Skip to main content
Applied and Environmental Microbiology logoLink to Applied and Environmental Microbiology
. 2019 May 16;85(11):e00239-19. doi: 10.1128/AEM.00239-19

Efficient Synthesis of Methyl 3-Acetoxypropionate by a Newly Identified Baeyer-Villiger Monooxygenase

Yuan-Yang Liu a, Chun-Xiu Li a, Jian-He Xu a, Gao-Wei Zheng a,
Editor: Volker Müllerb
PMCID: PMC6532031  PMID: 30926727

BVMOs are emerging as a green alternative to traditional oxidants in the BV oxidation of ketones. Although many BVMOs are discovered and used in organic synthesis, few are really applied in industry, especially in the case of aliphatic ketones. Herein, a highly soluble and relatively stable monooxygenase from Rhodococcus pyridinivorans (BVMORp) was identified with high activity and excellent regioselectivity toward most aliphatic ketones. BVMORp possesses unusually high substrate loading during the catalysis of the oxidation of biobased methyl levulinate to 3-hydroxypropionic acid derivatives. This study indicates that the synthesis of 3-hydroxypropionate from readily available biobased levulinate by BVMORp-catalyzed oxidation holds great promise to replace traditional fermentation.

KEYWORDS: 3-hydroxypropionate, Baeyer-Villiger monooxygenase, biocatalysis, ester synthesis, levulinic acid

ABSTRACT

Baeyer-Villiger monooxygenases (BVMOs) are an emerging class of promising biocatalysts for the oxidation of ketones to prepare corresponding esters or lactones. Although many BVMOs have been reported, the development of highly efficient enzymes for use in industrial applications is desirable. In this work, we identified a BVMO from Rhodococcus pyridinivorans (BVMORp) with a high affinity toward aliphatic methyl ketones (Km < 3.0 μM). The enzyme was highly soluble and relatively stable, with a half-life of 23 h at 30°C and pH 7.5. The most effective substrate discovered so far is 2-hexanone (kcat = 2.1 s−1; Km = 1.5 μM). Furthermore, BVMORp exhibited excellent regioselectivity toward most aliphatic ketones, preferentially forming typical (i.e., normal) products. Using the newly identified BVMORp as the catalyst, a high concentration (26.0 g/liter; 200 mM) of methyl levulinate was completely converted to methyl 3-acetoxypropionate after 4 h, with a space-time yield of 5.4 g liter−1 h−1. Thus, BVMORp is a promising biocatalyst for the synthesis of 3-hydroxypropionate from readily available biobased levulinate to replace the conventional fermentation.

IMPORTANCE BVMOs are emerging as a green alternative to traditional oxidants in the BV oxidation of ketones. Although many BVMOs are discovered and used in organic synthesis, few are really applied in industry, especially in the case of aliphatic ketones. Herein, a highly soluble and relatively stable monooxygenase from Rhodococcus pyridinivorans (BVMORp) was identified with high activity and excellent regioselectivity toward most aliphatic ketones. BVMORp possesses unusually high substrate loading during the catalysis of the oxidation of biobased methyl levulinate to 3-hydroxypropionic acid derivatives. This study indicates that the synthesis of 3-hydroxypropionate from readily available biobased levulinate by BVMORp-catalyzed oxidation holds great promise to replace traditional fermentation.

INTRODUCTION

The Baeyer-Villiger (BV) reaction, inserting an oxygen atom between the carbon atom of a carbonyl and the adjacent carbon atom to form an ester or lactone, is one of the most widely used oxidation reactions in organic synthesis. Although BV oxidation was reported by Baeyer and Villiger more than 100 years ago, it is still under continual development. BV monooxygenases (BVMOs) are emerging as a green alternative to using traditional oxidants in the BV oxidation of ketones (13). BVMOs can typically be categorized into three types, I, II, and O (3). Among them, the NADPH-dependent and flavin adenine dinucleotide (FAD)-dependent type I enzymes are the most frequently reported, which have two conserved sequence motifs (FxGxxxHxxxWP/D and G/AGxWxxxxF/YPG/MxxxD) (4, 5).

Due to their excellent chemo-, regio-, and enantioselectivity, as well as great potential for the synthesis of value-added chemicals, BVMOs are considered alternative green oxidative biocatalysts. The cyclohexanone monooxygenase from Acinetobacter calcoaceticus NCIMB 9871 (CHMOAcineto) was the first reported BVMO, and this enzyme is able to selectively oxidize a wide range of cyclic ketones to the corresponding lactones (6). Numerous naturally occurring BVMOs with various substrate specificities have since been identified, some of which can catalyze the conversion of aromatic ketones to aromatic esters, such as the thermostable phenylacetone monooxygenase (PAMO) from Thermobifida fusca and 4-hydroxyacetophenone monooxygenases from Pseudomonas fluorescens ACB and Pseudomonas putida JD1 (79). Some enzymes can oxidize aliphatic ketones to aliphatic esters, such as the BVMO from Pseudomonas fluorescens DSM 50106 (BVMOPf), the BVMO from Pseudomonas putida KT2440 (BVMOPp), and the BVMO from Aspergillus flavus NRRL3357 (BVMOAFL838) (1012). Also, several stable BVMOs have been identified (13, 14), and protein engineering has been employed to improve the properties of BVMOs (15), including expanding the substrate scope (1619), enhancing regioselectivity (2022) or enantioselectivity (2325), stimulating sulfoxidation activity (2628), changing the cofactor dependency (2931), and increasing stability (3234).

Recently, using the developed BVMOs as an attractive alternative to traditional chemical oxidants and/or in combination with other enzymes, several useful chemicals have been successfully synthesized from readily available starting materials. Park and coworkers constructed an elegant multienzymatic cascade reaction by combining BVMOPp or BVMOPf with other enzymes to cleave renewable fatty acids or plant oils into α,ω-dicarboxylic acids and ω-hydroxycarboxylic acids of varied carbon chain lengths. In this case, varying the position of the double bond in unsaturated fatty acids and the regioselectivity of hydratases and BVMOs resulted in the formation of diverse monomeric chemicals (35, 36). Fraaije and coworkers first identified several BVMOs that catalyze the partial monooxygenation of 2-butanone into the abnormal product methyl propanoate, an important precursor for the synthesis of polymethyl methacrylates (37). The research groups of Kroutil and Bornscheuer also developed two multienzymatic cascades for the synthesis of bulk chemicals 6-aminohexanoic acid and poly-ε-caprolactone, respectively, starting from readily available cyclohexanol (38, 39). In the first two steps, cyclohexanol (200 mM) is oxidized by an alcohol dehydrogenase to cyclohexanone, followed by BV oxidation by a variant of CHMOAcineto to ε-caprolactone with almost 100% yield. Fink and Mihovilovic synthesized 3-hydroxypropionic acid (3-HP) derivatives from inexpensive levulinic acid derivatives with a space-time yield (STY) of 2.4 g liter−1 day−1 using a CPMO variant from Comamonas sp. strain NCIMB 9872 (40). Oppermann’s group employed BVMOAFL838 to convert 50 mM octyl aldehyde for the synthesis of alkyl formates that are widely used in the flavor and fragrance industries (41). Mihovilovic’s group utilized a biocatalytic multistep cascade involving CHMOAcineto to produce chiral carvolactone starting from limonene (42).

Although numerous BVMOs have been reported and used in organic synthesis, new enzymes with potential for industrial applications are in high demand. In the present study, a highly soluble BVMO from Rhodococcus pyridinivorans DSM 44555 (BVMORp; NCBI RefSeq accession number WP_060655096) was cloned and overexpressed in Escherichia coli BL21(DE3). BVMORp displayed excellent regioselectivity for linear aliphatic ketones and their derivatives. Using crude cell extract containing BVMORp, 200 mM methyl levulinate was completely converted into 3-hydroxypropionate with an STY of 5.4 g liter−1 h−1, demonstrating great potential for the industrial production of 3-HP from renewable biobased levulinic acid.

RESULTS

Screening of BVMOs.

Seven putative BVMO genes from R. pyridinivorans DSM 44555 were cloned and heterologously expressed in E. coli BL21(DE3). In order to identify the best candidate enzyme, these enzymes were tested for their ability to oxidize several aliphatic ketones. The results showed that BVMORp (NCBI RefSeq accession number WP_060655096) was the most effective catalyst for the conversion of all substrates (Table 1) and was therefore chosen for further study.

TABLE 1.

Conversion of ketones by various BVMOsa

Enzyme NCBI RefSeq or GenBank accession no. Conversion (%) for:
2-Octanone (substrate 3) 3-Octanone (substrate 4) Methyl levulinate (substrate 12)
BVMORp WP_060655096 >99 98 >99
BVMO1 WP_064061488 95 70 66
BVMO2 WP_006550557 99 41 32
BVMO3 WP_064060773 0.3 0.3 0.4
BVMO4 WP_064061141 2.0 3.0 0.6
BVMO5 WP_006554318 6.0 5.0 <0.1
BVMO6 SEB82458 0.6 0.3 0.3
a

Reaction conditions were 100 mM substrate, 150 mM glucose, 0.5 mg glucose dehydrogenase, 0.5 mM NADP+, 0.1 mM FAD, cell extract from 10 mg cells (wet weight), and 400 mM potassium phosphate buffer (KPB [pH 7.5]), with 0.5 ml total volume at 30°C and 1,000 rpm for 4 h.

Characterization of BVMORp.

SDS-PAGE showed that BVMORp was present almost exclusively in the soluble fraction of the induced cell lysate, yielding a prominent band at 60 kDa (Fig. 1). BVMORp was active between pH 6.0 and 10.3, retaining over 40% of maximal activity at pH 8.0 to 9.0, with an optimal pH of 8.0 in potassium phosphate buffer (Fig. 2a). The optimum pH for stability was 7.0 to 7.5, and ∼85% of the initial activity was retained after incubation for 12 h at 30°C, compared with only 55% at pH 8.0 (Fig. 2b). Although the optimal temperature for activity in the 10-min assay was between 35°C and 40°C (Fig. 2c), BVMORp was relatively stable at 30°C, with a half-life of 23 h, compared with 10 h at 35°C and 6.1 min at 40°C (Fig. 2d). Therefore, pH 7.5 and 30°C were chosen for use in the subsequent BV oxidation catalyzed by BVMORp.

FIG 1.

FIG 1

SDS-PAGE analysis of purified recombinant BVMORp. Lane 1, protein molecular weight markers; lane 2, supernatant of induced cells; lane 3, flowthrough; lanes 4 to 8, elution fractions.

FIG 2.

FIG 2

Effects of pH and temperature on the activity and stability of BVMORp. (a) Effect of pH on activity determined by assaying the activity at 30°C in 100 mM NaAc-HAc buffer (pH 5.5 to 6.0 [●]), potassium phosphate buffer (pH 6.0 to 8.0 [▲]), and Gly-NaOH (pH 8.0 to 11.0 [■]). (b) Effect of pH on stability determined by assaying the residual activity of BVMORp after 12 h at 30°C and different pH values in 100 mM potassium phosphate buffer (▲), and Gly-NaOH (■). (c) Effect of temperature on activity. (d) Effect of temperature (30°C, ▲; 35°C, ■; 40°C, ●) on stability. Residual activity is expressed as a percentage of activity measured at 0 h.

Specific activities of BVMORp toward various ketones.

A variety of aliphatic ketones and their derivatives were used for exploration of the substrate scope of BVMORp (Fig. 3). Activity toward methyl aliphatic ketones (substrates 1, 3, and 6) was the highest among linear aliphatic ketones of equal chain length, and there was a slight decrease with increasing carbon chain length (Fig. 4).

FIG 3.

FIG 3

Selected ketone substrates used for BVMORp oxidation.

FIG 4.

FIG 4

Specific activity of BVMORp toward different ketone substrates.

According to previous literature, the position of the carbonyl group does not significantly affect the reactivity of BVMOs. For instance, BVMOPp, BVMOPf, and MekA display similar efficiencies toward 2-, 3-, and 4-decanones (10, 11, 43). For octanones, BVMORp also displayed similar specific activities toward 2-, 3-, and 4-octanone (substrates 3 to 5, respectively). Interestingly, for decanones, this enzyme exhibited significantly lower activity toward 3-decanone (substrate 7) than toward other 2-, 4-, and 5-decanones (substrates 6, 8, and 9, respectively) (Fig. 4).

In addition, enzyme activity was also observed toward methyl ketones (substrates 10 to 12) with different substituents at the end of the open chain (Fig. 4). Compared with nonpolar ester groups, polar substituents with carboxyl and hydroxyl groups were less efficient substrates for BVMORp, possibly due to incompatibility with the predominately hydrophobic cavity of the enzyme.

Kinetic parameters for BVMORp with various ketone substrates.

In order to explore how carbonyl group position and carbon chain length influence BVMORp performance, the kinetic parameters Km and kcat were determined (Table 2). The Km for BVMORp was <3.0 μM with aliphatic methyl ketones 1, 3, and 6, comparable to the value observed for MekA (6.0 μM) toward 2-butanone (43).

TABLE 2.

Kinetic parameters for BVMORp with various ketone substrates

Substrate Km (mM) kcat (s–1) kcat/Km (s–1 mM–1)
1 1.5 × 10−3 2.06 ± 0.05 1,372
2 1.4 ± 0.1 1.70 ± 0.04 1.3
3 1.8 × 10−3 1.99 ± 0.02 1,120
4 0.15 ± 0.01 1.72 ± 0.02 11.2
5 0.23 ± 0.01 1.85 ± 0.03 7.9
6 2.3 × 10−3 1.80 ± 0.04 776
7 0.48 ± 0.06 0.045 ± 0.001 0.093
8 0.17 ± 0.02 0.52 ± 0.02 3.1
9 0.16 ± 0.01 0.31 ± 0.01 1.9
12 0.35 ± 0.02 1.52 ± 0.02 4.3

For hexanones and octanones with carbonyl groups at different positions, BVMORp exhibited a similar kcat but dramatically decreased affinity for 3-ketone substrates 2 and 4 compared with 2-ketone substrates 1 and 3 (Table 2). A similar phenomenon was also observed previously for BVMOAFL838 (44). Because the enzyme was saturated by substrates at 2.0 mM, it displayed similar specific activities toward 2-, 3-, and 4-octanones (Fig. 4). For decanones, the position of the carbonyl group also had an influence on both Km and kcat values. Among all tested decanones (substrates 6 to 9), the Km value for 3-decanone (substrate 7) was the highest, but the kcat value was the smallest, which explains the worst reactivity toward 3-decanone.

Regioselectivity of BVMORp with various ketone substrates.

According to typical BV reactions, the so-called “normal” product is produced by introducing an oxygen atom next to the carbonyl group on the most substituted side (products 1a to 9a) (Fig. 3). However, “abnormal” products can be identified during the conversion of several chiral compounds and long aliphatic ketones. Biotransformation products of linear aliphatic ketones obtained using purified BVMORp were analyzed by gas chromatography (GC) or GC-mass spectrometry (GC-MS). As shown in Table 3, all bioconversion results were consistent with the specific activities. All methyl aliphatic ketones (substrates 1, 3, 6, and 12) were completely converted to normal acetates, as expected. For 3-hexanone (substrate 2), 4-ketone (substrates 5 and 8), and 5-decanone (substrate 9), normal products generally accounted for >90% of the products. Furthermore, even for 3-octanone (substrate 4) and 3-decanone (substrate 7), BVMORp displayed moderate regioselectivity (Table 3). These results clearly indicate that BVMORp identified in this work is a highly regioselective enzyme.

TABLE 3.

Regioselective oxidation of aliphatic ketones by purified BVMORp

Substrate Enzyme loading (mg ml–1) Conversion (%) Ratio of products (a:b)
1 0.04 100 100:0
2 0.20 84 97:3
3 0.04 100 100:0
4 0.04 100 83:17
5 0.04 96 95:5
6 0.04 100 100:0
7 0.20 24 73:26
8 0.20 97 99:1
9 0.20 75 92:8
12 0.20 100 100:0

Preparative biosynthesis of methyl 3-acetoxypropionate by BVMORp.

The BV oxidation of levulinate using traditional peroxyacid as a catalyst is not efficient because of the low conversion and undesirable by-product formation. On the contrary, the results in Tables 1 and 3 show that substrate 12 is a moderately good substrate for BVMORp, with a relatively high activity and excellent regioselectivity. A 200 mM concentration of substrate 12, which is a relatively high substrate concentration for BVMOs, was completely converted within 4 h, affording product 12a at 74% yield (2.15 g) as a colorless liquid (Fig. 5). 1H nuclear magnetic resonance (NMR) (400 MHz, CDCl3): δ = 4.32 to 4.22 (m, 2H), 3.65 (dd, J = 2.6, 0.9 Hz, 3H), 2.64 to 2.54 (m, 2H), 1.98 (dd, J = 2.7, 1.0 Hz, 3H). The results of 1H NMR were consistent with those from a previous report (40). The corresponding STY of 5.4 g liter−1 h−1 was based on the isolated product.

FIG 5.

FIG 5

Preparative oxidation of methyl levulinate to 3-hydroxypropionate by BVMORp.

DISCUSSION

In the present study, we cloned and expressed seven putative BVMOs from R. pyridinivorans DSM 44555. These enzyme genes could also be found in the genome of a methyl-ethyl-ketone-degrading bacterium Rhodococcus pyridinivorans SB3094 (45). Among them, BVMORp, BVMO1, and BVMO2 exhibited good conversion toward linear aliphatic ketones, and the BVMORp enzyme provided the highest conversion toward all the tested substrates (Table 1). On the basis of kinetic parameters and regioselectivity shown in Tables 2 and 3, we found that 3-ketones are a class of special substrates for BVMORp. BVMORp showed the worst affinity for 3-hexanone (substrate 2) and the lowest kcat for 3-decanone (substrate 7). The regioselectivities of BVMORp toward 4-octanone (substrate 5) and 4-decanone (substrate 8) were even higher than those toward 3-octanone (substrate 4) and 3-decanone (substrate 7). More investigations, such as molecular dynamics simulation, are required to explore the potential reasons. In addition, the kcat values of all substrates listed in Table 2 are within the range of 0.31 to 2.06 s−1 (except for 3-decanone), probably due to the limited release rate of NADP+, as originally observed for CHMOAcineto (46) and since reported for 4-hydroxyacetophenone monooxygenase (HAPMO) and PAMO (7, 8).

Compared to most reported BVMOs, BVMORp has the advantage of high-level expression of soluble protein. A considerable amount of BVMOs showed low functional expression in E. coli (10, 11, 43, 47). Many efforts have been devoted to the improvement of protein solubility, such as optimization of the gene expression systems and codon sequences and introduction of the molecular chaperones and fusion tags (10, 11, 4749). However, seldom have strategies been successfully used for increasing the expression level to a large extent. Luckily, high-level functional expression of BVMORp was achieved without extra optimization. However, the imbalance of high-level functional expression of BVMORp and insufficient intracellular FAD in recombinant E. coli resulted in the existence of inactive enzyme molecules, because the activity of cell extract of BVMORp increased significantly with the addition of 10 μM FAD to the standard assay mixture. Already reported methods in which genes of key enzymes involved in FAD synthesis were overexpressed can be adopted to increase intracellular FAD concentration. Also, BVMORp was relatively stable and has a half-life of 23 h at 30°C. With the exception of several BVMOs from thermophiles, BVMORp was found to be more stable than most reported BVMOs. Therefore, BVMORp is a promising candidate for further engineering efforts to develop a practical BVMO.

In previous studies, the most studied biological approach for 3-HP production is fermentation employing engineered Klebsiella pneumoniae or E. coli (5053). For example, 3-HP has been produced in the highest titer of 83.8 g/liter and an STY of 28 g liter−1 day−1 in K. pneumoniae (51). Fink and Mihovilovic (40) proposed an alternative approach for the production of 3-hydroxypropionates, with an STY of 2.4 g liter−1 day−1, starting from inexpensive renewable levulinic acid derivatives by Baeyer-Villiger oxidation. More efficiently, methyl 3-acetoxypropionate herein was synthesized by BVMORp-catalyzed oxidation of 200 mM methyl levulinate with an STY of 5.4 g liter−1 h−1. This represents the highest substrate loading reported to date for the BVMORp-catalyzed oxidation of methyl levulinate. Moreover, this substrate, methyl levulinate, is a derivative of renewable levulinic acid from biomass, and the product can be hydrolyzed to 3-HP, a well-known platform compound. Therefore, BVMORp-catalyzed oxidation holds great promise for industrial-scale production of 3-hydroxypropionate from readily available biobased levulinic acid to replace the traditional fermentation.

MATERIALS AND METHODS

Chemicals and materials.

All chemicals and reagents were purchased from authentic suppliers, were of reagent grade or higher, and were used without further purification, unless otherwise specified. Tryptone and yeast extract were obtained from Oxoid (Hampshire, UK). Primers were synthesized by Shanghai Generay Biotech Co., and DNA polymerases, restriction enzymes, and T4 DNA ligase were from TaKaRa (Dalian, China). The 5-ml HisTrap FF crude column was supplied by GE Healthcare. 1H NMR analysis was conducted on a Bruker Avance 400 MHz spectrometer, and chemical shifts (δ) are reported in parts per million (ppm), while coupling constants (J) are reported in Hz.

Gene cloning, expression, and protein purification.

The gene encoding BVMORp was amplified by PCR using genomic DNA isolated from R. pyridinivorans as the template and primers 5′-CCCAAGCTTGCATGACATCAACCCATCTCCCTGAGCACGC-3′ (forward) and 5′-CCGCTCGAGTCACGATACCCGCAGTCCGTCGGTGTATC-3′ (reverse). The PCR product was purified using a PCR product purification kit, followed by digestion with the HindIII and XhoI restriction enzymes. The restriction fragment was inserted into the pET-28a(+) expression vector digested by the same enzymes, resulting in the pBVMORp plasmid. After verification by DNA sequencing, the pBVMORp plasmid was transformed into E. coli BL21(DE3) cells prior to protein expression experiments.

Recombinant E. coli BL21(DE3) cells were cultured at 37°C with shaking at 180 rpm in Luria-Bertani (LB) medium containing 50 mg/liter kanamycin until the absorbance at 600 nm (OD600) reached 0.6 to 0.8. Protein expression was induced by the addition of 0.1 mM isopropyl-β-d-thiogalactopyranoside (IPTG), and culturing was continued at 16°C and 180 rpm for 24 h. Cells were harvested by centrifugation (4°C, 7,000 × g, 10 min), resuspended in binding buffer (20 mM phosphate, 500 mM NaCl, 10 mM imidazole [pH 7.4]), and disrupted by ultrasonication. The cell lysate was centrifuged for 30 min at 4°C at 18,000 × g. The supernatant was loaded onto a 5-ml HisTrap FF column preequilibrated with binding buffer and washed with buffer containing 20 mM phosphate and 500 mM NaCl (pH 7.4), accompanied by stepwise addition of imidazole at 10 mM, 60 mM, 160 mM, 310 mM, and 500 mM. After analysis by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), elution fractions containing high-purity BVMORp were collected and buffer exchanged into 20 mM phosphate (pH 7.4) using a 30-kDa-cutoff centrifugal concentrator (Millipore, USA). The purified enzyme was stored with 20% (vol/vol) glycerol and 10 mM FAD at −80°C.

Enzyme assay.

Enzyme activity was assayed spectrophotometrically at 30°C by monitoring the decrease in absorbance of NADPH at 340 nm (ε340 = 6.22 cm−1 mM−1). The standard assay mixture (1 ml) contained 100 mM potassium phosphate buffer (pH 8.0), 2.0 mM substrate, 0.1 mM NADPH, and an appropriate amount of enzyme. One unit of BVMORp was defined as the amount of enzyme catalyzing the oxidation of 1 μmol NADPH per min. The uncoupling NADPH was taken into consideration.

The kinetic parameters for BVMORp with different ketone substrates were determined by standard activity assays at 30°C at various substrate concentrations and a fixed NADPH concentration of 0.2 mM. Apparent Km and kcat values were calculated by nonlinear regression fitting of data to the Michaelis-Menten equation using the Origin software. Protein concentration was determined using a modified Bradford protein assay kit (Sangon Biotech, China).

Effect of pH and temperature on activity and stability.

Cyclohexanone was used as a model substrate for investigating the effects of pH and temperature on activity and stability. The effect of pH on the activity of BVMORp was determined by enzyme activity assays at 30°C in buffers of different pH (5.5 to 11.0). The effect of pH on the stability of BVMORp was examined by measuring the residual activity after incubating purified BVMORp (0.5 mg/ml) at 30°C for 12 h in buffers of different pH (6.0 to 10.0). The buffers were as follows: 100 mM NaAc-HAc (pH 5.5 to 6.0), 100 mM potassium phosphate (pH 6.0 to 8.0), and 100 mM Gly-NaOH (pH 8.0 to 11.0).

The effect of temperature on the activity of BVMORp was investigated by assaying enzyme activity in potassium phosphate buffer (100 mM [pH 7.5]) at different temperatures (20 to 50°C). Thermostability was examined by incubating purified BVMORp (0.75 mg/ml) in potassium phosphate buffer (100 mM [pH 7.5]) at 30, 35, and 40°C and measuring the residual activity for the designated time periods.

Biotransformation of aliphatic ketones by BVMORp.

Reaction mixtures (200 μl) for biotransformations were composed of potassium phosphate buffer (100 mM [pH 8.0]), 2 mM substrate, ethanol (5% [vol/vol]), 3 mM NADPH, and various quantities of purified BVMORp. Reaction mixtures were incubated in sealed bottles (2 ml) at 30°C with shaking at 1,000 rpm for 3 h and then extracted with an equivalent volume of ethyl acetate. Extracted samples were analyzed by gas chromatography (GC) or GC-mass spectrometry (MS) to measure the conversion of substrate and the ratio of the two ester products.

Analytical methods.

The conversion was analyzed by a Shimadzu GC-2014 instrument equipped with an Rxi-5Sil MS column (25 m by 0.25 mm by 0.25 μm; Restek). The two distinct products obtained from the oxidation of 2-hexanone, 3-hexanone, 2-octanone, 3-octanone, 2-decanone, and 3-decanone substrates were analyzed by GC (see Table S1 in the supplemental material).

The major and minor products formed from 4-octanone (substrate 5), 4-decanone (substrate 8), and 5-decanone (substrate 9) were analyzed using a Shimadzu GCMS-QP2010 instrument coupled to an InertCap 5MS/NP column (30 m by 0.25 mm by 0.25 μm; GL Science) using helium as the carrier gas. Mass spectra were collected using electrospray ionization. As the products of 5-decanone (substrate 9), pentyl pentanoate (product 9a) generates the characteristic ion at m/z 103, while butyl hexanoate (product 9b) produces the characteristic ion at m/z 99, and the proportion of product 9a in the product mixture was found to be linearly correlated with the ratio of the peak area of ion at m/z 103 to the sum of the peak areas of ions at m/z 103 and m/z 99 (Fig. S2). The regioselectivity of the enzyme was calculated using a calibration curve. Similarly for the products of 4-decanone (substrate 8), hexyl butyrate (product 8a) forms characteristic ions at m/z 71 and 89, whereas propyl heptanoate (product 8b) produces characteristic ions at m/z 113 and m/z 131. Meanwhile, as the products of 4-octanone (substrate 5), butyl butyrate (product 5a) generates characteristic ions at m/z 71 and 89, whereas propyl pentanoate (product 5b) has characteristic ions at m/z 85 and m/z 103.

Preparative synthesis of methyl 3-acetoxypropionate by BVMORp.

Methyl levulinate (2.6 g, 20 mmol) was added to 100 ml potassium phosphate buffer (400 mM [pH 7.5]) containing cell lysate from E. coli expressing BVMORp (2.0 g cells [wet weight]), 100 mg lyophilized crude glucose dehydrogenase (GDH; 2,500 U), d-glucose (5.4 g, 30 mmol), 0.5 mM NADP+, and 0.1 mM FAD. The reaction was performed in a sealed 1-liter reactor using pure oxygen instead of air at 30°C under magnetic stirring. After the conversion reached 99%, the reaction mixture was saturated with NaCl, extracted three times with ethyl acetate, and centrifuged (18,000 × g for 5 min) for phase separation. The combined ethyl acetate was washed twice with saturated NaCl solution, dried over anhydrous Na2SO4, and evaporated under reduced pressure. The crude product was purified by column chromatography on silica with a mixture of petroleum ether and ethyl acetate as the eluent (10:1 [vol/vol]).

Supplementary Material

Supplemental file 1
AEM.00239-19-s0001.pdf (461.9KB, pdf)

ACKNOWLEDGMENTS

This work was financially supported by the National Natural Science Foundation of China (no. 21536004, 21472045, and 21878085), the Shanghai Commission of Science and Technology (no. 15JC1400403), and the Fundamental Research Funds for the Central Universities (no. 22221818014).

Footnotes

Supplemental material for this article may be found at https://doi.org/10.1128/AEM.00239-19.

REFERENCES

  • 1.Roberts SM, Wan P. 1998. Enzyme-catalysed Baeyer-Villiger oxidations. J Mol Catal B Enzym 4:111–136. doi: 10.1016/S1381-1177(97)00027-1. [DOI] [Google Scholar]
  • 2.Torres Pazmiño DE, Dudek HM, Fraaije MW. 2010. Baeyer-Villiger monooxygenases: recent advances and future challenges. Curr Opin Chem Biol 14:138–144. doi: 10.1016/j.cbpa.2009.11.017. [DOI] [PubMed] [Google Scholar]
  • 3.Leisch H, Morley K, Lau P. 2011. Baeyer-Villiger monooxygenases: more than just green chemistry. Chem Rev 111:4165–4222. doi: 10.1021/cr1003437. [DOI] [PubMed] [Google Scholar]
  • 4.Fraaije MW, Kamerbeek NM, van Berkel WJH, Janssen DB. 2002. Identification of a Baeyer-Villiger monooxygenase sequence motif. FEBS Lett 518:43–47. doi: 10.1016/S0014-5793(02)02623-6. [DOI] [PubMed] [Google Scholar]
  • 5.Riebel A, Dudek HM, de Gonzalo G, Stepniak P, Rychlewski L, Fraaije MW. 2012. Expanding the set of rhodococcal Baeyer-Villiger monooxygenases by high-throughput cloning, expression and substrate screening. Appl Microbiol Biotechnol 95:1479–1489. doi: 10.1007/s00253-011-3823-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Donoghue NA, Norris DB, Trudgill PW. 1976. The purification and properties of cyclohexanone oxygenase from Nocardia globerula CL1 and Acinetobacter NCIB 9871. Eur J Biochem 63:175–192. doi: 10.1111/j.1432-1033.1976.tb10220.x. [DOI] [PubMed] [Google Scholar]
  • 7.Kamerbeek NM, Moonen MJH, van der Ven JGM, van Berkel WJH, Fraaije MW, Janssen DB. 2001. 4-Hydroxyacetophenone monooxygenase from Pseudomonas fluorescens ACB. A novel flavoprotein catalyzing Baeyer-Villiger oxidation of aromatic compounds. Eur J Biochem 268:2547–2557. doi: 10.1046/j.1432-1327.2001.02137.x. [DOI] [PubMed] [Google Scholar]
  • 8.Fraaije MW, Wu J, Heuts D, van Hellemond EW, Lutje Spelberg JH, Janssen DB. 2005. Discovery of a thermostable Baeyer-Villiger monooxygenase by genome mining. Appl Microbiol Biotechnol 66:393–400. doi: 10.1007/s00253-004-1749-5. [DOI] [PubMed] [Google Scholar]
  • 9.Rehdorf J, Zimmer CL, Bornscheuer UT. 2009. Cloning, expression, characterization, and biocatalytic investigation of the 4-hydroxyacetophenone monooxygenase from Pseudomonas putida JD1. Appl Environ Microbiol 75:3106–3114. doi: 10.1128/AEM.02707-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Kirschner A, Altenbuchner J, Bornscheuer UT. 2006. Cloning, expression, and characterization of a Baeyer–Villiger monooxygenase from Pseudomonas fluorescens DSM 50106 in E. coli. Appl Microbiol Biotechnol 73:1065–1072. doi: 10.1007/s00253-006-0556-6. [DOI] [PubMed] [Google Scholar]
  • 11.Rehdorf J, Kirschner A, Bornscheuer UT. 2007. Cloning, expression and characterization of a Baeyer-Villiger monooxygenase from Pseudomonas putida KT2440. Biotechnol Lett 29:1393–1398. doi: 10.1007/s10529-007-9401-y. [DOI] [PubMed] [Google Scholar]
  • 12.Ferroni FM, Smit MS, Opperman DJ. 2014. Functional divergence between closely related Baeyer-Villiger monooxygenases from Aspergillus flavus. J Mol Catal B Enzym 107:47–54. doi: 10.1016/j.molcatb.2014.05.015. [DOI] [Google Scholar]
  • 13.Romero E, Gómez Castellanos JR, Mattevi A, Fraaije MW. 2016. Characterization and crystal structure of a robust cyclohexanone monooxygenase. Angew Chem Int Ed Engl 55:15852–15855. doi: 10.1002/anie.201608951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Fürst M, Savino S, Dudek HM, Gómez Castellanos JR, Gutiérrez de Souza C, Rovida S, Fraaije MW, Mattevi A. 2017. Polycyclic ketone monooxygenase from the thermophilic fungus Thermothelomyces thermophila: a structurally distinct biocatalyst for bulky substrates. J Am Chem Soc 139:627–630. doi: 10.1021/jacs.6b12246. [DOI] [PubMed] [Google Scholar]
  • 15.Balke K, Beier A, Bornscheuer UT. 2018. Hot spots for the protein engineering of Baeyer-Villiger monooxygenases. Biotechnol Adv 36:247–263. doi: 10.1016/j.biotechadv.2017.11.007. [DOI] [PubMed] [Google Scholar]
  • 16.Bocola M, Schulz F, Leca F, Vogel A, Fraaije MW, Reetz MT. 2005. Converting phenylacetone monooxygenase into phenylcyclohexanone monooxygenase by rational design: towards practical Baeyer-Villiger monooxygenases. Adv Synth Catal 347:979–986. doi: 10.1002/adsc.200505069. [DOI] [Google Scholar]
  • 17.Reetz MT, Wu S. 2008. Greatly reduced amino acid alphabets in directed evolution: making the right choice for saturation mutagenesis at homologous enzyme positions. Chem Commun (Camb) 43:5499–5501. doi: 10.1039/b813388c. [DOI] [PubMed] [Google Scholar]
  • 18.Reetz MT, Wu S. 2009. Laboratory evolution of robust and enantioselective Baeyer-Villiger monooxygenases for asymmetric catalysis. J Am Chem Soc 131:15424–15432. doi: 10.1021/ja906212k. [DOI] [PubMed] [Google Scholar]
  • 19.Bisagni S, Abolhalaj M, de Brevern AG, Rebehmed J, Hatti-Kaul R, Mamo G. 2017. Enhancing the activity of a Dietzia sp. D5 Baeyer-Villiger monooxygenase towards cyclohexanone by saturation mutagenesis. ChemistrySelect 2:7169–7177. doi: 10.1002/slct.201701212. [DOI] [Google Scholar]
  • 20.Chen K, Wu S, Zhu L, Zhang C, Xiang W, Deng Z, Ikeda H, Cane DE, Zhu D. 2016. Substitution of a single amino acid reverses the regiospecificity of the Baeyer-Villiger monooxygenase PntE in the biosynthesis of the antibiotic pentalenolactone. Biochemistry 55:6696–6704. doi: 10.1021/acs.biochem.6b01040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.van Beek HL, Romero E, Fraaije MW. 2017. Engineering cyclohexanone monooxygenase for the production of methyl propanoate. ACS Chem Biol 12:291–299. doi: 10.1021/acschembio.6b00965. [DOI] [PubMed] [Google Scholar]
  • 22.Balke K, Bäumgen M, Bornscheuer UT. 2017. Controlling the regioselectivity of Baeyer-Villiger monooxygenases by mutation of active-site residues. ChemBioChem 18:1627–1638. doi: 10.1002/cbic.201700223. [DOI] [PubMed] [Google Scholar]
  • 23.Reetz MT, Brunner B, Schneider T, Schulz F, Clouthier CM, Kayser MM. 2004. Directed evolution as a method to create enantioselective cyclohexanone monooxygenases for catalysis in Baeyer-Villiger reactions. Angew Chem Int Ed Engl 43:4075–4078. doi: 10.1002/anie.200460272. [DOI] [PubMed] [Google Scholar]
  • 24.Reetz MT, Daligault F, Brunner B, Hinrichs H, Deege A. 2004. Directed evolution of cyclohexanone monooxygenases: enantioselective biocatalysts for the oxidation of prochiral thioethers. Angew Chem Int Ed Engl 43:4078–4081. doi: 10.1002/anie.200460311. [DOI] [PubMed] [Google Scholar]
  • 25.Kirschner A, Bornscheuer UT. 2008. Directed evolution of a Baeyer-Villiger monooxygenase to enhance enantioselectivity. Appl Microbiol Biotechnol 81:465–472. doi: 10.1007/s00253-008-1646-4. [DOI] [PubMed] [Google Scholar]
  • 26.Bong YK, Clay MD, Collier SJ, Mijts B, Vogel M, Zhang X, Zhu J, Nazor J, Smith D, Song S. November 2014. Synthesis of prazole compounds. US patent US8895271B2.
  • 27.Catucci G, Zgrablic I, Lanciani F, Valetti F, Minerdi D, Ballou DP, Gilardi G, Sadeghi SJ. 2016. Characterization of a new Baeyer-Villiger monooxygenase and conversion to a solely N-or S-oxidizing enzyme by a single R292 mutation. Biochim Biophys Acta Proteins Proteom 1864:1177–1187. doi: 10.1016/j.bbapap.2016.06.010. [DOI] [PubMed] [Google Scholar]
  • 28.Zhang Y, Liu F, Xu N, Wu Y-Q, Zheng Y-C, Zhao Q, Lin G, Yu H-L, Xu J-H. 2018. Discovery of two native Baeyer-Villiger monooxygenases for asymmetric synthesis of bulky chiral sulfoxides. Appl Environ Microbiol 84:e00638-18. doi: 10.1128/AEM.00638-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Kamerbeek NM, Fraaije MW, Janssen DB. 2004. Identifying determinants of NADPH specificity in Baeyer-Villiger monooxygenases. Eur J Biochem 271:2107–2116. doi: 10.1111/j.1432-1033.2004.04126.x. [DOI] [PubMed] [Google Scholar]
  • 30.Dudek HM, Torres Pazmiño DE, Rodríguez C, de Gonzalo G, Gotor V, Fraaije MW. 2010. Investigating the coenzyme specificity of phenylacetone monooxygenase from Thermobifida fusca. Appl Microbiol Biotechnol 88:1135–1143. doi: 10.1007/s00253-010-2769-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Beier A, Bordewick S, Genz M, Schmidt S, van den Bergh T, Peters C, Joosten H-J, Bornscheuer UT. 2016. Switch in cofactor specificity of a Baeyer-Villiger monooxygenase. ChemBioChem 17:2312–2315. doi: 10.1002/cbic.201600484. [DOI] [PubMed] [Google Scholar]
  • 32.Opperman DJ, Reetz MT. 2010. Towards practical Baeyer-Villiger-monooxygenases: design of cyclohexanone monooxygenase mutants with enhanced oxidative stability. ChemBioChem 11:2589–2596. doi: 10.1002/cbic.201000464. [DOI] [PubMed] [Google Scholar]
  • 33.van Beek HL, Wijma HJ, Fromont L, Janssen DB, Fraaije MW. 2014. Stabilization of cyclohexanone monooxygenase by a computationally designed disulfide bond spanning only one residue. FEBS Open Biol 4:168–174. doi: 10.1016/j.fob.2014.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Schmidt S, Genz M, Balke K, Bornscheuer UT. 2015. The effect of disulfide bond introduction and related Cys/Ser mutations on the stability of a cyclohexanone monooxygenase. J Biotechnol 214:199–211. doi: 10.1016/j.jbiotec.2015.09.026. [DOI] [PubMed] [Google Scholar]
  • 35.Song J-W, Jeon E-Y, Song D-H, Jang H-Y, Bornscheuer UT, Oh D-K, Park J-B. 2013. Multistep enzymatic synthesis of long-chain α,ω-dicarboxylic and ω-hydroxycarboxylic acids from renewable fatty acids and plant oils. Angew Chem Int Ed Engl 52:2534–2537. doi: 10.1002/anie.201209187. [DOI] [PubMed] [Google Scholar]
  • 36.Jeon E-Y, Seo J-H, Kang W-R, Kim M-J, Lee J-H, Oh D-K, Park J-B. 2016. Simultaneous enzyme/whole-cell biotransformation of plant oils into C9 carboxylic acids. ACS Catal 6:7547–7553. doi: 10.1021/acscatal.6b01884. [DOI] [Google Scholar]
  • 37.van Beek HL, Winter RT, Eastham GR, Fraaije MW. 2014. Synthesis of methyl propanoate by Baeyer-Villiger monooxygenases. Chem Commun (Camb) 50:13034–13036. doi: 10.1039/c4cc06489e. [DOI] [PubMed] [Google Scholar]
  • 38.Schmidt S, Scherkus C, Muschiol J, Menyes U, Winkler T, Hummel W, Gröger H, Liese A, Herz H-G, Bornscheuer UT. 2015. An enzyme cascade synthesis of ε-caprolactone and its oligomers. Angew Chem Int Ed Engl 54:2784–2787. doi: 10.1002/anie.201410633. [DOI] [PubMed] [Google Scholar]
  • 39.Sattler JH, Fuchs M, Mutti FG, Grischek B, Engel P, Pfeffer J, Woodley JM, Kroutil W. 2014. Introducing an in situ capping strategy in systems biocatalysis to access 6-aminohexanoic acid. Angew Chem Int Ed Engl 53:14153–14157. doi: 10.1002/anie.201409227. [DOI] [PubMed] [Google Scholar]
  • 40.Fink MJ, Mihovilovic MD. 2015. Non-hazardous Baeyer-Villiger oxidation of levulinic acid derivatives: alternative renewable access to 3-hydroxypropionates. Chem Commun (Camb) 51:2874–2877. doi: 10.1039/c4cc08734h. [DOI] [PubMed] [Google Scholar]
  • 41.Ferroni FM, Tolmie C, Smit MS, Opperman DJ. 2017. Alkyl formate ester synthesis by a fungal Baeyer-Villiger monooxygenase. ChemBioChem 18:515–517. doi: 10.1002/cbic.201600684. [DOI] [PubMed] [Google Scholar]
  • 42.Oberleitner N, Ressmann AK, Bica K, Gärtner P, Fraaije MW, Bornscheuer UT, Rudroff F, Mihovilovic MD. 2017. From waste to value–direct utilization of limonene from orange peel in a biocatalytic cascade reaction towards chiral carvolactone. Green Chem 19:367–371. doi: 10.1039/C6GC01138A. [DOI] [Google Scholar]
  • 43.Völker A, Kirschner A, Bornscheuer UT, Altenbuchner J. 2008. Functional expression, purification, and characterization of the recombinant Baeyer-Villiger monooxygenase MekA from Pseudomonas veronii MEK700. Appl Microbiol Biotechnol 77:1251–1260. doi: 10.1007/s00253-007-1264-6. [DOI] [PubMed] [Google Scholar]
  • 44.Ferroni FM, Tolmie C, Smit MS, Opperman DJ. 2016. Structural and catalytic characterization of a fungal Baeyer-Villiger monooxygenase. PLoS One 11:e0160186. doi: 10.1371/journal.pone.0160186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Dueholm MS, Albertsen M, D’Imperio S, Tale VP, Lewis D, Nielsen PH, Nielsen JL. 2014. Complete genome of Rhodococcus pyridinivorans SB3094, a methyl-ethyl-ketone-degrading bacterium used for bioaugmentation. Genome Announc 2:e00525-14. doi: 10.1128/genomeA.00525-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Sheng D, Ballou DP, Massey V. 2001. Mechanistic studies of cyclohexanone monooxygenase: chemical properties of intermediates involved in catalysis. Biochemistry 40:11156–11167. doi: 10.1021/bi011153h. [DOI] [PubMed] [Google Scholar]
  • 47.van Bloois E, Dudek HM, Duetz WA, Fraaije MW. 2012. A stepwise approach for the reproducible optimization of PAMO expression in Escherichia coli for whole-cell biocatalysis. BMC Biotechnol 12:31. doi: 10.1186/1472-6750-12-31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Seo J-H, Kim H-H, Jeon E-Y, Song Y-H, Shin C-S, Park J-B. 2016. Engineering of Baeyer-Villiger monooxygenase-based Escherichia coli biocatalyst for large scale biotransformation of ricinoleic acid into (Z)-11-(heptanoyloxy)undec-9-enoic acid. Sci Rep 6:28223. doi: 10.1038/srep28223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Song J-W, Woo J-M, Jung GY, Bornscheuer UT, Park J-B. 2016. 3′-UTR engineering to improve soluble expression and fine-tuning of activity of cascade enzymes in Escherichia coli. Sci Rep 6:29406. doi: 10.1038/srep29406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Valdehuesa KNG, Liu H, Nisola GM, Chung W-J, Lee SH, Park SJ. 2013. Recent advances in the metabolic engineering of microorganisms for the production of 3-hydroxypropionic acid as C3 platform chemical. Appl Microbiol Biotechnol 97:3309–3321. doi: 10.1007/s00253-013-4802-4. [DOI] [PubMed] [Google Scholar]
  • 51.Li Y, Wang X, Ge X, Tian P. 2016. High production of 3-hydroxypropionic acid in Klebsiella pneumoniae by systematic optimization of glycerol metabolism. Sci Rep 6:26932. doi: 10.1038/srep26932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Jiang J, Huang B, Wu H, Li Z, Ye Q. 2018. Efficient 3-hydroxypropionic acid production from glycerol by metabolically engineered Klebsiella pneumoniae. Bioresour Bioprocess 5:34. doi: 10.1186/s40643-018-0218-4. [DOI] [Google Scholar]
  • 53.Liu B, Xiang S, Zhao G, Wang B, Ma Y, Liu W, Tao Y. 2019. Efficient production of 3-hydroxypropionate from fatty acids feedstock in Escherichia coli. Metab Eng 51:121–130. doi: 10.1016/j.ymben.2018.10.003. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental file 1
AEM.00239-19-s0001.pdf (461.9KB, pdf)

Articles from Applied and Environmental Microbiology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES