• No results found

Exploring the selective demethylation of aryl-methyl ethers by a Pseudomonas Rieske monooxygenase

N/A
N/A
Protected

Academic year: 2021

Share "Exploring the selective demethylation of aryl-methyl ethers by a Pseudomonas Rieske monooxygenase"

Copied!
11
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Exploring the selective demethylation of aryl-methyl ethers by a Pseudomonas Rieske

monooxygenase

Lanfranchi, Elisa; Trajković, Miloš; Barta, Katalin; de Vries, Johannes G; Janssen, Dick B

Published in:

ChemBioChem

DOI:

10.1002/cbic.201800594

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Final author's version (accepted by publisher, after peer review)

Publication date:

2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Lanfranchi, E., Trajković, M., Barta, K., de Vries, J. G., & Janssen, D. B. (2019). Exploring the selective

demethylation of aryl-methyl ethers by a Pseudomonas Rieske monooxygenase. ChemBioChem, 20(1),

118-125. [cbic.201800594]. https://doi.org/10.1002/cbic.201800594

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

www.chembiochem.org

Accepted Article

A Journal of

Title: Exploring the selective demethylation of aryl-methyl ethers by a

Pseudomonas Rieske monooxygenase

Authors: Elisa Lanfranchi, Miloš Trajković, Katalin Barta, Johannes G

de Vries, and Dick B Janssen

This manuscript has been accepted after peer review and appears as an

Accepted Article online prior to editing, proofing, and formal publication

of the final Version of Record (VoR). This work is currently citable by

using the Digital Object Identifier (DOI) given below. The VoR will be

published online in Early View as soon as possible and may be different

to this Accepted Article as a result of editing. Readers should obtain

the VoR from the journal website shown below when it is published

to ensure accuracy of information. The authors are responsible for the

content of this Accepted Article.

To be cited as: ChemBioChem 10.1002/cbic.201800594

Link to VoR: http://dx.doi.org/10.1002/cbic.201800594

(3)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

Exploring the selective demethylation of aryl-methyl ethers by a

Pseudomonas Rieske monooxygenase

Elisa Lanfranchi,*

[a]§

Milos Trajkovic,

[a]

Katalin Barta,

[b]

Johannes G de Vries,

[b,c]

and Dick B Janssen

[a]

Abstract: Biocatalytic dealkylation of aryl-methyl ethers is an attractive reaction for valorization of lignin components and for deprotection of hydroxy functionalities in synthetic chemistry. We explored the demethylation of various aryl-methyl ethers by an oxidative demethylase from Pseudomonas sp. HR199. The Rieske monooxygenase (VanA) and its partner electron transfer protein (VanB) were recombinantly coexpressed in E. coli and they constituted at least 25% of the total protein content. Enzymatic conversions showed that VanB accepts NADH and NADPH as electron donor. The VanA-VanB system demethylates a number of aromatic substrates, the catalysis occurs selectively at meta position of the aromatic ring and the presence of a carboxylic moiety is essential. The reaction is inhibited by the byproduct formaldehyde. Therefore, we tested three different cascade/tandem reactions for cofactor regeneration and formaldehyde elimination; conversion was especially improved by adding formate dehydrogenase and formaldehyde dehydrogenase. Finally, the biocatalyst was applied for the preparation of protocatechuic acid from vanillic acid giving 77% yield of desired product. The described reaction may find application in the conversion of lignin components to diverse hydroxyaromatic building blocks and generally opens prospective to new mild methods for efficient deprotection of phenols.

Introduction

Phenolic groups occur in numerous natural bioactive compounds and are important functionalities during the synthesis of specialty chemicals. However, due to their high reactivity, protection of phenolic groups is often required prior to further syntheses, for example by preparing stable and unreactive methyl ethers.[1]

While such methylation strategies are highly effective for the protection of phenols, the subsequent deprotection of the inert methyl ethers generally requires harsh reaction conditions such as the use of strong acids or bases.[2,3] Alternative approaches

that allow more facile demethylation in a mild environment are therefore desired. Besides in synthetic chemistry, novel demethylation protocols could also be exploited in the valorization of compounds derived from lignin. This complex natural polymer is rich in aryl-methyl ethers groups derived from the precursors coniferyl and sinapyl alcohol.[4] In the past decade, great progress

was made regarding the development of catalytic strategies for lignin depolymerization to produce fine chemicals.[5–7]

Independent of the method used, the majority of these processes result in aromatic compounds that comprise the guaiacyl or syringyl moiety.[8,9] O-demethylation reactions would generate

hydroxylated aromatics that may serve as building blocks for the production of value-added chemicals using diverse synthetic strategies. For example, the demethylation of guaiacyl derivatives leads to catechols that are present in numerous pharmaceuticals, pesticides and flavors. Alternatively, catechols are converted by oxidation to muconic acid which canbe hydrogenated to adipic acid derivatives to be used in polymer synthesis.[10,11]

Green and environmentally friendly O-demethylating solutions can in principle be addressed by biocatalysis. However, whereas demethylases are widely present in nature, they have barely been explored for biocatalytic applications.[12] Reasons might be the

complex architecture of some of these multi-component systems as well as the requirement for expensive and/or delicate cofactors (e.g. cobalamin- and/or tetrahydrofolate-dependent enzymes). An example of such a demethylation system is cytochrome P450 CYP199A4 from Rhodopseudomonas palustris, a heme-dependent monooxygenase which catalyzes the oxidative demethylation of 4-methoxybenzoic acid to 4-hydroxybenzoic acid.[13,14] The enzyme was recombinantly expressed in E. coli

together with electron transfer proteins and whole cells were employed for the transformation of veratric acid to vanillic acid.[13]

CYP199A4 only demethylates methoxy groups at the para position of the phenyl ring. It tolerates other substituents at the

ortho and meta positions, whereas the carboxylate moiety is

critical for substrate recognition.[15] Interestingly, the single

Ser244Asp mutation in the active site of the enzyme unlocked the carboxylate dependence and extended the substrate scope to benzene derivatives.[16]

Lignin harbors most of the O-methyl groups at the 3 or 3-5 (meta) positions on the aromatic ring and accordingly demethylases with different regiospecificity have raised interest. The tetrahydrofolate (THF)- dependent O-demethylase LigM from

Sphingobium sp. SYK-6 was recently characterized and the

crystal structure was solved.[17–20] The expensive THF cofactor

was required in 10-fold molar excess over substrate in order to achieve good conversion.[18] Rosini and coworkers proposed a

tandem reaction exploiting a methionine synthase for cofactor recycling.[18] This enzyme (MetE) converts 5-methyl-THF back to

THF by the transfer of the methyl group to L-homocysteine, forming L-methionine. With the so developed biotransformation, the authors were able to obtain 5 mM protocatechuic acid (PCA, [a] Dr. E. Lanfranchi, Dr. M. Trajkovic, Prof. D. B. Janssen

Groningen Biomolecular Sciences and Biotechnology Institute (GBB)

University of Groningen

Nijenborgh 4, 9726 AG Groningen, The Netherlands E-mail: elisa.lanfranchi@ucc.ie

[b] Prof. K. Barta, Prof. J. G. de Vries

Synthetic Organic Chemistry - Stratingh Institute for Chemistry University of Groningen

Nijenborgh 4, 9726 AG Groningen, The Netherlands [c] Leibniz-Institut für Katalyse e.V.

Albert-Einstein-Strasse 29a, 18059 Rostock, Germany

§ Current address: School of Food and Nutritional Science Sciences University College Cork

College Road, Cork T12 YN60, Ireland

Supporting information for this article is given via a link at the end of the document.

Accepted

(4)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

1b) with only 0.1 mM THF added. Nevertheless, in view of the high cost of the cofactor, the availability of THF-independent

meta-demethylation systems could offer an attractive alternative.

One example is the new discovered cytochrome P450 from

Amycolatopsis sp. ATCC 39116, which catalyzes the

O-demethylation of guaiacol and its analogs at the meta position of the aromatic ring.[21]

Early genetic studies indicated that bacteria degrading lignin-derived components and dimeric model compounds express an additional type of demethylases to LigM, where demethylation of vanillic acid (1a) to PCA is catalyzed by a two-component system consisting of VanA and VanB.[22–24] Sequence similarities suggest

that VanA belongs to the Rieske non-heme iron-dependent monooxygenases and that it is a homolog of a demethylating dicamba (3,6-dichloro-2-methoxybenzoic acid) monooxygenase (DdmC) of which the structure and catalytic mechanism are known.[25–27] Rieske proteins are distinguished by the presence of

an iron-sulfur cluster [2Fe-2S] that functions as an electron transfer and electron storage organometallic complex. The active site contains a single free iron ion involved in the binding and activation of molecular oxygen, allowing substrate hydroxylation.[28,29]

One example of such a two-component system is the O-demethylase from Streptomyces (StVanA) which has been partially characterized by Nishimura and coworkers.[30] They

explored the substrate scope of StVanA after recombinant protein expression using whole-cell bioconversions with E. coli. This work showed the potential value of StVanA-VanB as a whole-cell biocatalyst, but also underlined the need to improve expression of soluble protein. Difficulties with heterologous production were also reported for other iron-sulfur proteins, which make it, together with protein instability, a major limitation for the exploitation of Rieske oxygenases (ROs).[31]

A demethylase system that is potentially attractive for applied biocatalysis is present in Pseudomonas sp. HR199 (DSM 7063), a strain able to degrade lignin-derived compounds to products of lower molecular weight.[32] It transforms eugenol to vanillin, which

is further oxidized to vanillic acid and then converted to PCA. The latter is further metabolized by dioxygenase-mediated cleavage of the aromatic ring.[33] The sequence of the monooxygenase

component of the vanillate demethylase system from

Pseudomonas HR199 (Ps(HR199)VanA) was identified in the late

nineties and the enzymatic activity was confirmed after recombinant expression in E. coli.[34] However, the properties of

the protein and selectivity of the enzyme have never been thoroughly investigated.

Here we describe the optimization of recombinant coexpression of Pseudomonas HR199 VanA and VanB in E. coli. Subsequently, the catalytic properties of the two-component system were determined using lyophilized cells and we explored its suitability for biocatalytic applications. Finally, VanA-VanB were employed in multi-enzyme cascade reactions for effective cofactor regeneration and byproduct removal.

Scheme 1. Proposed cascade for O-demethylation of aryl-methyl ethers

catalyzed by Rieske non-heme monooxygenase coupled with a cofactor regeneration system.

Results and Discussion

Amino acid sequence analysis

A multiple sequence alignment of the monooxygenase VanA from Pseudomonas HR199 (UniprotKB O05616) with StVanA and DdmC was performed (Figure S5, Supporting Information).

Ps(HR199)VanA is 36% identical to StVanA and 34% to DdmC.

The latter was the only appreciable hit obtained after a Blast search of Ps(HR199)VanA against the PDB database. Key residues and motifs were pinpointed by comparison to the reference sequence DdmC. Rieske domains are characterized by two conserved His and two Cys residues which coordinate the 2Fe-2S cluster. Cys47-His49 and Cys66-His69 are the first and second ligand pairs distinguishable by the typical CXH and CXXH motifs. Regarding the active site, the putative residues involved in non-heme Fe ion coordination are His156, His161 and Asp298. The so called bridging Asp is also conserved (Asp153).[25,35] This

observation suggests a multimeric structure of Ps(HR199)VanA and, similarly to other ROs enzymes, that the electron transport chains occurs inter-molecular from the 2Fe-2S cluster of monomer A to the active site of monomer B. Based on the results of D’Ordine and coworkers regarding DdmC, the proposed electron-transfer chain in Ps(HR199)VanA involves αHis49, αHis69, βAsp153, βHis156 and βHis161.[25] The multiple

sequence alignment suggests a different architecture of the active site and substrate coordination between DdmC and vanillate oxygenase. For example, the key residues His251, Asn230 and Trp285 for the coordination of the carboxylic moiety in DdmC are not conserved in Ps(HR199)VanA, in congruence with what has been reported by Dumitru and coworkers. [26] This is explainable

by the different substrates and dealkylation position of the two enzymes (DdmC catalyzes a demethylation at ortho position of the aromatic ring).

VanB (UniprotKB O05617) is the partner reductase enzyme, indispensable for the supply of electrons to the Rieske domain of VanA. A protein Blast search of its amino acid sequence against

10.1002/cbic.201800594

Accepted

Manuscript

ChemBioChem

(5)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

the PDB database retrieved only one hit with a high score (37%

identity and 100% query coverage): the phthalate reductase from

Burkholderia cepacia.[36] A pairwise alignment of the two amino

acid sequences was performed (Figure S6, Supporting Information). A conserved domain search (CDD database) identified a flavin and a nicotinamide cofactor binding domain, in agreement with a role in hydride transfer from NAD(P)H to the flavin cofactor during catalysis. Additionally, the sequence alignment showed a conserved ferredoxin domain recognizable by the four conserved Cys residues responsible of the coordination of the 2Fe-2S cluster (Cys266, Cys271, Cys274 and Cys304). These results suggested that, similar to other non-heme oxygenase enzymes, the partner reductase Ps(HR199)VanB reduces the flavin cofactor by oxidation of NAD(P)H and then transfers the electrons to the ferredoxin domain, which likely interacts with the Rieske domain of VanA and provides electrons necessary for the catalysis.

Coexpression of VanA and VanB in E. coli

To obtain appreciable amounts of the enzymatic demethylation system, the VanA and VanB proteins were coexpressed in E. coli BL21 Star (DE3) using a pET-derived vector named pE2T_Ps(HR199)VanA_VanB. In order to boost transcription, each coding DNA sequence was inserted downstream of an independent T7 promoter. Preliminary protein purification experiments by affinity chromatography failed due to the partial loss of the [2Fe-2S] complexes and low yields. Additionally, a detrimental effect of an N-terminally fused 6xHis-tag was observed on the activity of the purified enzymes (data not shown). Therefore, the two proteins were coexpressed in their native form, without any affinity tag. Various cultivation conditions for IPTG-induced protein production were examined, which revealed that high-level soluble overexpression of both proteins was best achieved when additional iron and sulfur sources were added, suggesting that maturation of the iron-sulfur clusters is a critical step in protein synthesis (Experimental Section). After optimization, the two enzymes were well expressed and constituted approximately 25% of the total soluble protein in cell lysates, as indicated by two prominent bands at 39 kDa (VanA) and 35 kDa (VanB) that were observed upon analysis by SDS-PAGE (Figure S1, Supporting Information).

The high-level recombinant coproduction made it possible to avoid protein isolation and prepare a stable biocatalyst by lyophilization of whole-cells. After centrifugation and washing, 14.2 g of wet cell weight (WCW) was obtained per l of culture, which corresponded to 3.0 g of dry cell weight (DCW) per l of culture after freeze-drying. The obtained powder containing the biocatalyst VanA-VanB was used for all following experiments.

Table 1. Initial screening of demethylation reaction conditions with 7.5

mg/mL of freeze-dried cells after 4 h of incubation.

Entry Reaction Conditions PCA Analytical

Yield (%)[a]

1 0.1 mM NADH, 0.1 mM FeSO4, 10% DMSO,

Regeneration system[b] 58.8 ± 2.6

2 0.1 mM NADPH, 0.1 mM FeSO4, 10%

DMSO, Regeneration system 44.4 ± 0.6

3 0.1 mM FeSO4, 10% DMSO, Regeneration

system 20.5 ± 0.2

4 0.5 mM NADH, 0.1 mM FeSO4, 10% DMSO,

Regeneration system 30.7 ± 0.2

5 0.1 mM NADH, 0.1 mM FeSO4, 10% DMSO 7.1 ± 0.8 6 0.1 mM NADH, 10% DMSO, Regeneration

system 46.7 ± 2.0

7 0.1 mM NADH, 1 mM FeSO4, 10% DMSO,

Regeneration system 60.2 ± 0.3

8 0.1 mM NADH, 0.1 mM FeSO4,

Regeneration system 53.3 ± 1.9

9 Control 1: No cells

(Conditions entry 1) ≤2

10 Control 2:[c] E. coli BL21 (DE3) Star

(Conditions entry 1) ≤2

11 Control 3:[d] PCA instead of vanillic acid

(Conditions entry 1) 80.5

12 Control 4:[d] No cells, PCA instead of vanillic

acid (Conditions entry 1) 80.6

13 Control 5:[d] E. coli BL21 (DE3) Star, PCA instead of vanillic acid (Conditions entry 1)

92 [a] Determined by HPLC. [b] Regeneration system: 10 µM PTDH and 10 mM phosphite. [c] Measured in triplicate.[d] Single experiment.

Biocatalytic reaction characterization

Since information on the biochemical properties of vanillate

O-demethylases is scarce, we examined the biocatalytic activity

of the Pseudomonas HR199 VanA-VanB system under various reaction conditions. In these experiments, 5 mM vanillic acid (1a) was chosen as the model substrate and the formation of protocatechuic acid (PCA, 1b) was studied. The conversion was carried out for 24 h and both substrate and product concentrations were determined by HPLC (Table 1 andTable S1, Supporting Information).

Initial experiments were performed by adding rehydrated freeze-dried cells containing VanA-VanB, NAD(P)H and the phosphite dehydrogenase (PTDH) cofactor regeneration system to the reaction mixture (Scheme 1, reaction conditions: Table 1, entries 1, 2 and 4).[37] The oxidoreductase VanB accepted both

NADPH and NADH as reducing electron donor, with a preference for the latter. This dual cofactor acceptance is in accordance to what was previously reported for other Pseudomonas species.[38]

Since omission of NADH and NADPH from reaction mixtures drastically reduced conversion of vanillic acid (Table 1, entry 3), we conclude that addition of the reduced nicotinamide-cofactor is essential to obtain high yields of the demethylated product PCA and that both NADH and NADPH are incorporated by rehydrated freeze-dried cells. In situ cofactor regeneration is commonly adopted for biocatalytic applications of nicotinamide-cofactor dependent oxidoreductases and avoids the need of stoichiometric addition of costly NAD(P)H (Scheme 1). As expected, O-demethylation coupled with PTDH for cofactor regeneration

Accepted

(6)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

allowed the reaction to proceed beyond stoichiometric

consumption of the reduced cofactor, while the formation of PCA was much lower when cofactor regeneration was omitted (Table 1, entry 5).

The addition of an excess of iron to the reaction mixture showed a beneficial effect on the overall conversion, which is in accordance with the involvement of the metal in catalysis through oxygen activation. Finally, VanA-VanB tolerated 10% DMSO, which is advantageous in case of poorly soluble substrates. Based on the above, we used 0.1 mM NADH, 0.1 mM FeSO4,

10% DMSO and the PTDH-based cofactor regeneration system for further characterization experiments (Table 1, entry 1).

The described conversions showed a molar imbalance between the consumption of the substrate and the production of PCA (Table S1, Supporting Information), suggesting a partial conversion of vanillic acid by unknown E. coli enzyme(s) or partial loss of the catechol due to its instability or further enzymatic degradation. To examine these possibilities, a series of controls were designed. Initially, the reaction mixture containing 5 mM vanillic acid was incubated without the biocatalyst by omission of cells or by using the E. coli strain lacking VanA-VanB (Table 1, entries 9 and 10). The latter reaction showed 37% substrate depletion but no specific product was identified (Table S1, Supporting Information). Additionally, a partial consumption of the product was observed when 5 mM PCA was added to the mixture instead of vanillic acid (Table 1, entries 11, 12 and 13; Table S2, Supporting Information), confirming the occurrence of some product loss due to a non-biocatalytic process by endogenous E. coli enzymes and compound instability.

Oxygen sensitivity

Whilst anaerobic conditions might be an optimal environment for the stability of iron-sulfur proteins, the demethylation reaction uses molecular oxygen as reactant that obviously cannot be left out. To examine the possibility that the enzymatic demethylation was restricted by detrimental effects of molecular oxygen on the activity and stability of the VanA-VanB system, we examined the biocatalytic formation of PCA in the presence of antioxidants at various concentrations (Figure 1).

In the absence of an antioxidant, a dramatic loss of enzyme activity was observed after 2 h of pre-incubation at 30°C, with only 29% of the initial activity retained and less than 1 mM PCA was detected. The biocatalyst was essentially inactive after 4 h of pre-incubation (Figure 1A). The addition of DTT substantially improved the stability for the first 2 h and more than 80% of the initial activity was retained (Figure 1A). The beneficial effect of the antioxidant was also visible when examining the effective conversion (Figure 1B). Higher product formation was achieved when glutathione (GSH) or DTT were added. The best result was recorded in the presence of 5 mM DTT, where 80% analytical yield was obtained (Figure 1B). These results suggested that oxygen has a large effect on protein stability.

Figure 1. Effect of antioxidants on the Pseudomonas HR199 VanA-VanB

dealkylation system. A: Effect on enzyme stability. Freeze-dried cells (7.5 mg/mL) were pre-incubated in 100 mM Tris-HCl, pH 7.5, with varying concentrations of DTT at 30°C and 100 rpm. At the end of the pre-incubation time, the biotransformation was initiated by addition all remaining components and product formation was determined after 2 h. PCA formation at 0.25 h pre-incubation (time required for cells rehydration) was set to 100% for each condition. B: Effect on bioconversion. Biotransformation was carried out as described under Experimental Section with addition of different reducing agents at various concentrations. After 4 h of incubation, the concentrations of product and substrate were determined by HPLC. Negative control (Neg Ctr): Table 1, entry 10. Positive control (Pos Ctrl): Table 1, entry 1. BME, b-mercaptoethanol; GSH, glutathione; DTT dithiothreitol. Error bars indicate the standard deviation obtained from two independent experiments.

Substrate scope

To investigate the substrate specificity of VanA, a panel of 19 compounds was selected (Scheme S1, Supporting Information). Each biocatalytic reaction was set up with cofactor regeneration as outlined in Scheme 1 and incubations were carried out for 7 h.

At first, the biocatalyst was tested for the conversion of vanillin, vanillyl alcohol, acetovanillone and methyl vanillate to their corresponding catechols. No activity was detected, which suggests a requirement for a carboxylic moiety for substrate recognition. The result is in accordance with previous in vivo studies on Acinetobacter, Streptomyces and P. fluorescens vanillate demethylases.[30,39,40] The presence of longer aliphatic

chains in the more bulky compounds was unfavorable as well, and homovanillic acid, ferulic acid, dihydroferulic acid, sinapyl acid and vanilpyruvic acid were not accepted and no product formation or substrate consumption was observed.

10.1002/cbic.201800594

Accepted

Manuscript

ChemBioChem

(7)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

Table 2 reports the obtained analytical yield of a series of methoxy aromatic acids and related products observed by HPLC. VanA showed high activity for 4-methyl-m-anisic acid (2a), veratric acid (3a) and m-anisic acid (4a), and ≥85% of substrate was converted to the corresponding phenol. Depletion of veratric acid is comparable to vanillic acid, but the amount of product increased by ca. 20% (Table S3, Supporting Information). This improvement can be explained by the higher stability of isovanillic acid (3b) vs the catechol functionality, resulting in a better product recovery. Additionally, a control reaction showed that 3a is not significantly subjected to breakdown by endogenous E. coli enzymes, whereas up to 32% consumption of vanillic acid was observed in the absence of the biocatalyst VanA-VanB (Table S1 and S4, Supporting Information).

In spite of the presence of two distinct methoxy functionalities, the conversion of 3a led to only one product, implying that the oxidative demethylation occurs regiospecifically at position 3 of the aromatic ring (meta). To confirm the high regio-selectivity of VanA, meta-, ortho- and para-anisic acid (4a, 5a and 6a) were assayed. No product formation was observed after incubation with 5a and 6a, while 4a showed 86.8% analytical yield. Interesting, when both positions 3 and 5 bear a methoxy substituent (7a), 3,4-dihydroxy-5-methoxy benzoic acid (8a) was obtained in 60% yield, but the reaction did not go any further and 8a was not accepted as substrate. The latter result distinguishes

Ps(HR199)VanA from both AcVanA and StVanA. The first cannot

convert 7a, while StVanA is less specific, and generates a mixture of 8a and gallate (StVanA) respectively.[30,40]

Biocatalytic production of PCA

To investigate the applicability of the biocatalytic reactions described here, a semi-preparative bioconversion was set up as depicted in Scheme 1. An amount of 10 mg of vanillic acid was added to 12 mL reaction mixture with 360 mg of freeze-dried cells (Experimental Section). After vanillic acid was completely consumed the product was isolated and purified and 7.1 mg of PCA (77% yield) was obtained as white crystals; the product identity was confirmed by NMR (Figure S13 and S14, Supporting Information).

Table 2. Substrate scope of VanA-VanB. Reaction mixtures contained 7.5

mg/mL of freeze-dried cells, 0.1 mM NADH, cofactor regeneration system, 0.1 mM FeSO4, 10 mM DTT and 5 mM substrate. Product formation was determined after 7 h of incubation.

Substrate Observed Product Analytical Product Yield (%)[a] 69 87.5 ± 0.7 88.6 ± 06 86.8 ± 2.6 n.d.[b] n.d. 60.0 ± 4.1 n.d. [c] n.d

[a] Determined by HPLC. [b] n.d.: no product was observed. [c] 15 mg/mL freeze-dried cells.

Multi enzymatic cascades

The byproduct of oxidative demethylation is formaldehyde, which is formed in stoichiometric amounts. Hibi and coworkers showed by deletion studies in E. coli that formaldehyde can be toxic to the cells, suggesting the necessity of a rapid detoxification machinery for efficient biotransformation.[38] This raises the

question if formaldehyde is toxic to the overall cell machinery or specifically affects VanA-VanB activity. To investigate its possible impact on enzymatic activity, 2 mM formaldehyde was added to the reaction mixture together with 5 mM vanillic acid. As a result, the multi enzymatic system was inhibited and conversion dropped by ~ 50% (Figure 2A).

Product inhibition or product toxicity can be overcome by continuous removal of the undesired compound along the course of the reaction. One method is to include an additional enzyme which accepts the product as substrate. Formaldehyde

Accepted

(8)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

dehydrogenase (FADH) catalyzes the oxidation of formaldehyde

to formic acid together with the reduction of NAD+ to NADH. We

hypothesized that the combination of the three, VanA, VanB and FADH, will result in a self-sufficient tandem reaction, where cofactor regeneration was concomitant to elimination of formaldehyde. The cascade could be further extended by introduction of formate dehydrogenase (FDH), driving the reaction to completion by converting formaldehyde to CO2. This system

would also favor the reaction workup and purification of the phenolic product.

These proposed enzymatic cascades in which PTDH-driven cofactor regeneration was omitted were tested for the synthesis of PCA from 5 mM vanillic acid and compared to the standard VanA-VanB system with NADH regeneration by PTDH (Figure 2). Good conversion was achieved by both of the designed cascades without the need of extensive optimization (one-step cascade with FADH 56.5±1.9% analytical yield, two-step cascade with FADH+FDH 84.4±3.3% analytical yield after 4 h reaction time). However, a 5-fold higher concentration of NADH was required compared to the 0.1 mM previously adopted. This can be explained by the different nature of cofactor recycling machinery: FADH and FDH activity are dependent on formaldehyde formation by VanA-VanB, while PTDH reduces the cofactor by phosphite that is present in excess from the beginning. Conversions stopped after 2 h, probably due to VanA-VanB instability. Whereas VanA-VanB-FADH seems comparable to VanA-VanB-PTDH, a significant improvement was observed for the 2-step aldehyde oxidation (VanA-VanB-FADH-FDH) and 4.1 mM PCA was obtained (Figure 2C).

Figure 2. Formaldehyde inhibition and cascade reaction. A: time course of

enzymatic demethylation of 5 mM vanillic acid under standard conditions (green) and with addition of 2 mM formaldehyde (white). Dashed lines indicate vanillic acid depletion and solid lines product formation. B: Cascade reaction for removal of formaldehyde together with cofactor regeneration. FADH, formaldehyde dehydrogenase; FDH, formate dehydrogenase. C: Bioconversion of 5 mM vanillic acid with three different NADH regeneration systems and 15 mg/mL of freeze dried cells. Standard conditions with PTDH (green); one-step cascade with 6 U/mL FADH (blue); and two-steps cascade with 6 U/mL FADH and 10 U/mL FDH (purple). 0.5 mM NADH was employed for the conversion with FADH and FDH. Dashed lines indicate vanillic acid depletion and solid lines product formation.

Conclusions

The two-component enzymatic system VanA-VanB from

Pseudomonas sp. HR199 was recombinantly coexpressed in E. coli in good amounts. The resulting biocatalyst governed oxidative

demethylation of methoxy ethers shows high conversion and high selectivity for small aromatic acids with methoxy groups at the

meta position on the aromatic ring. VanA-VanB conveniently

accepts both NADH and NADPH as cofactor. After a semi-preparative scale reaction, it was possible to isolate PCA in good yields. Finally, as proof of concept, multi-enzymatic cascades were established, demonstrating in situ cofactor regeneration

10.1002/cbic.201800594

Accepted

Manuscript

ChemBioChem

(9)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

coupled to simultaneous byproduct removal from the reaction

mixture.

The results suggest that VanA-VanB has good potential as biocatalyst for the removal of methoxy groups from lignin degradation products and other compounds, offering a valuable green tool for lignin valorization and synthetic chemistry. For future practical applications, structural studies would help enzyme engineering for the improvement of stability and substrate specificity of the biocatalyst.

Experimental Section

General

Solvents and organic substrates were purchased from Sigma Aldrich, TCI Europe and Abcr GmbH unless otherwise specified. NAD+ and NADH were acquired from Carl Roth. Molecular Biology reagents were purchased from New England Biolabs or ThermoFischer Scientific. DNA isolation and purification kits were obtained from Qiagen.

Cloning

The synthetic genes of VanA (UniprotKB O05616) and VanB (UniprotKB O05617) optimized for E. coli were purchased from GenScript. BsaI sites suitable for Golden Gate cloning were inserted up and downstream of the CDS by PCR. pET28(b)_GG was developed as universal and versatile vector for an easy and highly efficient cloning of multiple fragments using the Golden Gate method. The vector was prepared by Gibson cloning using pET28(b)+ as template and a stuffer fragment was inserted. The two CDS flanked by BsaI sites and a designed synthetic fragment containing an additional T7 promoter were cloned into the pET28_GG vector in a one pot reaction by Golden Gate cloning. The final product pE2T_Ps(HR199)VanA_VanB was transformed in E. coli NEB 10 Beta and positive clones were selected on LB kanamycin (50 mg/L). Primers, were obtained from Sigma. Gene fragments were purchased from Integrated DNA Technology (gBlock) and ThermoFisher Scientific (GeneArt strings). Details are reported in Supporting Information. Phosphite dehydrogenase was recombinantly expressed and purified as previously described.[41]

Coexpression of VanA and VanB and biocatalyst preparation A previously reported protocol was optimized for the heterologous coexpression of VanA and VanB in E. coli.[31] Competent cells of E. coli BL21 Star (DE3) were transformed with the pE2T_Ps(HR199)VanA_VanB vector. A sample of 200 µL of regenerated culture was immediately added to 50 mL LB media supplemented with kanamycin (50 mg/L of culture) followed by overnight growth at 37°C. Then, 5 mL of pre-culture were transferred to 0.5 L TB medium with a double amount of glycerol (8% v/v) and cultivation was done with kanamycin selection at 37°C under strong agitation. At OD600 1, cells were cooled at 20°C for 20-30 min and then protein expression was induced by addition of 0.5 mM IPTG, FeSO4 (0.1 mg/mL), Fe(III) ammonium citrate (0.1 mg/mL) and 1 mM cysteine. Protein production was performed at 20°C for 20 h under moderate shaking. At the end of the cultivation, cells were harvested by centrifugation (4000 rpm, 4°C, 10 min) and washed once with 50 mM Tris-HCl, pH 7.5. Next, the cell pellet was resuspended in 100 mM Tris-HCl, pH 7.5, (250 mL) and shock frozen drop by drop in liquid nitrogen. Finally, frozen cell-drops were dried for 48 h (ice condenser CHRIST Alpha 2-4 LDplus). The obtained fluffy powder (0.5 g of dried cells/g of powder) was kept at -20°C or at -80°C for longer storage.

Biocatalytic reaction

Stock solutions of 50 mM substrate were prepared in DMSO and stored at 4°C. Reaction mixtures (2 mL) were set up as follows unless otherwise indicated. Samples containing 15, 30 or 60 mg of freeze-dried cells expressing VanA-VanB (corresponding to 30, 60 or 120 mg of powder) were weighted into a 15 mL glass vial and allowed to rehydrate in 100 mM Tris-HCl pH 7.5 containing 10 mM DTT for 15-20 min at 30°C and 100 rpm (shaker: Innova 40, New Brunswick). Afterwards, 0.1 mM FeSO4, 10 mM phosphite, 0.01 mM PTDH and 5 mM of methoxy substrate were added to the mixture. The reaction was initiated by the addition of 0.1 mM NADH and it was incubated at 30°C at 100 rpm. Each experiment was performed in duplicate.

HPLC analysis

At different time points 200 µL of the reaction mixture was quenched with 200 µL of 100% acetonitrile (Boom B.V.). The sample was centrifuged for 10 mins at 16,000 g at 4°C and 200 µL of supernatant were transferred to an HPLC vial. Analysis was performed with Jasco 980-31 HPLC apparatus, Jasco 2075 Plus UV detector and Jasco MD-2010 Plus PDA detector. An Alltima HP C18 5u length 250 mm ID 4.6 mm column was used for sample separation. A solution of 2 mM sodium acetate containing 0.05% formic acid (Eluent A) and 100% acetonitrile (Eluent B) were chosen as mobile phase with the following method: 0.5 mL/min flow; 5-10 µL injection; 5 min isocratic 95% A – 5% B; 15 min gradient to 5% A – 95% B; 5 min isocratic 5% A – 95% B; 5 min gradient to 95% A – 5% B; 10 min isocratic 95% A – 5% B. Substrate and products were detected at 254 nm and 280 nm. Preparative scale and product isolation

A 12 mL reaction mixture was set up in a 50 mL bottle. An amount of 720 mg of biocatalyst powder (30 mg of freeze-dried cells expressing VanA-VanB per mL of reaction mixture) was rehydrated in 100 mM Tris-HCl pH 7.5 (7.31 mL) and 10 mM DTT (0.120 mL of 1 M stock) at 30°C and 100 rpm. Next, the mixture was supplemented with 0.1 mM FeSO4, 10 mM phosphite and 0.025 mM PTDH. Finally, the reaction was initiated with 0.2 mM NADH and 10 mg vanillic acid. The mixture was incubated overnight at 30°C at 100 rpm. The complete substrate consumption was confirmed by HPLC. Then, the reaction mixture was quenched by addition of a 10% aqueous solution of citric acid, adjusted to pH~3 and extracted with ethyl acetate (3 times). The combined organic extracts were washed with brine, dried with anh. MgSO4, filtered and concentrated under reduced pressure. The crude product was purified by flash chromatography on silica gel SiliaFlash P60 40–63 µm (230-400 mesh) (SiO2; dichloromethane/methanol = 95:5). NMR spectra were recorded on Agilent Technologies 400 NMR (1H NMR at 400 MHz, 13C NMR at 100 MHz). Chemical shifts are expressed in ppm (δ).

Acknowledgements

This work was supported by the European Commission (Marie Skłodowska Curie Initial Research Training Network SUBICAT, Call FP7-PEOPLE-2013-ITN, grant no. 607044) and the BE-Basic R&D Program (http://www.be-basic.org/), which is financially supported by an EOS Long Term grant from the Dutch Ministry of Economic Affairs, Agriculture and Innovation (EL&I). The authors thank Dr. Ulrich Commandeur and Selin Ece (RWTH Aachen University) for providing E. coli BL21 Star (DE3) strain, and Dr. Peter Deuss (University of Groningen) for his inputs.

Accepted

(10)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

The authors declare no financial or commercial conflict of interest.

Keywords: biocatalysis • methoxy group • O-demethylation • protein expression • Rieske monooxygenase

[1] P. G. M. Wuts, T. W. Greene, in Greene’s Prot. Groups Org. Synth., John Wiley & Sons, Inc., 2006, pp. 367–430.

[2] L. Zuo, S. Yao, W. Wang, W. Duan, Tetrahedron Lett. 2008, 49, 4054–4056.

[3] S. B. Waghmode, G. Mahale, V. P. Patil, K. Renalson, D. Singh,

Synth. Commun. 2013, 43, 3272–3280.

[4] G. de Gonzalo, D. I. Colpa, M. H. M. Habib, M. W. Fraaije, J.

Biotechnol. 2016, 236, 110–119.

[5] A. J. Ragauskas, G. T. Beckham, M. J. Biddy, R. Chandra, F. Chen, M. F. Davis, B. H. Davison, R. A. Dixon, P. Gilna, M. Keller, et al.,

Science (80-. ). 2014, 344, 1246843.

[6] W. Schutyser, T. Renders, S. Van den Bosch, S.-F. Koelewijn, G. T. Beckham, B. F. Sels, Chem. Soc. Rev. 2018, 47, 852–908. [7] Z. Sun, B. Fridrich, A. de Santi, S. Elangovan, K. Barta, Chem. Rev.

2018, 118, 614–678.

[8] P. C. A. Bruijnincx, B. M. Weckhuysen, Nat. Chem. 2014, 6, 1035– 1036.

[9] P. J. Deuss, M. Scott, F. Tran, N. J. Westwood, J. G. de Vries, K. Barta, J. Am. Chem. Soc. 2015, 137, 7456–7467.

[10] R. Jastrzebski, E. J. van den Berg, B. M. Weckhuysen, P. C. A. Bruijnincx, Catal. Sci. Technol. 2015, 5, 2103–2109.

[11] T. Sonoki, K. Takahashi, H. Sugita, M. Hatamura, Y. Azuma, T. Sato, S. Suzuki, N. Kamimura, E. Masai, ACS Sustain. Chem. Eng.

2018, 6, 1256–1264.

[12] N. Richter, F. Zepeck, W. Kroutil, Trends Biotechnol. 2015, 33, 371– 373.

[13] S. G. Bell, A. B. H. Tan, E. O. D. Johnson, L.-L. Wong, Mol. BioSyst.

2009, 6, 206–214.

[14] S. G. Bell, R. Zhou, W. Yang, A. B. H. Tan, A. S. Gentleman, L.-L. Wong, W. Zhou, Chem. Eur. J. 2012, 18, 16677–16688. [15] T. Coleman, R. R. Chao, J. J. De Voss, S. G. Bell, Biochim.

Biophys. Acta - Proteins Proteomics 2016, 1864, 667–675.

[16] R. R. Chao, I. C.-K. Lau, J. J. De Voss, S. G. Bell, ChemCatChem

2016, 8, 3626–3635.

[17] T. Abe, E. Masai, K. Miyauchi, T. Abe, E. Masai, K. Miyauchi, Y. Katayama, M. Fukuda, J. Bacteriol. 2005, 187, 2030–2037. [18] E. Rosini, P. D’Arrigo, L. Pollegioni, Catal. Sci. Technol. 2016, 6,

7729–7737.

[19] A. Harada, N. Kamimura, K. Takeuchi, H. Y. Yu, E. Masai, T.

Senda, FEBS J. 2017, 284, 1855–1867.

[20] A. C. Kohler, M. J. L. Mills, P. D. Adams, B. A. Simmons, K. L. Sale,

Proc. Natl. Acad. Sci. 2017, 114, E3205–E3214.

[21] S. J. B. Mallinson, M. M. Machovina, R. L. Silveira, M. Garcia-Borràs, N. Gallup, C. W. Johnson, M. D. Allen, M. S. Skaf, M. F. Crowley, E. L. Neidle, et al., Nat. Commun. 2018, 9, 2487. [22] F. Brunel, J. Davison, J. Bacteriol. 1988, 170, 4924–4930. [23] J. A. Buswell, D. W. B. T.-M. in E. Ribbons, in Biomass Part B

Lignin, Pectin, Chitin, Academic Press, 1988, pp. 294–301.

[24] V. Venturi, F. Zennaro, G. Degrassi, B. C. Okeke, C. V. Bruschi,

Microbiology 1998, 144, 965–973.

[25] R. L. D’Ordine, T. J. Rydel, M. J. Storek, E. J. Sturman, F. Moshiri, R. K. Bartlett, G. R. Brown, R. J. Eilers, C. Dart, Y. Qi, et al., J. Mol.

Biol. 2009, 392, 481–497.

[26] R. Dumitru, W. Z. Jiang, D. P. Weeks, M. A. Wilson, J. Mol. Biol.

2009, 392, 498–510.

[27] W. Jiang, M. A. Wilson, D. P. Weeks, ACS Chem. Biol. 2013, 8, 1687–1691.

[28] D. J. Ferraro, L. Gakhar, S. Ramaswamy, Biochem. Biophys. Res.

Commun. 2005, 338, 175–190.

[29] S. M. Barry, G. L. Challis, ACS Catal. 2013, 3, 2362–2370. [30] M. Nishimura, Y. Nishimura, C. Abe, M. Kohhata, 2014, 37, 1564–

1568.

[31] S. Jaganaman, A. Pinto, M. Tarasev, D. P. Ballou, Protein Expr.

Purif. 2007, 52, 273–279.

[32] J. Rabenhorst, Appl. Microbiol. Biotechnol. 1996, 46, 470–474. [33] J. Overhage, H. Priefert, A. Steinbuchel, Appl. Environ. Microbiol.

1999, 65, 4837–4847.

[34] H. Priefert, J. Rabenhorst, A. Steinbuchel, J. Bacteriol. 1997, 179, 2595–2607.

[35] M. Tarasev, A. Pinto, D. Kim, S. J. Elliott, D. P. Ballou, Biochemistry

2006, 45, 10208–10216.

[36] C. C. Correll, C. J. Batie, D. P. Ballou, M. L. Ludwig, Science (80-. ).

1992, 258, 1604–1610.

[37] T. W. Johannes, R. D. Woodyer, H. Zhao, Society 2005, 71, 5728– 5734.

[38] M. Hibi, T. Sonoki, H. Mori, FEMS Microbiol. Lett. 2005, 253, 237– 242.

[39] E. M. T. El-Mansi, S. C. K. Anderson, World J. Microbiol. Biotechnol.

2004, 20, 827–832.

[40] B. Morawski, A. Segura, L. N. Ornston, J. Bacteriol. 2000, 182, 1383–1389.

[41] N. Beyer, J. K. Kulig, A. Bartsch, M. A. Hayes, D. B. Janssen, M. W. Fraaije, Appl. Microbiol. Biotechnol. 2017, 101, 2319–2331.

10.1002/cbic.201800594

Accepted

Manuscript

ChemBioChem

(11)

FULL PAPER

For internal use, please do not delete. Submitted_Manuscript

FULL PAPER

Enzymes for free phenols: The enzymatic demethylation system VanA-VanB from Pseudomonas was characterized for the oxidative demethylation of methoxy aromatic acids into catechols and phenols. Three different cascade options for cofactor regeneration paired to byproduct removal are proposed, offering perspectives for innovative biocatalytic methods for the dealkylation of aryl-methyl ethers.

E Lanfranchi,* M Trajkovic, K Barta, JG de Vries, DB Janssen

Page No. – Page No. Title

Accepted

Referenties

GERELATEERDE DOCUMENTEN

8 TABLE 8: LINEAR REGRESSION RESULTS FOR THE VARIABLES OF THE FORMAL INSTITUTIONS PERSPECTIVE USING FDI/GDP AS DEPENDENT VARIABLE.

Specifically, the study examined whether perceived Twitter brand account features (information quality, entertainment, vividness and interactivity) predicted the

The w lines following 1c, 2c, and 3c in the listing show the minimum column widths specified by the ‘w’ keys in the format; 35 pt is 7 times TABLE’s default column width unit of 0.5

The LaTeX package decision-table provides a command \dmntable, which allows for an easy way to generate decision tables in the Decision Model and Notation (DMN) format. 1 ) This

This example is based upon a nice example in the Pythontex gallery,

Section 16: Binomial coefficients and probabilities ⋆ Section 17: Tossing coins on a computer, part 1 ⋆ Section 18: Tossing coins on a computer, part 2 ⋆⋆ Section 19: Statistics,

IBP, inflammatory back pain; NSAIDs, Non-Steroidal Anti-Inflammatory Drugs; IBD, inflammatory bowel disease; HLA-B27, Human Leucocyte Antigen B27; ESR, erythrocyte sedimentation

Marie-Charlotte Ibanez, Judith Berendsen Ex 5.1.8 generator Ornstein-Uhlenbeck May 20 Matteo Quattropani, Giulia Pederzani Lemma 5.2.10 plus analogy with 4.4.5. Jian-He, Xavier