based plate assay, as described in the Experimental Section. The screening results of mutagenesis libraries are listed in Table 1. In the first round, a double mutation variant, K326C/F432L (M1), was identified, showing new trace activity towards omeprazole sulfide. Subsequently, using M1 as the parental enzyme, we performed ISM repeatedly, resulting in an octuple mutation variant (M2) that exhibited a specific activity of 0.61

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 36

U/gprotein toward omeprazole sulfide (Table 1). The CASTing experiments of Library C and Library D were also implemented using M2 as parent. As a result, two variants, M3 and M4, showed nearly six-fold greater omeprazole sulfide sulfoxidation activity than that of parental M2. The simple combination of M3 and M4 created a better variant, M5, which presented a near-20-fold improvement compared with M2, suggesting that a strong synergistic effect might exist between M3 and M4. Error-prone

PCR

Mutagenesis

of

Full-length

AcCHMOM5.

Random

mutagenesis can be used to effectively identify new hot sites. Therefore, error-prone PCR was employed in the second round of mutagenesis using M5 as a template to generate a random library. After screening approximately 5,000 clones and confirming the specific activity, eight variants showing specific activity improvements of more than 1.3-fold were identified (Table S3). Among them, the best variant was M6 (M5L143P/K269E), which showed 2.6-fold higher activity compared with M5 (Table 1). Although these improvements were not remarkable, many new mutation sites beneficial for activity improvements were identified in this round. Two of the mutation sites located just within the core region (L143, G185) were newly identified despite being missed in the first round of semi-rational mutagenesis. The remaining sites were situated on the surface region (Fig. S4). Furthermore, to achieve a combinatorial improvement in catalytic efficiency, we also performed DNA shuffling

58,59

using the genes found in the epPCR library. Unfortunately, no superior

variants were obtained after a screening of 3,000 colonies in total. Crystal Structure and Substrate Specificity Analysis of AcCHMO Variants. To

ACS Paragon Plus Environment

Page 15 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

elucidate the mechanism responsible for substrate specificity changes in AcCHMO variants, we attempted to solve the crystal structures of AcCHMOs WT, M1, M2, and M5. However, only the crystal structure of AcCHMOM2 complexed with an FAD prosthetic group was successfully resolved in the space group P21 at a 2.2 Å resolution using the molecular replacement method (Table S4). This structure contained two Rossmann folds and exhibited typical BVMO domain organization with an FAD domain (residues 1–141 and 387–542), an NADP domain (residues 152–210 and 334–380), and a helical domain (residues 224–333) (Fig. S5A). CHMOs are known to have two forms, namely, CHMOopen and CHMOclosed structures. The closest structures to AcCHMO were those reported for TmCHMO and RmCHMO (both with 58% sequence identity). When compared with the CHMOopen (PDB ID: 3GWF) and CHMOclosed (PDB ID: 3GWD) forms of RmCHMO, the resulting root-mean-square deviation (rmsd) values were 0.572 and 0.825 Å, respectively. This indicated that the structure of AcCHMOM2 was much closer to the “open” form, which was consistent with the real structure in a non-catalytic state. The location of the FAD prosthetic group was also consistent with that of RmCHMO (Fig. S5B). To gain insight into the origin of the dramatically altered substrate specificity, the structures of AcCHMOM2 (PDB ID: 6A37) and AcCHMOWT (modelled based on the crystal structure of AcCHMOM2) were compared. As shown in Fig. 3, four mutation sites replaced by smaller amino acids (K326C, F432L, T433A, and L435S) in AcCHMOM2 occurred at locations close to the substrate tunnel causing the substrate tunnel to significantly broaden, which was beneficial for accepting sterically bulky

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

substrates. The newly formed interaction between hydroxy Y246 and the sulfurous substrate aided substrate binding in the catalytic pocket. The three mutation sites (E488K, S489C, and W490R) of M4 were either surrounding the NADP+ prosthetic group or located on a flexible loop (residues 487–504), which has been demonstrated to have interactions with the NADP+ molecule.60 Here, three additional hydrogen bonds formed between the three mutation sites and NADP+, resulting in an enzyme with drastically increased activity (Fig. 4). Similarly, the dramatically improved activity of variant M3 was attributed to two additional hydrogen bonds formed between FAD and the newly created residues (Fig. 4), which were beneficial for FAD binding. Through two rounds of directed evolution, the catalytic activity of AcCHMO variants specific to pharmaceutically relevant omeprazole sulfide increased gradually, as shown in Fig. 5. The combinatorial multiple mutation variant M6 showed 5,611-fold higher activity compared with variant M1. To explore the evolutionary footprints of AcCHMO variants with dramatically altered substrate specificity, four distinct substrates with varying sizes and structures, including cyclohexanone (1a), thioanisole (1b), 5-methoxy-2-methylthio-1H-benzimidazole (1c), and omeprazole sulfide (1d), were selected to fingerprint the activities of AcCHMOWT and six variants. Interestingly, compared with the bulky omeprazole sulfide, the catalytic activity of the variants toward small substrates cyclohexanone and thioanisole decreased obviously during the evolutionary course. In particular, for native substrate cyclohexanone, variant M2 had lost its natural activity completely. However, the activity of variants

ACS Paragon Plus Environment

Page 16 of 36

Page 17 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

toward substrate 1c initially increased and then decreased later. These results showed an evolutionary route of AcCHMO, demonstrating that the substrate specificity was altered stepwise from the natural small substrate (cyclohexanone) to the pharmaceutically important bulky omeprazole sulfide. Catalytic Performance of AcCHMO Variants for (S)-omeprazole Synthesis. The engineered AcCHMOs were capable of catalyzing the asymmetric sulfoxidation of prochiral omeprazole sulfide into pharmacologically active (S)-omeprazole, accompanied by the consumption of NADPH to form NADP+. Therefore, a glucose dehydrogenase from Bacillus megaterium was introduced to regenerated the desired cofactor NADPH from NADP+ using glucose as a cheap cosubstrate. The enzymatic sulfoxidation of omeprazole sulfide (3 g/L) was conducted comparatively using the best two variants, M5 and M6, using a 1 g/L dose of each purified enzyme under the same conditions. As shown in Fig. 6, the reaction catalyzed by variant M6 achieved >95% conversion after 22 h. Meanwhile, variant M5 reached only 14% conversion after 22 h. In contrast to the chemical oxidation of omeprazole sulfide using metal catalysts and peracids or hydrogen peroxide, this enzymatic process used oxygen (air) and glucose as sustainable cosubstrates, affording water as a clean byproduct, making this method a promising green chemistry approach to esomeprazole production.

CONCLUSIONS In summary, we applied structure-guided rational and random strategies for

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

evolving a putative cyclohexanone monooxygenase to activate its potential activity toward the stereoselective sulfoxidation of bulky omeprazole sulfide. An effective halo-based plate assay method was designed and constructed based on the solubility difference between the substrate (omeprazole sulfide) and product (esomeprazole). Compared with the common HPLC/GC assays or microtiter-based screening methods of monooxygenases, which monitor the NADPH absorbance at 340 nm, the plate assays significantly improved the screening efficiency of mutant libraries. As the solubilities of sulfoxides and sulfones are generally higher than those of the corresponding sulfides, this halo-based plate assay method could be applied extensively. Consequently, a 5,611-fold improved variant (M6) was hit after two rounds of evolution, including a combination of CASTing and ISM of the core sites and epPCR mutation of the full-length protein. The footprints on the AcCHMO evolution road were explored with migratory substrate specificity toward four different substrates with varying sizes and structures. The substrate specificity of AcCHMOs was shifted from the small and natural substrate (cyclohexanone) to the bulky and unnatural omeprazole sulfide. We also determined the crystal structure of AcCHMOM2 complexed with an FAD prosthetic group, and analyzed the structural transformation between AcCHMOWT and its variants. In the majority of flavoenzymes, the cofactor is tightly but noncovalently bound.61 However, in our crystallization experiments, even when no external FAD was added, the FAD prosthetic group was still observed in the crystal structure of AcCHMOM2. Furthermore, activity experiments showed that a less than two-fold

ACS Paragon Plus Environment

Page 18 of 36

Page 19 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

increase was observed even when saturated FAD was added. This was consistent with the literature (cofactor is tightly but noncovalently bound) described above. The resulting best variant (M6) converted omeprazole sulfide into (S)-omeprazole with greatly enhanced efficiency, making it a promising candidate for practical applications. Further analysis and reaction optimization studies are ongoing to better elucidate the unknown mechanism behind the substrate specificity shifts of CHMOs and to promote their industrial application.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Table S1-S4 and Figure S1-S7. AUTHOR INFORMATION Corresponding Author H-L Yu. Email: [email protected] Notes The authors declare no competing financial interest.

ACKNOWLEDGMENTS The work was financially supported by the National Natural Science Foundation of

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

China (Nos. 21536004, 21672063 and 21871085) and the Fundamental Research Funds for the Central Universities (No. 22221818014). We thank Yue-Peng Shang and Feng Liu at East China University of Science and Technology for their insightful discussions. We thank Simon Partridge, PhD, from Liwen Bianji, Edanz Editing China (www.liwenbianji.cn/ac), for editing the English text of a draft of this manuscript.

ACS Paragon Plus Environment

Page 20 of 36

Page 21 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

REFERENCES 1. Fernández, I.; Khiar, N. Recent developments in the synthesis and utilization of chiral sulfoxides. Chem. Rev. 2003, 103 (9), 3651-3706, DOI 10.1002/chin.200348251. 2. Talsi, E. P.; Rybalova, T. V.; Bryliakov, K. P. Isoinversion behavior in the enantioselective oxidations of pyridylmethylthiobenzimidazoles to chiral proton pump inhibitors on titanium salalen complexes. ACS Catalysis. 2015, 5 (8), 4673-4679, DOI 10.1021/acscatal.5b01212. 3. Cotton, H.; Elebring, T.; Larsson, M.; Li, L.; Sörensen, H.; Unge, S. Asymmetric synthesis of esomeprazole. Tetrahedron: Asymmetry. 2000, 11: 3819-3825, DOI 10.1016/S0957-4166(00)00352-9. 4. van Berkel, W. J.; Kamerbeek, N. M.; Fraaije, M. W. Flavoprotein monooxygenases, a diverse class of oxidative biocatalysts. J. Biotechnol. 2006, 124 (4), 670-689, DOI 10.1016/j.jbiotec.2006.03.044. 5. Riebel, A.; Fink, M. J.; Mihovilovic, M. D.; Fraaije, M. W. Type II flavin-containing monooxygenases: A new class of biocatalysts that harbors Baeyer-Villiger monooxygenases with a relaxed coenzyme specificity. ChemCatChem 2014, 6 (4), 1112-1117, DOI 10.1002/cctc.201300550. 6. Stewart, J. D. Cyclohexanone monooxygenase: a useful reagent for asymmetric Baeyer-Villiger reactions. Curr. Org. Chem. 1998, 2 (3), 195-216, DOI 10.1002/chin.199844320. 7. Ottolina, G.; Pasta, P.; Carrea, G.; Colonna, S.; Dallavalle, S.; Holland, H. L. A predictive active site model for the cyclohexanone monooxygenase catalyzed oxidation of sulfides to chiral sulfoxides. Tetrahedron: Asymmetry 1995, 6 (6), 1375-1386, DOI 10.1016/0957-4166(95)00170-T. 8. Stefano Colonna, N. G.; Carrea G.; Ottolina G.; Pasta P.; Zambianchi, F. First asymmetric epoxidation catalysed by cyclohexanone monooxygenase. Tetrahedron Letters 2002, 43, 1797-1799, DOI 10.1016/S0040-4039(02)00029-1. 9. Zambianchi, F.; Pasta, P.; Carrea, G.; Colonna, S.; Gaggero, N.; Woodley, J. M. Use of isolated cyclohexanone monooxygenase from recombinant Escherichia coli as a biocatalyst for Baeyer-Villiger and sulfide oxidations. Biotechnol. Bioeng. 2002, 78 (5), 489-496, DOI 10.1002/bit.10207. 10. Branchaud, B. P.; Walsh, C. T. Functional group diversity in enzymic oxygenation reactions catalyzed by bacterial flavin-containing cyclohexanone oxygenase. J. Am. Chem. Soc. 1985, 107 (7), 2153-2161, DOI 10.1002/chin.198533141. 11. de Gonzalo, G.; Mihovilovic, M. D.; Fraaije, M. W. Recent developments in the application of Baeyer-Villiger monooxygenases as biocatalysts. Chembiochem 2010, 11 (16), 2208-2231, DOI 10.1002/cbic.201000395. 12. Leisch, H.; Morley, K.; Lau, P. C. K. Baeyer-Villiger monooxygenases: more than just green chemistry. Chem. Rev. 2011, 111 (7), 4165-4222, DOI 10.1021/cr1003437. 13. Yoshida T.; Kito M.; Tsujii M.; Nagasawa T. Microbial synthesis of a proton pump inhibitor by enantioselective oxidation of a sulfide into its corresponding sulfoxide by Cunninghamella echinulata MK40. Biotechnol. Lett. 2001, 23(15), 1217-1222, DOI 10.1023/a:1010521217954. 14. Babiak, P.; Kyslikova, E.; Stepanek, V.; Valesova, R.; Palyzova, A.; Maresova, H.; Hajicek, J.; Kyslik, P. Whole-cell oxidation of omeprazole sulfide to enantiopure esomeprazole with Lysinibacillus sp. B71. Bioresour. Technol. 2011, 102 (17), 7621-7626, DOI

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10.1016/j.biortech.2011.05.052. 15. Zhang, Y.; Liu, F.; Xu, N.; Wu, Y. Q.; Zheng, Y. C.; Zhao, Q.; Lin, G. Q.;Yu, H. L.; Xu, J. H. Discovery of two native Baeyer-Villiger monooxygenases for asymmetric synthesis of bulky chiral sulfoxides. Appl. Environ. Microbiol. 2018, 84 (14), e00638-18, DOI 10.1128/AEM.00638-18. 16. Malito, E.; Alfieri, A.; Fraaije, M. W.; Mattevi, A. Crystal structure of a Baeyer-Villiger monooxygenase. Proc. Natl. Acad. Sci. U. S. A. 2004, 101 (36), 13157-13162, DOI 10.1073/pnas.0404538101. 17. Fraaije, M. W.; Wu, J.; Heuts, D. P. H. M.; van, H. E. W.; Spelberg, J. H. L.; Janssen, D. B. Discovery of a thermostable Baeyer-Villiger monooxygenase by genome mining. Appl. Microbiol. Biotechnol. 2005, 66 (4), 393-400, DOI 10.1007/s00253-004-1749-5. 18. Bocola, M.; Schulz, F.; Leca, F.; Vogel, A.; Fraaije, M. W.; Reetz, M. T. Converting phenylacetone monooxygenase into phenylcyclohexanone monooxygenase by rational design: Towards practical Baeyer-Villiger monooxygenases. Adv. Synth. Catal. 2005, 347 (7-8), 979-986, DOI 10.1002/adsc.200505069. 19. Mirza, I. A.; Yachnin, B. J.; Wang, S.; Grosse, S.; Bergeron, H.; Imura, A.; Iwaki, H.; Hasegawa, Y.; Lau, P. C. K.; Berghuis, A. M. Crystal structures of cyclohexanone monooxygenase reveal complex domain movements and a sliding cofactor. J. Am. Chem. Soc. 2009, 131 (25), 8848-8854, DOI 10.1021/ja9010578. 20. Beam, M. P.; Bosserman, M. A.; Noinaj, N.; Wehenkel, M.; Rohr, J. Crystal structure of Baeyer-Villiger monooxygenase MtmOIV, the key enzyme of the mithramycin biosynthetic pathway. Biochemistry 2009, 48 (21), 4476-4487, DOI 10.1021/bi8023509. 21. Clouthier, C. M.; Kayser, M. M. Increasing the enantioselectivity of cyclopentanone monooxygenase (CPMO): profile of new CPMO mutants. Tetrahedron: Asymmetry 2006, 17 (18), 2649-2653, DOI 10.1016/j.tetasy.2006.10.001. 22. Clouthier, C. M.; Kayser, M. M.; Reetz, M. T. Designing new Baeyer-Villiger monooxygenases using restricted CASTing. J. Org. Chem. 2006, 71 (22), 8431-8437, DOI 10.1002/chin.200703194. 23. Mihovilovic M. D.; Rudroff F.; Winninger A.; Schneider T.; Schulz F.; Reetz M. T. Microbial Baeyer−Villiger oxidation:  stereopreference and substrate acceptance of cyclohexanone monooxygenase mutants prepared by directed evolution. Org. Lett. 2006, 8 (6), 1221-1224, DOI 10.1021/ol0601040. 24. Pazmino, D. E. T.; Snajdrova, R.; Rial, D. V.; Mihovilovic, M. D.; Fraaije, M. W. Altering the substrate specificity and enantioselectivity of phenylacetone monooxygenase by structure-inspired enzyme redesign. Adv. Synth. Catal. 2007, 349 (8-9), 1361-1368, DOI 10.1002/adsc.200700045. 25. Kirschner, A.; Bornscheuer, U. T. Directed evolution of a Baeyer-Villiger monooxygenase to enhance enantioselectivity. Appl. Microbiol. Biotechnol. 2008, 81 (3), 465-472, DOI 10.1007/s00253-008-1646-4. 26. Reetz, M. T.; Wu, S. Laboratory evolution of robust and enantioselective Baeyer-Villiger monooxygenases for asymmetric catalysis. J. Am. Chem. Soc. 2009, 131 (42), 15424-15432, DOI 10.1021/ja906212k. 27. Zhang, Z. G.; Parra, L. P.; Reetz, M. T. Protein engineering of stereoselective Baeyer-Villiger monooxygenases. Chemistry 2012, 18 (33), 10160-10172, DOI

ACS Paragon Plus Environment

Page 22 of 36

Page 23 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

10.1002/chem.201202163. 28. Bisagni, S.; Summers, B.; Kara, S.; Hatti-Kaul, R.; Grogan, G.; Mamo, G.; Hollmann, F. Exploring the substrate specificity and enantioselectivity of a Baeyer-Villiger monooxygenase from Dietzia sp. D5: oxidation of sulfides and aldehydes. Top. Catal. 2014, 57 (5), 366-375, DOI 10.1007/s11244-013-0192-1. 29. Wang, J. B.; Li, G.; Reetz, M. T. Enzymatic site-selectivity enabled by structure-guided directed evolution. Chem. Commun. 2017, 53 (28), 3916-3928, DOI 10.1039/c7cc00368d. 30. Reetz, M. T.; Brunner, B.; Schneider, T.; Schulz, F.; Clouthier, C. M.; Kayser, M. M. Asymmetric catalysis: Directed evolution as a method to create enantioselective cyclohexanone monooxygenases for catalysis in Baeyer-Villiger reactions. Angew. Chem., Int. Ed. 2004, 43 (31), 4075-4078, DOI 10.1002/anie.200460272. 31. Reetz, M. T.; Daligault, F.; Brunner, B.; Hinrichs, H.; Deege, A. Asymmetric catalysis: Directed evolution of cyclohexanone monooxygenases: Enantioselective biocatalysts for the oxidation of prochiral thioethers. Angew. Chem., Int. Ed. 2004, 43 (31), 4078-4081, DOI 10.1002/anie.200460311. 32. Mihovilovic, M. D. Enzyme mediated Baeyer-Villiger oxidations. Curr. Org. Chem. 2006, 10 (11), 1265-1287, DOI 10.2174/138527206777698002. 33. Kayser, M. M. ‘Designer reagents’ recombinant microorganisms: new and powerful tools for organic synthesis. Tetrahedron 2009, 65 (5), 947-974, DOI 10.1016/j.tet.2008.10.039. 34. Torres Pazmino, D. E.; Dudek, H. M.; Fraaije, M. W. Baeyer-Villiger monooxygenases: Recent advances and future challenges. Curr. Opin. Chem. Biol. 2010, 14 (2), 138-144, DOI 10.1016/j.cbpa.2009.11.017. 35. van Beek, H. L.; Wijma, H. J.; Fromont, L.; Janssen, D. B.; Fraaije, M. W. Stabilization of cyclohexanone monooxygenase by a computationally designed disulfide bond spanning only one residue. FEBS Open Bio. 2014, 4, 168-174, DOI 10.1016/j.fob.2014.01.009. 36. Sicard, R.; Chen, L. S.; Marsaioli, A. J.; Reymond, J. L. A fluorescence-based assay for Baeyer-Villiger monooxygenases, hydroxylases and lactonases. Adv. Synth. Catal. 2005, 347 (7-8), 1041-1050, DOI 10.1002/adsc.200505040. 37. Linares-Pasten, J. A.; Chavez-Lizarraga, G.; Villagomez, R.; Mamo, G.; Hatti-Kaul, R. A method for rapid screening of ketone biotransformations: detection of whole cell Baeyer-Villiger monooxygenase activity. Enzyme Microb. Technol. 2012, 50 (2), 101-106, DOI 10.1016/j.enzmictec.2011.10.004. 38. Sass, S.; Kadow, M.; Geitner, K.; Thompson, M. L.; Talmann, L.; Boettcher, D.; Schmidt, M.; Bornscheuer, U. T. A high-throughput assay method to quantify Baeyer-Villiger monooxygenase activity. Tetrahedron 2012, 68 (37), 7575-7580, DOI 10.1016/j.tet.2012.05.098. 39. Bong, Y. K.; Clay, M. D.; Collier, S. J.; Mijts, B.; Vogel, M.; Zhang, X.; Zhu, J.; Nazor, J.; Smith, D.; Song, S. Engineered cylohexanone monooxygenases for synthesis of prazole compounds. WO2011071982A2, 2011. 40. Bong, Y. K.; Clay, M. D.; Collier, S. J.; Mijts, B.; Vogel, M.; Zhang, X.; Zhu, J.; Nazor, J.; Smith, D.; Song, S. Synthesis of prazole compounds. US8895271B2, 2014. 41. Marti-Renom, M. A.; Stuart, A. C.; Fiser, A.; Sanchez, R.; Melo, F.; Sali, A. Comparative protein structure modeling of genes and genomes. Annu. Rev. Biophys. Biomol. Struct. 2000, 29, 291-325, DOI 10.1146/annurev.biophys.29.1.291. 42. Trott, O.; Olson, A. J. AutoDock Vina: Improving the speed and accuracy of docking with a

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31 (2), 455-461, DOI 10.1002/jcc.21334. 43. Seeliger, D.; Groot, B. L. Ligand docking and binding site analysis with PyMOL and Autodock/Vina. J. Comput.-Aided Mol. Des. 2010, 24 (5), 417-422, DOI 10.1007/s10822-010-9352-6. 44. Chovancova, E.; Pavelka, A.; Benes, P.; Strnad, O.; Brezovsky, J.; Kozlikova, B.; Gora, A.; Sustr, V.; Klvana, M.; Medek, P.; Biedermannova, L.; Sochor, J.; Damborsky, J. CAVER 3.0: a tool for the analysis of transport pathways in dynamic protein structures. PLoS Comput. Biol. 2012, 8 (10), e1002708, DOI 10.1371/journal.pcbi.1002708. 45. Zheng, L.; Baumann, U.; Reymond, J. L. An efficient one-step site-directed and site-saturation mutagenesis protocol. Nucleic Acids Res. 2004, 32 (14), e115, DOI 10.1093/nar/gnh110. 46. Chen, J.; Luo, X. J.; Chen, Q.; Pan, J.; Zhou, J.; Xu, J. H. Marked enhancement of Acinetobacter sp. organophosphorus hydrolase activity by a single residue substitution Ile211Ala. Bioresources and Bioprocessing 2015, 2 (39), DOI 10.1186/s40643-015-0067-3. 47. Li, N.; Li, X.; Wang, Y.; Liu, G.; Zhou, P.; Wu, H.; Hong, C.; Bian, F.; Zhang, R. The new NCPSS BL19U2 beamline at the SSRF for small-angle X-ray scattering from biological macromolecules in solution. J. Appl. Crystallogr. 2016, 49 (5), 1428-1432, DOI 10.1107/S160057671601195X. 48. Kabsch, W. Integration, scaling, space-group assignment and post-refinement. Acta Crystallogr. D Biol. Crystallogr. 2010, 66 (Pt 2), 133-144, DOI 10.1107/S0907444909047374. 49. Otwinowski, Z.; Minor, W. [20] Processing of X-ray diffraction data collected in oscillation mode. Methods in Enzymology 1997, 276, 307-326, DOI 10.1016/S0076-6879(97)76066-X. 50. McCoy, A. J.; Grosse-Kunstleve, R. W.; Adams, P. D.; Winn, M. D.; Storoni, L. C.; Read, R. J. Phaser crystallographic software. J. Appl. Crystallogr. 2007, 40 (Pt 4), 658-674, DOI 10.1107/S0021889807021206. 51. Terwilliger, T. C.; Grosse-Kunstleve, R. W.; Afonine, P. V.; Moriarty, N. W.; Zwart, P. H.; Hung, L. W.; Read, R. J.; Adams, P. D. Iterative model building, structure refinement and density modification with the PHENIX AutoBuild wizard. Acta Crystallogr. D Biol. Crystallogr. 2008, 64 (Pt 1), 61-69, DOI 10.1107/S090744490705024X. 52. Emsley, P.; Lohkamp, B.; Scott, W. G.; Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 2010, 66 (Pt 4), 486-501, DOI 10.1107/S0907444910007493. 53. Afonine, P. V.; Grosse-Kunstleve, R. W.; Echols, N.; Headd, J. J.; Moriarty, N. W.; Mustyakimov, M.; Terwilliger, T. C.; Urzhumtsev, A.; Zwart, P. H.; Adams, P. D. Towards automated crystallographic structure refinement with phenix.refine. Acta Crystallogr. D Biol. Crystallogr. 2012, 68 (Pt 4), 352-367, DOI 10.1107/S0907444912001308. 54. Zhang, Y.; Pan, J.; Luan, Z. J.; Xu, G. C.; Park, S.; Xu, J. H. Cloning and characterization of a novel esterase from Rhodococcus sp. for highly enantioselective synthesis of a chiral cilastatin precursor. Appl. Environ. Microbiol. 2014, 80 (23), 7348-7355, DOI 10.1128/AEM.01597-14. 55. Reetz, M. T. Laboratory evolution of stereoselective enzymes: A prolific source of catalysts for asymmetric reactions. Angew. Chem., Int. Ed. 2011, 50 (1), 138-174, DOI 10.1002/anie.201000826. 56. Acevedo-Rocha, C. G.; Reetz, M. T. Assembly of designed oligonucleotides: a useful tool in synthetic biology for creating high-quality combinatorial DNA libraries. Methods Mol. Biol. 2014,

ACS Paragon Plus Environment

Page 24 of 36

Page 25 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

1179, 189-206, DOI 10.1007/978-1-4939-1053-3_13. 57. Sun, Z.; Wikmark, Y.; Baeckvall, J. E.; Reetz, M. T. New concepts for increasing the efficiency in directed evolution of stereoselective enzymes. Chem. Eur. J. 2016, 22 (15), 5046-5054, DOI 10.1002/chin.201623263. 58. Antikainen, N. M.; Martin, S. F. Altering protein specificity: techniques and applications. Bioorg. Med. Chem. 2005, 13 (8), 2701-2716, DOI 10.1016/j.bmc.2005.01.059. 59. Jackson, S. E.; Craggs, T. D.; Huang, J. R. Understanding the folding of GFP using biophysical techniques. Expert Rev. Proteomics 2006, 3 (5), 545-559, DOI 10.1586/14789450.3.5.545. 60. Yachnin, B. J.; Lau, P. C. K.; Berghuis, A. M. The role of conformational flexibility in Baeyer-Villiger monooxygenase catalysis and structure. Biochim. Biophys. Acta, Proteins Proteomics 2016, 1864 (12), 1641-1648, DOI 10.1016/j.bbapap.2016.08.015. 61. Joosten, V.; van Berkel, W. J. Flavoenzymes. Curr. Opin. Chem. Biol. 2007, 11 (2), 195-202, DOI 10.1016/j.cbpa.2007.01.010.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table Captions Table 1. Properties of the purified AcCHMO variants.a

Figure Legends Scheme 1. Asymmetric oxidation of omeprazole sulfide by BVMOs for the production of S-omeprazole. Figure 1. The enzyme structure is divided into two regions for focused evolution: 1) The core region (FAD binding, NADPH binding and substrate tunnel); and 2) The surface region. CASTing and ISM were employed to mutate the residues in the core region. epPCR was further employed to discover further reinforcing mutations in the surface region. Figure 2. Poses of omeprazole sulfide docked in the modelled structure of AcCHMOWT. The target amino acid residues of Libraries A to D are shown in sticks. The substrate tunnel is shown in surface. The substrate omeprazole sulfide is shown in salmon; FAD prosthetic group in orange; and NADP+ in yellow. Libraries A to D are shown in red, green, marine and cyan, respectively. Figure 3. Structural comparison between AcCHMOWT (A) and AcCHMOM2 (B). The structure of AcCHMOWT was built based on the crystal structure of AcCHMOM2 (PDB ID: 6A37). The amino acid residues, substrate and FAD are shown in sticks. The substrate tunnel is shown in surface. Figure 4. Comparison of hydrogen bonding in AcCHMOWT (A) and its variant AcCHMOM5 (B). The amino acid residues, NADP+ and FAD are shown in sticks. The hydrogen bonds are shown in imaginary lines. Figure 5. Activity fingerprints of AcCHMOWT and its variants (M1 to M6) with different substrates. The activity was measured by GC or HPLC and the relative activity was expressed as a percentage of the maximum activity toward each substrate. Figure 6. Progress curves of omeprazole sulfide sulfoxidation catalyzed by the purified enzymes of two AcCHMO variants: M5 (♦) and M6 (■). The enzymatic sulfoxidation of omeprazole sulfide (3 g/L) was performed at pH 9.0 and 25 °C with a dose of each 1 g/L purified enzymes under the same condition. GDH was used to recycle NADPH.

ACS Paragon Plus Environment

Page 26 of 36

Page 27 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

TOC Synopsis: We applied structure-guided hierarchical strategies and an effective halo-based plate assay method for evolving AcCHMO to alter its substrate specificity from the native substrate cyclohexanone to the pharmaceutically important bulky omeprazole sulfide.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

Page 28 of 36

Table 1. Properties of the purified AcCHMO variants.a Enzyme and targeted area

Mutation sites

Specific activity (U/gprotein)a

% ee (Config.)b

WT, starting enzyme

None

ndc

nd -3

1 (M1), substrate tunnel

K326C/F432Ld

(5.4 ±0.5)  10

29 (S)

2, substrate tunnel

K326F/F432L

nd

nd

3, substrate tunnel

K326C/F432I

(4.4 ±0.4)  10

4, substrate tunnel

K326C/F432L/L435S/S438I

(2.5 ±0.5)  10

5, substrate tunnel

K326C/F432L/T433C/L435S/S438I

(3.2 ±0.8)  10

6, substrate tunnel

K326C/F432L/T433A/L435S/S438I

(4.3 ±0.1)  10

7, substrate tunnel

K326C/L426F/F432L/T433A/L435S/S438I

(9.0 ±0.3)  10

-3

4.9 (S)

-2

57 (S)

-2

84 (S)

-2

83 (S)

-2

69 (S)

-2

8, substrate tunnel

K326C/L426P/F432L/T433A/L435S/S438I

(9.9 ±0.2)  10

16 (S)

9, substrate tunnel

K326C/L426F/F432L/T433A/L435S/S438I/F505L

0.12 ± 0.01

93 (S)

10 (M2), substrate tunnel

F246Y/K326C/L426F/F432L/T433A/L435S/S438I/F505L

0.61 ± 0.02

97 (S)

11 (M3), FAD binding

F246Y/K326C/L426F/F432L/T433A/L435S/S438I/F505L/N386S/I388K/M390I

3.8 ± 0.1

99 (S)

12 (M4), NADPH binding

F246Y/K326C/L426F/F432L/T433A/L435S/S438I/F505L/E488K/S489C/W490R

3.5 ± 0.1

99 (S)

12 ± 1

99 (S)

30 ± 1

99 (S)

13 (M5), simple combination

14 (M6), surface a

F246Y/K326C/L426F/F432L/T433A/L435S/S438I/F505L/N386S/I388K/M390I/E488K/S4 89C/W490R L143P/F246Y/K269E/K326C/L426F/F432L/T433A/L435S/S438I/F505L/N386S/I388K/M 390I/E488K/S489C/W490R

Specific activity was determined at pH 9.0 and 30 °C using purified enzyme and omeprazole sulfide. b ee value was determined by HPLC. c Not detected. d The bold

mutations indicate newly involved mutation(s) in each round.

ACS Paragon Plus Environment

Page 29 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 1. The enzyme structure is divided into two regions for focused evolution: 1) The core region (FAD binding, NADPH binding and substrate tunnel); and 2) The surface region. CASTing and ISM were employed to mutate the residues in the core region. epPCR was further employed to discover further reinforcing mutations in the surface region. 29x16mm (600 x 600 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Poses of omeprazole sulfide docked in the modelled structure of AcCHMOWT. The target amino acid residues of Libraries A to D are shown in sticks. The substrate tunnel is shown in surface. The substrate omeprazole sulfide is shown in salmon; FAD prosthetic group in orange; and NADP+ in yellow. Libraries A to D are shown in red, green, marine and cyan, respectively. 31x15mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 30 of 36

Page 31 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 3. Structural comparison between AcCHMOWT (A) and AcCHMOM2 (B). The structure of AcCHMOWT was built based on the crystal structure of AcCHMOM2 (PDB ID: 6A37). The amino acid residues, substrate and FAD are shown in sticks. The substrate tunnel is shown in surface. 31x15mm (600 x 600 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4. Comparison of hydrogen bonding in AcCHMOWT (A) and its variant AcCHMOM5 (B). The amino acid residues, NADP+ and FAD are shown in sticks. The hydrogen bonds are shown in imaginary lines. 31x15mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 32 of 36

Page 33 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 5. Activity fingerprints of AcCHMOWT and its variants (M1 to M6) with different substrates. The activity was measured by GC or HPLC and the relative activity was expressed as a percentage of the maximum activity toward each substrate. 15x11mm (600 x 600 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6. Progress curves of omeprazole sulfide sulfoxidation catalyzed by the purified enzymes of two AcCHMO variants: M5 (♦) and M6 (■). The enzymatic sulfoxidation of omeprazole sulfide (3 g/L) was performed at pH 9.0 and 25 °C with a dose of each 1 g/L purified enzymes under the same condition. GDH was used to recycle NADPH. 19x11mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 34 of 36

Page 35 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Asymmetric oxidation of omeprazole sulfide by BVMOs for the production of S-omeprazole. 77x16mm (600 x 600 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Synopsis: We applied structure-guided hierarchical strategies and an effective halo-based plate assay method for evolving AcCHMO to alter its substrate specificity from the native substrate cyclohexanone to the pharmaceutically important bulky omeprazole sulfide. 22x13mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 36 of 36