Abstract
Many aldehydes are volatile compounds with distinct and characteristic olfactory properties. The aldehydic functional group is reactive and, as such, an invaluable chemical multi‐tool to make all sorts of products. Owing to the reactivity, the selective synthesis of aldehydic is a challenging task. Nature has evolved a number of enzymatic reactions to produce aldehydes, and this review provides an overview of aldehyde‐forming reactions in biological systems and beyond. Whereas some of these biotransformations are still in their infancy in terms of synthetic applicability, others are developed to an extent that allows their implementation as industrial biocatalysts.
Keywords: aldehyde, aliphatic aldehydes, aromatic aldehydes, aryl‐aliphatic aldehydes, biocatalysis, enzymes, green leaf volatiles, vanillin
Nature has evolved a number of enzymatic reactions to produce aldehydes, and this review provides an overview of aldehyde‐forming reactions in biological systems and beyond.

1. INTRODUCTION
Aldehydes are reactive compounds and can undergo chemical transformations to numerous other functional groups. 1 The aldehyde is therefore an invaluable chemical multi‐tool to make all sorts of products. 2 As final products, aldehydes find application in the flavour and fragrance sector, because they are often volatile with characteristic olfactory properties. The most abundant biomass‐derived aldehydes are furfural from cellulose, 5‐hydroxymethylfurfural (HMF, 5‐(hydroxymethyl)‐2‐furaldehyde) from hemicellulose, aromatic aldehydes like vanillin and syringaldehyde from lignin 3 as well as short‐, medium‐ and long‐chain aldehydes from oil and fat‐derived (polyunsaturated) fatty acids. Lignin, for example, is a largely untapped resource, since nowadays it is primarily burnt to generate heat. 4
The flavour and fragrance industry is interested in these kinds of molecules, and enzymatic and biotechnological processes towards ‘bio‐aldehydes’ are gaining relevance. 2 Valuable compounds can be produced from renewable resources based, for example, on the degradation of lignin, that produces monolignols. Biotechnological production decreases the dependence on plant material, whose availability greatly varies with season, weather, political developments, need as food or feed and many factors more. While in the consumer's eye, natural ingredients are clearly favoured over chemically synthesised ingredients, biocatalysis, 5 biotechnology and metabolic engineering may close gaps, especially for desirable compounds from rare and endangered species. The ‘naturality’ of ingredients is a richly facetted chapter. 6 Whether an ingredient can be labelled ‘natural’ or not is a matter of definitions and regulations. Naturality is not easily measurable. It is clear that an apple is natural, but when this apple is processed to generate a flavour ingredient, it depends on each processing step and its associated auxiliaries whether the final ingredient can still be considered natural. Phrasing it in a very simplistic way, the ‘greenness’ of each processing step helps to preserve the naturality status of an ingredient. Mild and biobased methodologies play a key role in this context.
Nature is equipped with a portfolio of proteins with enzymatic activity to transform various precursor molecules to the respective aldehydes, and these are the focus of this review (Scheme 1).
SCHEME 1.

Enzymatic reactions to aldehydes. Aldehyde‐forming aldolases were omitted due to their need for aldehydes as substrates.
2. ENZYME CATALYSED ALDEHYDE FORMATION
2.1. From carboxylic acids
Carboxylic acids are stable and abundant in biomass and they can be produced in good amounts by acidogenic fermentation. 7 Volatile organic acids are, for example, accessible by anaerobic fermentation from organic wastes. A broad spectrum of fatty acids is available from fats and oils, including also waste streams from edible oil production. Such carboxylic acids are a stable and renewable pool of aldehyde precursors.
Two distinct enzyme families are known to reduce acids to aldehydes: carboxylic acid reductases (CARs) and aldehyde oxidoreductases (AORs).
2.1.1. Aldehyde oxidoreductase
Aldehyde oxidoreductases (EC 1.2.99.6) catalyse the reduction of a carboxylate to an aldehyde and the reverse reaction. 8 AORs exhibit a relatively broad substrate spectrum, using short‐chain aliphatic aldehydes, branched‐chain aliphatic aldehydes and aromatic aldehydes and their corresponding acids (Scheme 2). AORs are oxygen sensitive, tungsten containing enzymes from bacterial and archaeal origin. They harbour Fe–S clusters and operate under strictly anaerobic conditions in the presence of redox mediators. 9 , 10 Due to this sensitivity, AORs have mainly been studied in the form of whole‐cell biocatalysts in their natural hosts (e.g. Pyrococcus furiosus DSM 3638) or as purified enzymes from their native organisms. In the strictly anaerobic natural hosts, further reduction of the emerging aldehyde to alcohols was predominant. Aldehyde:ferredoxin oxidoreductase, for example, reduces acetate to acetaldehyde via reduced ferredoxin as a redox carrier. 11 In P. furiosus, hydrogen (H2) is oxidised by a hydrogenase enzyme and serves as a reducing equivalent. 12 A major scientific challenge that—to the best of our knowledge—has not been solved so far is to functionally express AORs in a heterologous host such as Escherichia coli or yeast.
SCHEME 2.

Aldehyde oxidoreductases AORs (EC 1.2.99.6) reduce short‐chain carboxylic acids.
2.1.2. Carboxylic acid reductase
Carboxylic acid reductases (EC 1.2.1.30) also catalyse the single‐step reduction of a carboxylate to an aldehyde (Scheme 3). The reverse reaction has not been reported. CARs are relaxed regarding structural features of the acids they reduce and can cope with aliphatic, aromatic, heteroaromatic and aryl‐aliphatic mono‐ and di‐carboxylates. Halogen, hydroxy, methyl, methoxy and amino substitutions are tolerated well, in case they are two carbon atoms or more distant to the carboxylic acid moiety. 13 Due to their broad product scope, CARs hold great promise as a tool for the synthesis of various aldehydes.
SCHEME 3.

(A) Carboxylic acid reductase (CAR EC 1.2.1.30) mediated enzymatic aldehyde synthesis; for application in cell‐free systems, ATP and NADPH are recycled. (B) Isolated products form cell‐free aldehyde synthesis.
Acid activation is accomplished directly in the CAR enzyme by adenosine triphosphate (ATP), and the nicotinamide cofactor NADPH serves as the reductant (Scheme 3). These cofactors are required in stoichiometric amounts. When CARs are utilised as isolated enzymes, cofactor supply can be accomplished by catalytic amounts of ATP and NADPH and their recycling in vitro (Scheme 3). Well‐established oxidoreductase recycling systems are available for NADPH (e.g. glucose dehydrogenase/glucose). ATP recycling is typically accomplished with kinase enzymes at the expense of polyphosphate. The reader is referred to a recent review by Tavanti et al. 14 for detailed information. New kinases are currently becoming available, as ATP recycling is in the research focus of both academia 15 , 16 and industry. 17 For aldehyde synthesis, the in vitro strategy has been shown successful, for example, for the preparation of 4‐methoxy‐benzaldehyde (anisaldehyde) 17 and N‐carbobenzoxylated 4‐formylpiperidine 18 on a gram scale (Scheme 3). Vanillin can be produced in 2.86 g L−1 concentration using an E. coli‐based whole‐cell biocatalyst equipped with Mycobacterium abscessus CAR MaCAR2, 19 or as vanillyl‐glycoside on industrial scale in yeast. 20
2.2. From thioesters
Thioesters are carboxylic acid derivatives which are formed in living systems in the course of the biosynthesis of fatty acids, polyketides and non‐ribosomal peptides, and other related metabolic processes. The thio‐donor in living systems is Coenzyme A (CoA), or the phosphopantetheine unit of CoA that is attached to protein. Also, proteinogenic cysteines may function as thio‐donors.
2.2.1. Thioester reductase and thioester reductase domains in larger proteins
Thioester reductases (TERs) catalyse the NAD(P)H‐dependent reduction of thioesters to aldehydes. Acyl protein thioester reductases (EC 1.2.1.50, EC 1.2.1.80) produce long‐chain aldehydes from protein‐bound fatty acids. Fatty aldehydes are furthermore accessible through CoA‐bound fatty acids, which are reduced by fatty acyl‐CoA reductases (FARs, EC 1.2.1.42 and EC 1.2.1.B25). Short‐chain aldehydes as primary metabolites are the products of several enzyme classes which are typically named according to their natural product: acetaldehyde dehydrogenase (EC 1.2.1.10) produces acetaldehyde from acetyl‐CoA. Others act on CoA‐bound malonates (EC 1.2.1.18, EC 1.2.1.27 and EC 1.2.1.75), to deliver malonate‐semialdehyde or CoA‐bound succinate (EC 1.2.1.76), or glyoxylates (EC 1.2.1.17, EC 1.2.1.58), respectively.
Thioester reductases also occur as domains in multi‐domain proteins. One example thereof was outlined for CARs (chapter 2.1.2), in which the R‐domain is in fact catalysing thioester reduction. Similarly, NRPS and PKS with thioester reductase domains were reported to produce complex aldehydes 21 , 22 and amino aldehydes which eventually tend to condense to give pyrazines. 23 , 24 Thioester reductase domains may be promiscuous in the sense that not only the thioester delivered by its associated transthiolation domain is reduced, but also unbound surrogates, albeit with less efficiency. 25
Notably, 3‐ketoacyl‐thioester reductase (EC 1.1.1.100) and 2‐enoyl (thioester) reductases (EC 1.3.1.10 and EC 1.3.1.38) do not produce aldehydes but act on other reducible groups in protein‐bound thioesters.
Due to the need for coenzyme A or a particular protein for thioester formation, thioester reductases are mostly targeted in metabolic engineering campaigns. Aldehydes themselves are rarely the desired product but appear as transient key components, for example, for the production of the corresponding alcohols. 26 An exception is the synthesis of cinnamaldehyde, which is generated, for example, by a cinnamoyl‐CoA reductase from Arabidopsis thaliana (Scheme 4) as the final product. 27
SCHEME 4.

Last step of the biosynthetic pathway to cinnamaldehyde catalysed by cinnamoyl‐CoA reductase (CCR).
2.2.2. Aldehyde dehydrogenase and aldehyde reductase
The thioester may also be formed by a proteinogenic cysteine. A recent publication showed that the adenosine monophosphate (AMP)‐anhydride of hydroxylated 2,2′‐bipyridine‐1‐carboxylate reacted with a C‐terminal cysteine of CaeB2, a protein with high similarity to NADPH‐dependent aldehyde dehydrogenases (ALDHs). CaeB2 hence catalysed the transthiolation of an AMP‐anhydride followed by NADPH‐dependent thioester reduction to 2,2′‐bipyridine‐1‐carbaldehyde. 28
2.3. From primary alcohols
Alcohols are an abundant source of compounds from renewable sources, already widely used in many industries. Higher molecular weight alcohols (e.g. geraniol, cuminol, 2‐phenylethanol) are often isolated from plant material by steam distillation. Fatty alcohols are usually obtained from fats and waxes by base hydrolysis followed by reduction, but can also be produced by microbial fermentation. 29
The most commonly used enzymes for the oxidation of alcohols are alcohol dehydrogenases (ADHs) and alcohol oxidases (AOxs). Peroxidases and oxygenases can also be used but have received little attention because of narrow substrate scope or low selectivity. 30 , 31
2.3.1. Alcohol oxidase
Alcohol oxidases (EC 1.1.3.x) oxidise alcohols to aldehydes or ketones using O2 as a terminal electron acceptor. The reaction produces H2O2 as a side product. To avoid enzyme deactivation, H2O2 is usually removed in situ by an accessory enzyme: a catalase or a peroxidase. Thus, the net reaction becomes irreversible (Scheme 5). In addition, horseradish peroxidase (HRP) can be conveniently coupled with 2,2′‐azino‐bis(3‐ethylbenzothiazoline‐6‐sulfonic acid) (ABTS) as an electron acceptor in a spectrophotometric assay for screening purposes. 32 Except for some cofactor‐lacking oxidases, electron transfer is mediated by a tightly bound redox‐active prosthetic group, usually a copper or FAD. 32 , 33 A major advantage of the AOx over ADH enzymes is the lack of dependence on the costly NAD(P)+ cofactor. The major drawbacks are their comparably lower abundance in nature and often they are of eukaryotic origin, making recombinant expression more challenging.
SCHEME 5.

Oxidation of primary alcohols catalysed by alcohol oxidase (AOx).
Copper radical alcohol oxidase
Copper radical alcohol oxidases (CRO‐AOxs) contain a single copper(I) ion bound in a shallow active site. Substrate oxidation involves the formation of a tyrosine radical that is stabilised by a cross‐linked tyrosine–cysteine residue and ultimately forms a metalloradical complex. 32 , 34
The archetypal CRO‐AOx is galactose oxidase (GAOx, EC 1.1.3.9), whose natural activity is oxidation of the C‐6 OH group of d‐galactose to give the aldehyde d‐galacto‐hexodialdose (Scheme 6). 35 The enzyme is highly regioselective and next to galactose itself, is able to oxidise galactose‐containing saccharides such as lactose. The recognition motif of the enzyme seems to be the C‐4 to C‐6 diol, as certain primary alcohols like, for example, the N‐acetylalcohol of N‐glycolylneuraminic acid (Neu5Gc) are also oxidised.
SCHEME 6.

Oxidation of the primary alcohol of galactose by the galactose oxidase. The substrate recognition motif is highlighted in red.
Galactose oxidase enzymes are applied for the functionalization of saccharides, as biosensors and as key enzymes in various enzymatic cascades designed for the transformation of HMF into furandicarboxylic acid (FDCA). 36 , 37 , 38 In this context, (di)aldehydes are transitional intermediates. Overoxidation of aldehydes to carboxylates was observed as a side reaction of GAOx, and was also purposefully engineered into GAOx enzymes. 39
Efforts to expand the substrate scope and increase the catalytic efficiency of GAOx have predominantly been carried out on the GAOx from the phytopathogenic fungus Fusarium graminearum, which was the first described and remains the most prominent member of the CRO‐AOx family. The substrate scope was expanded not only to other carbohydrates such as glucose, 40 fructose, 41 mannose and N‐acetylglucosamine, 42 but also secondary alcohols, 43 amino alcohols 44 and benzyl alcohols. 39 Notably, a highly evolved variant was developed for industrial biocatalytic cascade synthesis of the drug islatravir. 45
In the last decade, new types of CRO‐AOx enzymes were found based on the structural similarity to the GOx, but with surprisingly low activity towards carbohydrates. Most of these enzymes were identified in the phytopathogenic genera Colletotrichum and Fusarium, and catalyse the oxidation of diverse activated and non‐activated aliphatic and aromatic alcohols—activity that classifies them either as specific alcohol oxidases (EC 1.1.3.13) or aryl alcohol oxidases (AAOx, EC 1.1.3.7). 46 , 47 , 48 , 49 , 50 , 51
Alcohol oxidases from Colletotrichum graminicola (CgrAOx) and Colletotrichum destructivum (CdeAOx) were used for oxidation of C6 and C8 primary aliphatic alcohols into the corresponding fragrance aldehydes. Upon longer reaction times, overoxidation of >C6 substrates to corresponding carboxylic acids was observed and associated to the propensity of the aldehyde to undergo hydration and form a geminal diol. The presence of an electron‐withdrawing group increased aldehyde overoxidation (24% aldehyde hydrate formation in aqueous solution), while an electron‐donating group produced only aldehyde (0% hydrate formation). Enzyme inhibition by the long‐chain fatty aldehyde hydrates was also described. 50 , 51 A gram‐scale reaction using a crude preparation of CgrAOx yielded 0.72 g of octanal (Scheme 7A). 51 CgrAOx was also applied in a sequential one‐pot cascade for oxidizing geraniol to geranial, which was subsequently reduced by an ene‐reductase to (R)‐ or (S)‐citronellal. A quantity of 62 mg of geraniol was converted to (R)‐citronellal with 98% conversion and 72% isolated yield (Scheme 7B). 52
SCHEME 7.

Primary alcohol oxidations catalysed by the CRO‐AOx from Colletotrichum graminicola (CgrAOx). (A) Oxidation of octan‐1‐ol to octanal. (B) Oxidation of geraniol to geranial by CgrAOx followed by C=C reduction by enoate‐reductase OYE2 to produce (R)‐citronellal.
The overoxidation phenomenon does not only depend on the propensity of the aldehyde to form the aldehyde hydrate, but also on the ability of the specific enzyme to accommodate the geminal diol in the active site. For example, while the AAOx from Fusarium species overoxidizes HMF to 5‐formyl‐2‐furoic acid (FFCA), respectively, 48 the AAOx from C. graminicola can chemoselectively oxidise HMF to 2,5‐diformylfuran (Scheme 8). 47
SCHEME 8.

Oxidations of HMF catalysed by CRO‐AOx aryl alcohol oxidases (AAOx). The AAOx from Fusarium species (FgrAAOx and FoxAAOx) overoxidize HMF to 5‐formyl‐2‐furoic acid (FFCA). The AAOx from Colletotrichum graminicola (CgrAAOx) produces the di‐aldehyde 2,5‐diformylfuran (DFF) without overoxidation.
Flavin‐dependent alcohol oxidase
Flavin‐dependent alcohol oxidases (FAD‐AOx; EC 1.1.3.13) contain a tightly, sometimes even covalently, bound flavin cofactor (generally FAD) that serves as the primary hydride acceptor. Substrate oxidation generates a reduced flavin which is re‐oxidised using molecular oxygen via a hydroxyperoxyflavin intermediate, generating H2O2. Often, the oxidation of primary alcohols does not stop at the aldehyde stage, but proceeds to the corresponding carboxylic acid. 32
Based on the structural fold, these enzymes can be classified into two main families: the glucose‐methanol‐choline oxidase family (known as the GMC family) and the vanillyl‐alcohol oxidase family (known as the VAO family) which can further be divided based on the presence and number of covalent bonds present between the FAD cofactor and the protein part of the specific oxidase. 53
Members of the GMC family include short‐chain alcohol oxidases (SCAOx), long‐chain alcohol oxidases (LCAOx) and aryl alcohol oxidases (AAOx). SCAOxs are key enzymes of methanol metabolism of methylotrophic yeasts such as Komagataella phaffii (Pichia pastoris). The AOx from K. phaffii has high activity towards short‐chain aliphatic alcohols (methanol, ethanol), but activity decreases with increasing chain length. However, the oxidation of longer chains (C6–C11) and aromatic alcohols is possible in biphasic systems. 54 , 55
An interesting enzyme belonging to the GMC family is the 5‐hydroxymethylfurfural oxidase from Methylovorus sp. strain MP688 (EC 1.1.3.47). The enzyme is highly selective in accepting only hydrated aldehydes, primary alcohols and thiols as substrates and at the same time strikingly promiscuous concerning the side chain. Except for HMF, it accepts various furanic, benzylic and cinnamic alcohols, with ring substituents of different sizes, polarities and charges, oxidizing them to carboxylic acids. 56 , 57
In terms of aldehyde synthesis, overoxidation is undesirable, but it may be eliminated through enzyme engineering, as shown for the choline oxidase (CO, EC 1.1.3.17) from Arthrobacter chlorophenolicus which catalyses the two‐step oxidation of choline into trimethylglycine. Structure‐guided directed evolution of the AcCO produced a six amino acid variant (AcCO6) with much broader substrate specificity towards a range of primary alcohols and increased thermostability and solvent tolerance. Saturated alcohols with C6–C10 chain lengths and unsaturated alcohols were especially good substrates. Hexan‐1‐ol (100 mg) was completely converted to hexanal in 72% isolated yield (Scheme 9). Overoxidation was still detected with other substrates (e.g. cinnamyl alcohol) but could be minimized by applying a biphasic system. 58 In a follow‐up study, oxidation of hexan‐1‐ol by AcCO6 in continuous flow did lead to the overoxidation to the carboxylic acid when performed in buffer, but could be completely avoided by using neat solvent as the reaction medium. 59
SCHEME 9.

Oxidation of hexan‐1‐ol to hexanal catalysed by the sixfold variant of choline oxidase from Arthrobacter chlorophenolicus (AcCO6).
Aryl alcohol oxidase from the fungus Pleurotus eryngii (PeAAOx) was used for biocatalytic preparation of the green note aroma compounds trans‐hex‐2‐enal from trans‐hex‐2‐en‐1‐ol (Scheme 10). Initially, biocatalytic synthesis of 200 mg of trans‐hex‐2‐enal in a continuous flow reactor was reported, with turnover frequencies up to 38 s−1 and turnover numbers >300,000. 60 A follow‐up study described a two‐liquid phase system approach using the substrate itself as the organic phase. trans‐Hex‐2‐enal accumulated over 14 days with multiple additions of PeAAOx and catalase to 255 g L−1. Although the conversion was only 31%, the total turnover number of PeAAOx reached >2.2 × 106. 61
SCHEME 10.

Oxidation of trans‐hex‐2‐en‐1‐ol to trans‐hex‐2‐enal catalysed by the aryl alcohol oxidase from Pleurotus eryngii (PeAAOx).
Another aryl alcohol oxidase was found in the same fungus (PeAAOx2) and shown to oxidise cumic alcohol and piperonyl alcohol to cuminaldehyde and piperonal with high catalytic efficiencies of 84.1 and 600.2 mM−1 s−1, respectively. 62 In a follow‐up study, conditions for a preparative scale reaction were optimized to reach a space–time yield of 9.5 g L−1 h−1 for piperonal (244.6 mg; 85% yield) and further applied for the oxidation of cumic alcohol, thiophene‐2‐yl‐methanol, (E,E)‐2,4‐heptadienol and (E)‐2‐(Z)‐6‐nonadienol (Figure 1). Up to 300 mM of these substrates were converted to the corresponding aldehydes within 20 h. 63
FIGURE 1.

Aldehydes produced by the oxidation of the corresponding primary alcohols by the aryl alcohol oxidase from Pleurotus eryngii (PeAAOx2).
A new thermotolerant aryl alcohol oxidase from Moesziomyces antarcticus (MaAAOx) was recently described as accepting a broad range of primary benzylic alcohols, aliphatic allylic alcohols and furan derivatives like HMF. Among oxidation products were odorous compounds such as benzaldehyde, cuminaldehyde, piperonal and perillaldehyde. Moreover, MaAAOx showed unique activity towards oxidizing 5‐hydroxymethyl‐2‐furancarboxylic acid (HMFCA) to FFCA. 64
Recently, a high‐throughput screening endeavour identified a new long‐chain alcohol oxidase (LCAOx, EC 1.1.3.20) that is active with a range of fatty alcohols, with 1‐dodecanol being the preferred substrate. 65
The VAO family is represented by its name‐giving member: the vanillyl alcohol oxidase, first isolated from the fungus Penicillium simplicissimum (PsVAO, EC 1.1.3.38). The fungal flavoenzyme can convert a variety of para‐substituted phenols and produce high‐value aromatic compounds, for example, vanillin and coniferyl alcohol. 66 Except for alcohol oxidations, the VAO‐type oxidases show a remarkable spectrum of activity: amine oxidations, hydroxylations, ether bond cleavage and even C–C bond formation, which are described elsewhere. 67 PsVAO is characterized by an autocatalytically formed covalently linked FAD. The prototype reaction is the oxidation of vanillyl alcohol to vanillin (Scheme 11), even though also other substrates can be used as vanillin precursors (e.g. vanillyl amine and creosol). Since fungal VAOs are typically poorly expressed in bacteria, a quest for a bacterial VAO identified a PsVAO homologue in Rhodococcus sp. strain RHA1, which could be expressed at high levels in E. coli. The bacterial enzyme readily accepted vanillyl alcohol and 5‐indanol as substrates, but the highest activity was observed for the hydroxylation of eugenol to coniferyl alcohol. Therefore, the enzyme was named eugenol oxidase (EUGO). 68 The optimization of biosynthesis of vanillin using EUGO lead to an organic solvent‐free process with space–time yield of 9.9 g L−1 h−1, producing 11.2 g of vanillin from a 250 mL scale reaction. Notably, overoxidation to vanillic acid was not detected under the applied conditions. 69
SCHEME 11.

Oxidation of vanillyl alcohol to vanillin catalysed by the vanillyl alcohol oxidase from fungus Penicillium simplicissimum (PsVAO) or eugenol oxidase from bacterium Rhodococcus sp. strain RHA1 (EUGO).
2.3.2. Alcohol dehydrogenase
Alcohol dehydrogenases (ADHs, EC 1.1.1.1) are classical redox catalysts, mediating the reversible transfer of a hydride from an alcohol carbon atom to its associated nicotinamide cofactor (NAD(P)+; Scheme 12). Compared to their use in reductive processes (generating chiral centres of secondary alcohols), oxidative applications are far less common.
SCHEME 12.

Oxidation of primary alcohols catalysed by alcohol dehydrogenase (ADH).
Alcohol dehydrogenases are the most intensely studied group of oxidoreductases, with numerous known enzymes. The major disadvantage of ADHs is their need for cofactor NAD(P)+ which is temporarily loosely bound, relatively unstable and expensive, therefore, a lot of research has been done on efficient cofactor recycling using a molar equivalent of a stoichiometric oxidant (e.g. acetaldehyde or acetone). However, the poor thermodynamic driving force of oxidation reactions often necessitates significant molar surplus of the stoichiometric oxidant in order to drive the reaction equilibrium in the desired direction. 70 The equilibrium can also be influenced by the pH: usually reduction is preferred at neutral pH, whereas higher pH promotes oxidation.
Despite the large number of known ADHs and studies involving them, the predominant focus is on their use for asymmetric reduction of ketones and enantioselective oxidation of secondary alcohols. Consequently, examples of ADHs applied for aldehyde production are relatively scarce. Overall, many studies reporting expression and characterization of novel ADHs report activity on primary alcohols in the initial substrate screenings, however, often without analysing the oxidation product. 71
From the well‐known ADHs, the horse liver ADH (HLADH), yeast ADH from Saccharomyces cerevisiae (YADH) and the ADH from Geobacillus stearothermophilus (BsADH) are known to have broad substrate spectra towards primary alcohols. 31 HLADH has the best activity with C8 alcohol, with activity detected up to C24, 72 and was described with, for example, benzyl alcohol 73 and amino alcohols. 74 YADH isozymes showed activity with primary unbranched aliphatic C2–C12 alcohols, as well as allyl and cinnamyl alcohol. 75 Recently, purified YADH was used in a microreactor for production of hexanal reaching volumetric productivity of 4.8 mol L−1 day−1, without overoxidation that was characteristic for whole cells of S. cerevisiae (Scheme 13). 76 Addition of a second microreactor unit enabled integrated NAD+ recycling using the same enzyme with acetaldehyde as the stoichiometric oxidant. 77
SCHEME 13.

Oxidation of hexan‐1‐ol to hexanal catalysed by alcohol dehydrogenase from Saccharomyces cerevisiae (YADH).
The BsADH is an industrially applicable enzyme that can resist ionic detergents, organic solvents, chaotropic agents, temperatures up to 65°C, but it poorly tolerates oxidative conditions. Recently, it was used for the synthesis of ω‐oxo lauric acid methyl ester (HLAMe) from the corresponding long‐chain alcohol with K M = 86 μM and specific activity of 44 U mg−1 (Scheme 14). The limited stability of BsADH under oxidation conditions was improved by a single mutation of an oxidation‐prone residue Cys257 to Leu. 78
SCHEME 14.

Oxidation of ω‐hydroxy lauric acid methyl ester to ω‐oxo lauric acid methyl ester catalysed by alcohol dehydrogenase from Geobacillus stearothermophilus (BsADH).
Lyophilized cells of Janibacter terrae were used as a catalyst for the chemoselective oxidation of primary alcohols to the corresponding aldehydes with the use of acetaldehyde as the hydrogen acceptor. The substrate spectrum encompassed substituted benzyl alcohols (e.g. producing benzaldehyde and piperonal), n‐alkanols and allylic alcohols. A total of 95% conversion of 97 mM benzyl alcohol to benzaldehyde was reported. 73 , 79 However, isolation of the responsible ADH was not described to the best of our knowledge.
Propanediol oxidoreductase from E. coli (FucO, EC 1.1.1.77) catalyses regioselective oxidation of vicinal diols to produce α‐hydroxy aldehydes. The enzyme was subjected to a directed‐evolution study to access larger aryl‐substituted α‐hydroxy aldehydes. 80
2.4. From amines
The amine functional group is ubiquitous in nature in many forms, such as amino acids and derivatives, amino alcohols, amino sugars and aliphatic mono‐ and diamines. In addition, amine functionalities are crucial in alkaloids, complex nitrogen‐containing structures (often heterocyclic) in plants. Several enzyme classes are very effective in the reversible or irreversible interconversion of primary/secondary amines and the corresponding carbonyls, and such enzymes have been extensively studied and developed. Bioderived primary amines are rare and considered more valuable than aldehydes, so the reverse reaction is much more often applied. Many amine‐converting biocatalysts have become commercially available on large scale, and a few relevant applications will be discussed in this section.
2.4.1. Transaminase/aminotransferase
Transaminases (TAs, EC 2.6.1.x) also known as aminotransferases (ATs), catalyse the transfer of an amino group from an amino donor to an amino acceptor, which contains a carbonyl group (Scheme 15). 81 Most commonly, in native enzymes, the transfer occurs between an α‐amino acid and an α‐keto acid, and enzymes with this activity are known as α‐transaminases (α‐TAs). In contrast, there are also enzymes classified as ω‐transaminases (ω‐TAs) which do not require the carboxylic acid functionality to be adjacent to the amine group that undergoes transfer, or which can even accept amines and carbonyl substrates which do not contain a carboxylic acid at all. ω‐TAs are considerably more appealing and useful in synthetic applications because of their broader substrate scope. 82 TAs have been classified into six subgroups according to the types of substrates accepted and the sequence/structure homologies: classes I–II include most l‐α‐TAs (such as l‐alanine TAs and l‐aspartate TAs), class III includes ω‐TAs, class IV includes d‐α‐TAs and branched chain α‐TAs, class V includes l‐serine TAs and class VI sugar TAs. 83 Transaminases rely on the presence of a pyridoxal 5′‐phosphate (PLP) cofactor (Scheme 15), which is covalently bound to a lysine residue in the active site in the resting state of the enzyme. The typical mechanism is based on a specular sequence of two half‐reactions. Firstly, the amine donor forms an imine with enzyme‐bound PLP, which undergoes de‐protonation and re‐protonation in the active site, forming a different imine, with the C=N bond shifted to the opposite side of the nitrogen atom. Hydrolysis of the latter imine affords the amine version of the cofactor (pyridoxamine 5′‐phosphate, PMP) and the carbonyl compound corresponding to the amine donor, which is released from the enzyme. The second half of the reaction is the mirror image of the first half, where the carbonyl acceptor binds to PMP to form an imine and all steps are reversed to yield the corresponding amine and PLP is regenerated. Kinetic studies demonstrated that PMP is easily lost from the intermediate state, leading often to insoluble and inactive high‐order apoprotein aggregates through intermolecular electrostatic interactions, but the supplementation of extra PLP in solution effectively circumvents the problem. 81 , 84 , 85
SCHEME 15.

Conversion of primary amines to aldehydes mediated by transaminases (EC 2.6.1.x) and the chemical structure of the cofactor pyridoxal 5′‐phosphate.
The availability of a huge range of well‐characterized and studied wild‐type TAs, as well as broad panels of engineered variants with broadened substrate scope, make those enzymes one of the most efficient and versatile groups of biocatalysts for the synthesis of primary and secondary (often enantiomerically enriched) amines from carbonyls. In spite of the extremely vast literature regarding the use of TAs from laboratory scale to industrial production in this direction, only a small number of examples are reported for the opposite reaction, where the starting material is a terminal primary amine to be converted to an aldehyde.
The formation of aldehydes from a range of hydroxylated arylethylamines has been investigated as a means of supplying such molecules as starting materials or intermediates for multi‐step enzymatic cascades for the synthesis of complex chiral alkaloids, such as benzylisoquinolines and protoberberines. 86 , 87 Recently, the applicability of TAs for the production of small aromatic and aryl‐substituted aliphatic aldehydes of interest for the flavour and fragrance industry, starting from the corresponding amines, has been demonstrated in excellent yield (>80%) and purity (>99%). The process was optimized and scaled up using an immobilized TA from the halo‐adapted bacterium Halomonas elongata, which was used in a biphasic system under continuous flow with co‐immobilized PLP cofactor. 88 The same process was also exploited in the production of phenylacetaldehyde and other substituted arylacetaldehydes starting from the corresponding arylethylamines in 90%–95% yield. 89
Transaminase‐mediated degradation of amino acids is regarded as a key step in the development of flavour in fermented foods, such as cheese and dairy products. Keto acids produced by transamination of amino acids are subsequently degraded to aldehydes or shorter‐chain carboxylic acids and metabolized further. 90 , 91 While the addition of isolated TAs is not generally a viable strategy in the food industry, for reasons related to cost and regulatory aspects, the selection of appropriate microorganisms for the process, with their own specific fingerprint of enzymatic activities including TAs, is instrumental to the quality and flavour profile of fermented food.
2.4.2. Monoamine oxidase
The irreversible oxidation–deamination of primary amines to the corresponding aldehydes, at the expense of the reduction of molecular oxygen to hydrogen peroxide, is catalysed by monoamine oxidases (MAOs, EC 1.4.3.4, Scheme 16). 92 MAOs are common in mammals, including humans, where they play a key role in the catabolism of food‐derived monoamines (mainly the isoform MAO‐A) and in the metabolism and regulation of monoamine neurotransmitters such as dopamine and norepinephrine, modulating their concentrations in the brain and peripheral tissues (mainly the isoform MAO‐B). 93 MAO was discovered in 1928 for its ability to oxidise tyramine to p‐hydroxyphenylacetaldehyde, and thus originally named tyramine oxidase. 94
SCHEME 16.

Conversion of primary amines to imines mediated by monoamine oxidase (MAO), followed by spontaneous deamination.
Besides mammalian enzymes, several MAOs have been identified and studied in bacteria and fungi, particularly the homologues present in Aspergillus niger (MAO‐N) 95 and in Micrococcus luteus. 96 MAOs are flavoproteins, with a covalently bound essential FAD prosthetic group. 97 Due to the richness of the reactivity of the flavin functionality, several mechanisms have been proposed for the oxidative–deamination (single‐electron transfer, step‐wise hydride transfer, two‐step hydride transfer and a polar nucleophilic mechanism). 98 , 99 Structurally, MAOs are often characterized by a hydrophobic active site at the bottom of a narrow substrate channel, sometimes with bulky side chains acting as gating residues to control substrate selection and recognition.
The applications of MAO activity to convert primary amines to aldehydes dates back to many decades ago, using partially purified enzyme 100 or whole cells of A. niger overproducing the protein, 101 for the conversion of dopamine and other biogenic amines. The substrate scope has also been investigated and expanded to the oxidative deamination of a broad range of simple and complex amines. 92
One of the most powerful drives to the development of engineered MAO variants has been the deracemization of cyclic secondary amines, exploiting an enantioselective MAO for the oxidation and a non‐selective reducing agent. 102 , 103 Several of the engineered MAOs thus obtained also showed improved activity against primary and aromatic amines. 104
MAOs can also be employed for the dehydrogenative synthesis of pyridines and pyrroles. 105 Lastly, beyond the area of synthetic preparative applications, MAO activity against small benzylamines and arylethylamines is useful for specific analytical assays with aldehydes as intermediates. 106
2.4.3. Diamine oxidase
Diamine oxidases (DAO, EC 1.4.3.22) are isofunctional to MAOs and catalyse oxidative deamination of primary amines at the expense of molecular oxygen, which is reduced to hydrogen peroxide (Scheme 17). DAOs show a very high substrate specificity for linear diamines and polyamines such as putrescine, cadaverine and spermidine, although they were formerly known as histaminases due to their biological role in the degradation of histamine to imidazole acetaldehyde.
SCHEME 17.

Conversion of primary amines to imines mediated by diamine oxidase (EC 1.4.3.22), followed by spontaneous deamination. The chemical structure of the cofactor topaquinone is also shown.
Diamine oxidases are homodimeric proteins containing one Cu(II) ion and one trihydroxyphenylalanine quinone (topaquinone, TPQ) molecule in each active site (Scheme 17). The latter is derived from the post‐translational modification of a tyrosine residue in the presence of oxygen, in a spontaneous oxidation reaction mediated by the Cu(II) ion. 107 The enzymatic reaction involves first the addition of the amine to TPQ followed by dehydration to afford a quinone‐imine intermediate, which undergoes isomerisation and hydrolysis in a similar fashion to the PLP‐dependent transamination mechanism. In the second half‐reaction, the resulting TPQ‐amine undergoes oxidative deamination to regenerate TPQ at the expense of molecular oxygen in the presence of Cu(II). 108
Human DAO, first produced recombinantly in 2002 109 has been studied extensively due to its implication in the degradation of histamine linked to food‐related immunological response. 110 Inhibition studies suggested that an aspartate residue, conserved in all DAOs but not present in other amine oxidase families, is responsible for the specificity towards diamines because it interacts with the second amine group of the most active substrates.
DAOs catalyse the oxidative deamination of many substrates, such as N‐alkylputrescines, 111 triazolyl alkyl amines, 112 p‐chlorophenylethylamine and β‐substituted ethylenediamines. 113 For instance, a DAO from Phialemonium sp. AIU 274 was screened against a panel of amine substrates, showing good conversion of long‐chain alkylamines and aliphatic amino alcohols. 114 Recently, the DAO isolated from chickpea shoots (Lathyrus cicera) gave very high to quantitative conversion with a broad range of amine substrates, and, in order to improve the scalability of the biotransformation, it was purified using a chromatography‐free protocol developed specifically for this enzyme. 115 The same DAO was also employed in a cascade system to convert the resulting aldehydes into benzylisoquinoline alkaloids. 113
A relevant area of application of DAO in the food industry is the development of electrochemical biosensors to quantify biogenic amines, linked to food safety issues (spoilage and bacterial contamination) and to potential health problems (allergic reactions or asthma). As representative examples, DAO from Lathyrus sativus was applied for electrochemical sensing of biogenic amines in wine and beer samples, 116 and a DAO from Arthrobacter crystallopoietes has been engineered to improve its already high specificity for histamine to be integrated in a biosensor to quantify histamine in fish samples rapidly and inexpensively. 117
2.5. From alkenes
Alkenes are molecules containing at least one C=C double bond and they can be categorised into terminal and internal double bonds. Alkene reagents are essential bulk chemicals and are among the most important raw materials in a plethora of reactions and synthesis processes. Nowadays, most alkenes are produced from petrol and natural gas by processes such as hydrocarbon cracking 118 , 119 or from short olefins via the Ziegler Natta 120 process or the Shell Higher Olefin Process (SHOP). 121 , 122 These alkenes can be used, in turn, to produce aldehydes and ketones via an oxidative cleavage. The most relevant alkene‐cleaving enzymes are described in the following sections. Further notable enzymes that are able to catalyse C=C double bond cleavage are: chloroperoxidases 123 (CPO), horseradish peroxidases (HRP), myeloperoxidase (MPO) 124 and lignin peroxidases (LiP). It should be noted that alkene cleavage is a side activity of those enzymes.
2.5.1. Non‐heme iron or Mn‐dependent alkene‐cleaving enzymes
Alkene‐cleaving enzymes can be further divided into enzymes using iron as cofactor and non‐iron metal‐dependent alkene cleaving enzymes with, for example, manganese in their active centre as cofactor. The metal ion is chelated in the catalytic centre by amino acid side chains like, for example, four histidines in the case of most carotenoid‐cleaving oxygenases. 125 Distinct alkene‐cleaving enzymes are described in the following.
Carotenoid cleavage oxygenase
Carotenoid cleavage oxygenases (CCOs) or carotenoid cleavage dioxygenase (CCD, both EC 1.13.11.51/67/70/71/82) have their iron cofactor chelated by four histidines in the active site. 125 A wide substrate scope is reported for this family, spanning from relatively small molecules such as, for example, isoeugenol to large and bulky substrates like their name‐giving natural substrates (carotenoids). As shown in Scheme 18, these enzymes generate dioxetanes as intermediates that decay and give rise to the corresponding aldehydes. Enzymes using this mechanism are considered dioxygenases, since two atoms of oxygen are integrated into the product. 126 On the other hand, there are CCOs forming an epoxide as an intermediate that can be classified as monooxygenases. 127
SCHEME 18.

Alkene cleavage of β‐apo‐8′‐carotenal to β‐ionone with a carotenoid cleavage oxygenase/carotenoid cleavage dioxygenase.
One of the best‐known and most important reactions catalysed by CCOs is the synthesis of vanillin as an essential fragrance and flavour molecule. 128 , 129 , 130 This is usually achieved with a two‐step reaction cascade combined with another enzyme to provide a suitable substrate for the alkene cleavage, such as isoeugenol or 4‐vinylguaiacol, starting from lignin‐derived hydrolytic aromatics. 128 , 131 , 132 , 133 , 134 Other interesting products that can be obtained by the direct oxidative cleavage of carotenoids are β‐ionone and dihydro‐β‐ionone, both molecules with a flowery odour. β‐Ionone can be synthesised directly in a one‐step reaction. Dihydro‐β‐ionone is accessible through a one‐pot reaction with a CCO and an additional reductase to reduce the double bond of β‐ionone. 135 Here, approximately 13.3 mg L−1 product was obtained in a one‐pot biosystem, corresponding to over 85% conversion. 135
Ionones are considered essential fragrance components of scent of tea, grapes, roses, tobacco and wine. 136 , 137 CCOs like AtCCD1 from A. thaliana have already attracted interest in biocatalysis for the production of those and other aroma molecules. 123 Reaction velocity of AtCCD1 can be increased up to 3.8‐fold using Triton X‐100 as a surfactant to increase substrate solubility 138 and formation of over 90% yield ionone and perfect regioselectivity was reported in reactions using crude cell lysate. 139 Furthermore, using lysates with the dioxygenases NOV1 and NOV2, complete conversion of 1 mM of resveratrol and piceatannol were achieved after 20 and 60 min, respectively. 127
Isoeugenol monooxygenase
Isoeugenol monooxygenases (IEMs, EC 1.13.11.88) are often compared to CCOs 133 since they also chelate an Fe2+ between four histidine residues and have generally a similar structure to carotenoid oxygenase. IEMs even possess the same monooxygenase mechanism, as some CCOs first form an epoxide that hydrolyses to the diol before releasing the aldehyde. 128 It perhaps could be considered to group IEMs to CCOs.
Studies using IEM in combination with a sol–gel chitosan membrane report production of up to 4.5 g L−1 (75% conversion) vanillin from 6 g L−1 isoeugenol. 140
Lutein cleavage dioxygenase
Lutein cleavage dioxygenase (LCD) is also similar to CCOs where the catalytic site is comprised of an iron ion, coordinated by four histidines. Like IEM, LCDs could also be considered as a subgroup of CCOs or more precisely CCDs.
The substrate scope is related to carotenoids, as mainly the name‐giving lutein is converted to 3‐hydroxy‐β‐ionone, a fruity, violet‐like compound. 141 This is achieved through the cleavage of the double bond between C‐9 and C‐10. 142 For example with EhLCD from Enterobacter hormaechei YT‐3637, 2 mg L−1 (87% conversion) product was obtained within 60 min with 1.5 U mL−1 of purified enzyme (Scheme 19). 141 8‐Methyl‐β‐ionone is another reported product of the degradation of lutein by E. hormaechei A20. 143
SCHEME 19.

Alkene cleavage of lutein to give 3‐hydroxy‐β‐ionone.
Cupins
Cupins are a vast family of enzymes containing various different metals as cofactor such as iron, copper, zinc, cobalt, nickel or manganese and are able to catalyse 50–100 different biochemical reactions, for example, isomerisation reactions, hydrolysis or oxygenations. Further exploration and explanation of the diversity of cupins can be found in previous works. 144 , 145 Cupins are short enzymes, consisting of 100–150 amino acids on average and four histidines to chelate the metal cofactor. The manganese‐dependent cupin TM1459 from Thermotoga maritima is capable to cleave isosafrole and piperine to yield piperonal. 146 TM1459 is reported to achieve 76% conversion of 4‐chloro‐α‐methylstyrene in biphasic systems using organic hydroperoxide as oxidant. Various other similar styrene derivatives are also reported as accepted substrates. 147
Protease A homologues
The Protease A homologues are a scarcely investigated group of enzymes for C=C cleavage, though members such as AlkCE show promising properties in this field. AlkCE was originally found in the white‐rot basidiomycete Trametes hirsuta in 2006. It has a Mn3+ centre ion chelated between two aspartic acid residues and one threonine. Unlike most other enzymes, the metal is bound exclusively to oxygen atoms, not to a nitrogen ligand as found in the side chains of histidine. 48 Alkene‐cleaving reactions can be performed with crude extract and molecular oxygen serves as the sole oxidant. Even freeze‐dried cells exhibit activity on substrates such as t‐anethole. Substrates suitable for transformation are 1,2‐dihydronaphthalene, indene, isosafrole as well as several styrene derivatives, and lead to several interesting for flavour and fragrance products such as piperonal or anisaldehyde. 148
Catechol oxygenase
A group of enzymes capable of ring‐opening reactions are catechol oxygenases and they serve as key catalysts in the oxidative degradation of aromatic compounds. These non‐heme iron enzymes use substituted aromatic rings such as catechol as substrate and can be divided into two different classes: intradiol dioxygenases, like catechol 1,2‐dioxygenase (pyrocatechase) and extradiol dioxygenases, such as catechol 2,3‐dioxygenase (also metapyrocatechase) (EC 1.13.11.2). The latter group cleaves the neighbouring double bond of the hydroxy bonds which forms 2‐hydroxymuconaldehyde 123 , 139 (Scheme 20) or other derivatives depending on substitutions of the ring. Catechol 2,3‐dioxygenases are Fe(II)‐dependent enzymes with 285 amino acids 149 that use two histidines and one glutamic acid for coordination of the Fe ion. 139 Mechanistically, the two oxygens of the substrate are coordinated to the Fe(II), before the oxygen is most likely added by an electrophilic attack, forming a Fe(II)‐semiquinone complex. Then, this leads to a Fe(II) proximal hydroperoxide over a diradical reaction producing an epoxide as an intermediate that undergoes a Criegee rearrangement followed by subsequent hydrolysis to yield the product. 150 , 151 , 152 Using a combination of catechol 2,3‐dioxygenase (cell‐free extract or crude extract) and a bisulfite nucleophilic addition for stabilization of the 2‐hydroxymuconic semialdehyde product, around 75% conversion is achieved starting from 6 g L−1 catechol (Scheme 20). 153
SCHEME 20.

Oxidative ring‐cleavage of catechol to 2‐hydroxymuconaldehyde.
Glyoxal‐forming diketone‐cleaving enzyme
Diketone‐cleaving enzyme (Dke1; EC 1.13.11.50) is a 153 amino acid long enzyme that is capable to cleave 2,4‐pentanedione (acetylacetone) to acetate and methylglyoxal with the consumption of one equivalent of molecular oxygen. 154 , 155 The enzyme is very specific, since the β‐diketone has to be present in its enol for the addition of oxygen as depicted in Scheme 21. Polar aldehydes such as methylglyoxal tend to be hydrolysed in aqueous solution. Examples of substrates that can be consumed are 2,4‐octanedione, 2‐acetylcyclohexanone and 3,5‐heptanedione. The presence of two ketones in beta position are essential, since their exchange to other moieties such as alcohols or carboxylic acids results in no conversion. 154 The Fe2+ cation in Dke1 serving as the catalytic centre, is bound by a triad consisting of histidines 156 , 157 , 158 instead of the 2‐His‐1‐carboxylate facial triad commonly found in other mononuclear non‐heme iron enzymes.
SCHEME 21.

Cleaving reaction of diketone‐cleaving enzyme.
2.5.2. Heme‐dependent alkene‐cleaving enzymes
Heme‐dependent alkene cleaving enzymes possess ferric protoporphyrin IX or protoheme and are capable of using molecular oxygen or hydrogen peroxide as the oxidant. Additionally to the heme complex, the iron can be coordinated by a histidine or serine from below the plane of the porphyrin. Two examples for oxygen‐requiring enzymes that are able to catalyse oxidative ring cleavage are tryptophan 2,3‐dioxygenase (TDO) and indoleamine 2,3‐dioxygenase (IDO). Their substrates scope is wide and varies from cyclic systems to polymers such as rubber.
Tryptophan 2,3‐dioxygenase and Indoleamine 2,3‐dioxygenase
Tryptophan 2,3‐dioxygenases 159 (TDO, EC 1.13.11.11) and indoleamine 2,3‐dioxygenases 160 , 161 (IDO, EC 1.13.11.17) are classified as heme‐dependent alkene‐cleaving enzymes and use molecular oxygen for a ring‐opening reaction via the cleavage of a C=C double bond. Both serve in tryptophan catabolism through the kynurenine pathway and are also produced in cells in response to inflammation. They are especially highly expressed in tumour cells to reduce detection by the immune system. 162 , 163 TDOs are 35–45 kDa proteins and are tetramers, while IDOs are monomeric enzymes with an approximate mass of 45 kDa. Though isofunctional, they possess a low sequence identity with 16% between human TDO and human IDO1. 161 An important difference in the sequence is the presence of a histidine that interacts with the NH group of the indole ring of l‐tryptophan, aiding in substrate binding in TDOs, while IDOs lack this residue and have a serine in this place. 164 Heme iron serves to activate the oxygen to allow the rection to occur. Nowadays, a radical addition mechanism is postulated as the first step, 161 but alternative mechanisms for the initial oxygen addition have also been suggested. 165 Over a peroxo transition state, an epoxide intermediate and a ferryl heme [Fe(IV)] are formed. The presence of a ferryl heme was confirmed in IDOs. 166 , 167 , 168 Protonation opens the epoxide ring and triggers the addition of oxygen chelating the iron to C2, leading to the formation of N‐formyl‐l‐kynurenine (Scheme 22). 161 Both TDO and IDO have already been shown to be able to completely convert several substrates such as tryptophan at low concentrations. 123
SCHEME 22.

Oxidative ring‐cleaving reaction catalysed by tryptophan 2,3‐dioxygenase.
Rubber oxygenases RoxA and RoxB
Rubber oxygenases (Rox, EC 1.13.11.x) are extracellular c‐type diheme dioxygenases 169 found in rubber‐degrading bacteria that can cleave the C=C double bond of poly(cis‐1,4‐isoprene). While the cleavage product of RoxA is mainly 2‐oxo‐4,8‐dimethyltrideca‐4,8‐diene‐1‐al (ODTD), a C15 oligoisoprenoid, 170 RoxB possesses a distinctive product spectrum of C20, C25, C30 and higher oligo‐isoprenoids. 171 RoxA consists of 678 amino acids and features two binding sites for covalent attachment of heme. 137 , 172 The reaction is performed without additional cofactors and uses O2 as oxidant. First, the double bond attacks the oxygen that deprotonates the substrate, forming temporarily a hydroperoxide and shifting temporarily the double bond between C1 and C2. Then, spontaneously, a 1,2‐dioxetane is formed as an intermediate that releases the corresponding aldehydes upon rearrangement. 169
2.6. From cyanohydrins
α‐Hydroxynitriles (alternative name: cyanohydrins) occur in plants in the form of cyanogenic glycosides for the purpose of defence against herbivores. Approximately 25 cyanogenic glycosides are known in plants, whereby the cyanohydrin moiety is attached to a mono‐ or disaccharide. 173 Also, certain insects harbour such glycosides. 174
2.6.1. Hydroxynitrile lyase
Hydroxynitrile lyases (HNL; EC 4.2.1.X) catalyse the cleavage of cyanohydrins, which are released from cyanogenic glycosides by glycosidases in the first step. 175 In this manner, simple aldehydes like benzaldehyde and isobutyraldehyde are accessible from renewable sources. For synthetic purposes, the reverse reaction is evidently much more relevant, as it affords a chiral, bi‐functional molecule, which can undergo a great variety of follow‐up reactions. 176 , 177 In this sense, HNLs have considerably more impact in aldehyde transformation than they do in aldehyde synthesis.
2.7. From hydroperoxides and epoxides
Hydroperoxides and epoxides are formed in nature by reaction of unsaturated compounds such as fatty acids with reactive oxygen species in an untargeted manner or with metal‐dependent redox enzymes in a targeted fashion, and they play a major role in redox‐dependent signalling. Reactive epoxides emerge in the degradation of aromatic compounds. 178 Plants, for example, utilize polyunsaturated fatty acids as substrates for lipoxygenases (LOX) 179 , 180 or α‐dioxygenases (αDOX), which produce fatty acid hydroperoxides as precursor molecules for aldehydes that are formed in a second step and often give rise to the typical odour of the respective plant.
2.7.1. α‐Dioxygenase
α‐Dioxygenases (α‐DOXs, EC 1.13.11.92) oxidise medium‐chain fatty acids to 2‐hydroperoxy fatty acids (Scheme 23). Spontaneous decarboxylation and dehydration give rise to carbon chain‐shortened fatty aldehydes and one equivalent each of H2O and CO2. 181 The substrate scope of α‐DOXs differs depending on their origin. Plant enzymes accept longer chain fatty acids (>C14), whereas cyanobacterial α‐DOXs studied up to date show preference for shorter chain fatty acids (C10–C16). 182 , 183 , 184 From the mechanistic viewpoint, α‐DOXs are heme‐dependent and carry an essential tyrosine residue close to the heme. Similar as in CRO‐AOx, a tyrosyl radical is formed. The α‐hydrogen of the fatty acid substrate is transferred to the tyrosyl radical, followed by O2 radical trapping and non‐enzymatic decarboxylation of the formed α‐hydroperoxy fatty acid. 185 For more comprehensive information about α‐DOXs and its applications for fatty aldehyde synthesis, the reader is referred to a very recent review. 185 Notably, some applications target the formation of a mixture of aldehydes. In a recent example, an α‐DOX was coupled to a fatty aldehyde dehydrogenase, which lead to the iterative decarboxylation of long‐chain fatty acids to mixtures of long‐ and medium‐chain aldehydes. 186
SCHEME 23.

α‐Dioxygenase‐mediated fatty acid hydroperoxidation followed by spontaneous decarboxylation to yield carbon chain shortened fatty acid.
2.7.2. Hydroperoxide lyase
Hydroperoxide lyases (HPLs, EC 4.1.2.x) are known to catalyse the cleavage of hydroperoxides of polyunsaturated fatty acids to a short‐chain aldehyde and an ω‐oxo acid respectively. HPLs are a particular subclass of heme proteins of the cytochrome P450 family (CYP74). They do not function as monooxygenases, but isomerize the hydroperoxide into a short‐lived hemi‐acetal, 187 which finally decomposes to two aldehyde species. In plants, HPL is acting in concert with lipoxygenases of matching product selectivity, depending on the (poly)unsaturated fatty acid that is available as a substrate. 180 In essence, the overall reaction sequence resembles that of alkene‐cleaving enzymes (Section 2.5). Typical substrates of LOX/HPL cascades are linoleic acid, α‐linolenic acid, oleic acid and arachidonic acid. In context of arachidonic acid metabolism, cyclooxygenase (COX) is worth to be mentioned as a producer of hydroperoxyl arachidonic acid. 188 Depending on the regio‐ and stereoselectivities of the enzymes, green leaf volatiles (e.g. C6, C9) are formed (Scheme 24). On analytical level, aldehyde production with recombinant HPL in E. coli appears to be promising. 189 When utilizing the LOX/HPL system recombinantly in microbial hosts, the aldehydes are often (intentionally) metabolized further, hence, mostly alcohol formation has been published in synthetic applications. In a recent example, tomato HPL was expressed in E. coli and used for (Z)‐3‐hexenol synthesis. 190 Impressive space–time yields of >8 g L−1 h−1 were published for the same product in non‐optimized batch reactions using bacterial cell suspensions with engineered guava HPL in combination with a commercial ADH for aldehyde reduction. 191
SCHEME 24.

Hydroperoxide lyase mediated isomerisation of fatty acid hydroperoxides to short‐lived hemi‐acetals, followed by spontaneous hydrolysis into two aldehyde species.
2.7.3. Styrene oxide isomerase
Styrene oxide isomerase (SOI, EC 5.3.99.7) catalyses the isomerisation of styrene oxide to phenylacetaldehyde (Scheme 25). 192 In case of trans‐1‐phenylpropylene oxide, the terminal methyl group undergoes a 1,2 shift and enantiomerically pure 2‐phenylpropanal is formed. The stereochemistry of the product depends on the absolute configuration of the epoxide. 193 The substrate scope of SOIs appears to be quite restricted to epoxides adjacent to a phenyl (styrene and a handful of substituted styrene derivatives). 194 SOIs are small (20 kDa) membrane‐bound proteins, which might be the reason that a protein structure is not available yet, and also the reaction mechanism remains largely elusive. 194 Nevertheless, the enzyme finds application in various cascade reactions, 194 the production of phenylacetaldehyde (vide infra) 195 and 2‐phenylethanol production. 196
SCHEME 25.

Name‐giving reaction of styrene oxide isomerases. The reaction resembles the Meinwald rearrangement.
2.7.4. Benzoyl‐CoA dihydrodiol lyase (BDL)
Benzoyl‐CoA dihydrodiol lyase (BoxC; EC 4.1.2.44) catalyses the hydration of the non‐aromatic 2,3‐epoxide of benzoyl‐CoA that is a metabolite in the anaerobic degradation of benzoate. Two molecules of H2O are required to generate 3,4‐dehydroadipyl‐CoA semialdehyde and one equivalent of formate (Scheme 26). 197 Relevance of this reaction in a synthetic context has not been demonstrated yet.
SCHEME 26.

Aldehyde‐forming step in anaerobic benzoate degradation.
2.8. From keto acids
Keto acids are essential organic compounds that contain a keto group and a carboxyl group. 198 They have different natural functions, and their stability depends on the position of the carbonyl group. Currently, the majority of α‐keto acids are produced by chemical synthesis, but their biotechnological production, mainly biotransformation, and de novo synthesis from inexpensive renewable carbohydrates are promising alternatives. 199 , 200 , 201 Therefore, α‐keto acids are used as a source of aldehyde precursors in non‐oxidative decarboxylation in which a corresponding aldehyde and carbon dioxide are formed. 202
2.8.1. Pyruvate decarboxylase
Pyruvate decarboxylase (PDC; EC 4.1.1.1) is an extensively examined class of decarboxylases, involved in the ethanol production pathway. PDCs are prevalent in yeast and plants and are seldom present in prokaryotes. Their primary function is to decarboxylate pyruvate to acetaldehyde and carbon dioxide (Scheme 27). 203 Moreover, PDCs are able to utilize other aliphatic as well as aromatic 2‐oxoacids. The catalytic activity of PDC depends on the presence of the cofactor thiamine diphosphate (ThDP), which is bound mainly via a Mg2+ ion to the protein moiety at the interface of two subunits. 204
SCHEME 27.

Decarboxylase‐mediated disintegration of α‐oxoacids to aldehyde and carbon dioxide.
Since acetaldehyde is an inhibitory and toxic by‐product for microorganisms, intracellular concentrations are typically low. However, Lactobacillus lactis was engineered to accumulate high yields of acetaldehyde (0.418 g L−1) produced by decarboxylation of pyruvate by PDC originating from Z. mobilis. 205 Not only L. lactis but also E. coli was engineered to overproduce acetaldehyde starting from glucose. Acetaldehyde concentration was doubled to 0.725 g L−1 by knocking out the competing metabolic pathways. 206 Acetaldehyde is not only a valuable compound used as a substrate for further reactions, but also an important aroma in yoghurts and other dairy products. 205
2.8.2. Other α‐keto acid decarboxylases
Benzoylformate decarboxylase (BFD, EC 4.1.1.7) was originally isolated from Pseudomonas putida but was also purified from Acinetobacter species. These enzymes are also ThDP‐Mg2+‐dependent with certain similarities to ZmPDC. BFD plays an inevitable role in the mandelate pathway, where it catalyses the removal of CO2 from benzoyl formate to generate benzaldehyde. 207 Benzaldehyde is one of the most frequently applied aromas in the food industry. It is interesting due to its bitter almond odour with cherries, malt or roasted pepper tones. 208 However, BFD is also active towards certain aliphatic α‐keto acids, such as pyruvic acid, 2‐oxobutyric acid, 4‐methyl‐2‐oxopentanoic acid that are converted at lower rates. The products of these substrates are acetaldehyde, propanal and 3‐methylbutanal, respectively. 209
Branched‐chain keto acid decarboxylases (KdcAs, EC 4.1.1.72) have been found in different L. lactis strains. The substrate scope of KdcAs is wider in comparison to PDCs, but both enzymes are ThDP‐dependent. 210 KdcA is involved in the Ehrlich pathway, which eventually leads to the formation of alcohols from branched‐chain amino acids. 211 In addition to the main pathway, amino acid catabolism may result in hydroxy acids, esters and likewise in high‐impact aldehyde flavours formed from valine, isoleucine, phenylalanine and from other amino acids. 212 , 213 Other aldehydes accessible through KdcA‐mediated decarboxylation worth mentioning are methional (potato flavour present in crisps) or phenylacetaldehyde with a sweet, honey‐like odour resembling hyacinths and frequently added to perfumes. 1
Ketoisovalerate decarboxylase (KID) is the second decarboxylase isolated from L. lactis with almost 90% amino acid sequence identity with KdcA. 212 Therefore, KID also participates in the metabolism of amino acids, where it is involved in the production of 3‐methylbutanal starting from leucine. 214 3‐Methylbutanal is the aroma with caramel cocoa flavour, which contributes to the nutty flavour of Cheddar cheese. 214 , 215 Various strains of L. lactis were isolated from fermented products and tested for the activity of KID. The highest yields showed strain 408 that produced 12.1 ± 0.6 mg L−1 of 3‐methylbutanal. 216 , 217 Alternatively, recombinant KID was produced together with leucine dehydrogenase for in vitro cascade reactions from valine to 3‐methylbutanal. In continuous mode, the best catalyst formulation gave 340 mg L−1 product concentration. 218
2.9. From other precursors
2.9.1. Chloroacrylic acid dehalogenase
Chloroacrylic acid dehalogenases (CaaDs) are tautomerase enzymes that catalyse the hydrolytic dehalogenation of cis‐3‐chloroacrylic acid to afford malonate semialdehyde (Scheme 28). 219
SCHEME 28.

Tautomerase catalysed dehalogenation.
3. FLAVOUR AND FRAGRANCE ALDEHYDES TARGETED BY ENZYMATIC REACTIONS
3.1. Aromatic aldehydes
Vanillin is the spotlight aldehyde and various routes to this compound have been explored in academia and industry, up to production level. 220 Biocatalytic vanillin production from ferulic acid was, for example, commercialized more than 20 years ago using wild‐type organisms. 5 The most recent report shows vanillin formation from isoeugenol with a recombinant isoeugenol monooxygenase (Jin1 from Pseudomonas nitroreducens) at 0.3 M substrate load in a space–time yield of 4.8 g L−1 h−1. 221
Piperonal (heliotropin) is an aromatic aldehyde that is structurally related to vanillin. It is a floral‐type odour with creamy cherry and vanilla smell. 136 This aldehyde was enzymatically prepared from piperonylic acid with CAR as the key enzyme in an E. coli‐based living cell catalyst with as space–time yield of 1.5 g L−1 h−1. 222 Starting from the respective alcohol, space–time yield of 9.5 g L−1 h−1 was achieved using AAOx. 63 In both reports, piperonal was isolated in high yield and >99% purity via simple crystallization from n‐hexane extracts.
Further benzaldehyde derivatives are accessible through the action of monooxygenases via oxidation of their primary products, the respective alcohols. The non‐heme iron monooxygenase XylM, for example, was used for the production of 3,4‐dimethylbenzaldehyde from pseudocumene in a space–time yield of 1.6 g L−1 h−1 after systematic optimization. 223 One of the key tasks here was the suppression of further oxidation to the respective acid and this was—amongst others—achieved by in situ removal of the aldehyde, the most frequently used strategy in bioaldehyde production.
3.2. Aliphatic aldehydes
Acetaldehyde is a short‐chain aldehyde with a pungent, ethereal, fresh and fruity odour. This aldehyde inhibits microbial growth at millimolar concentrations and therefore is a rather challenging fermentation product. Zymomonas mobilis shake flask cultivation gave up to 1.6 g L−1 acetaldehyde with its native PDC. 224 The AOx of Pichia pastoris was used in biotransformation mode with resting cells and delivered up to 70 g L−1 of acetaldehyde when ISPR with TRIS was used. 225
The odour perception of isobutyraldehyde is described as fresh, aldehydic, floral, pungent and green. Rodriguez and Atsumi engineered E. coli towards isobutyraldehyde production using a L. lactis keto acid decarboxylase (KID) for the aldehyde‐forming step from 2‐ketoisovalerate. To suppress further reduction to isobutanol, altogether 15 genes were knocked out. The final strain reached a product titre of 35 g L−1 after 5 days (space–time yield of 0.29 g L−1 h−1), using gas stripping as ISPR strategy. 226
Hexenal and hexanal are referred to as typical ‘green leaf volatiles’ (GLVs) because their smell is associated to freshly cut grass, green fruits and vegetables. Traditionally, plant‐derived enzyme preparations were used for GLV production from polyunsaturated fatty acids: Akacha and Gargouri, for example, used olive leaf‐derived HPL to produce hexenal in a liquid/gas reactor and achieved 50% hexenals with respect to the added substrate 13‐(S)‐hydroperoxylinolenic acid, which corresponded 0.36 g kg−1 of reaction medium. 227 Recombinant, engineered guava HPL was published by Firmenich SA for green leaf alcohol synthesis (8 g L−1), 191 and with further optimization is likely an applicable biocatalyst for production scale of both, GLV aldehydes and alcohols.
The odour of octanal is described as waxy, fatty, citrus, orange, peely, green, herbal, aldehydic and fresh. High aldehyde yield was achieved for the conversion of octanoic acid to octanal using a recombinant CAR from Mycobacterium marinum in engineered E. coli cells in the presence of n‐heptane as ISPR solvent. 228 When opting for this strategy, ISPR has the double function of protecting the aldehyde product from cell‐mediated follow‐up reactions and protecting the cells from the cytotoxic product compound. Biocompatible solvents like hexadecane hold even more promise, as they are able to sustain cell viability and therefore ATP supply even better.
3.3. Aryl‐aliphatic aldehydes
Phenylacetaldehyde is a green‐type odour with a sweet and floral smell, also reminding of honey and cocoa/chocolate. This aldehyde can be produced from styrene with SOI. High productivity of this isomerase enzyme was showcased by attaching a small protein fusion tag to enhance its expression, resulting in the highest reported phenylacetaldehyde production of (405 g L−1) from styrene oxide so far. 195
The spicy odour type cinnamaldehyde is perceived as cinnamon, sweet, aldehydic, aromatic, balsamic, resinous and powdery, sometimes also honey. Cinnamaldehyde was in the spotlight of a recent study which utilised an engineered cinnamic acid producing Corynebacterium glutamicum strain. The same C. glutamicum strain was equipped with Mycobacterium phlei CAR (MpCAR). Undesired aldehyde metabolism was suppressed by deletion of four genes. Cinnamaldehyde (2.2 g L−1 h−1) was obtained after full conversion of crude cinnamic acid (1.2 g L−1) from a C. glutamicum culture supernatant. 229 Higher product titres were obtained in the laboratory by a heavily engineered E. coli strain in combination with MmCAR, however, space–time yields were considerably lower (0.049 g L−1 h−1). 230
3.4. Heteroaromatic aldehydes
Even though furfural and 5‐hydroxymethylfurfural (HMF) belong to the most abundant biomass‐derived aldehydes, there is also research on their synthesis using chemoenzymatic pathways including glucose isomerase 231 and microbial biosynthesis by introducing heterologous pathways into microorganisms. 232 Furthermore, oxidation of HMF by AOx was discussed in Section 2.3.1.
Pyrrole‐2‐carbaldehyde is a musty‐type odour found in the highest concentration not only in beer but also in teas and coffees. Synthesis from pyrrole and CO2 was reported using a one‐pot system with a UbiD‐type decarboxylase from Pseudomonas aeruginosa HudA/PA0254 in combination with CAR from Segniliparus rotundus (SroCAR) as a whole‐cell biocatalyst (Scheme 29). Low maximum product yield of 2.14 ± 0.16 mM was assigned to the instability of the target product. Except pyrrole‐2‐carboxylic acid, the SroCAR was capable of reducing furan‐ and thiophene‐2‐carboxylic acids to their corresponding aldehydes. 233 A similar two‐step approach (using a decarboxylase from Arthrobacter nicotianae and CAR from S. rugosus) was used to produce indole‐3‐carbaldehyde from indole. 234
SCHEME 29.

One‐pot production of pyrrole‐2‐carbaldehyde from pyrrole, with a carboxylation performed with a UbiD‐type decarboxylase from Pseudomonas aeruginosa (PA0254) and CAR from Segniliparus rotundus (SroCAR) as a whole‐cell biocatalyst.
4. CONCLUSIONS
Aldehydes are challenging chemicals to make, due to their high reactivity. In many cases, there is a kinetic preference for aldehyde reactions in comparison to the reactivity of their precursor molecules. This is a fact in chemical as well as in enzymatic or biocatalytic production of aldehydes. Herein, we described enzymatic routes from 11 classes of precursor molecules that may be treated with enzymes to produce aldehydes as their products. A 12th possibility was omitted deliberately, which is the formation of aldehydes by aldolases, which interconvert one aldehyde into another. While some of the presented reactions may be perceived as laboratory curiosities, others have certainly the potential to be used for synthesis (Table 1). We devoted a chapter to those reactions which delivered the best yields of molecules perhaps most relevant for the flavour and fragrance sector. In this respect, the most often selected reactions are alcohol oxidation by alcohol oxidases, acid reduction by carboxylic acid reductases, keto acid decarboxylation by decarboxylases and the combination of lipoxygenases with hydroperoxide lyases. In terms of catalyst form, the most prevalent formulations are resting cell or living cell biocatalysts. On the one hand, this is a necessity in case metabolic reactions are required, and on the other hand, it is owed to handling simplicity and enzyme stability.
TABLE 1.
An overview of biocatalytic flavour and fragrance aldehyde‐forming reactions.
| Products | Substrates | Enzymes | Catalyst form | Volumetric productivity or STY | Reference |
|---|---|---|---|---|---|
| Vanillin | Isoeugenol | IEM | Resting cells | 4.8 g L−1 h−1 | 221 |
| Piperonal | Piperonylic alcohol | AAOx | Purified enzyme | 9.5 g L−1 h−1 | 63 |
| Piperonal | Piperonylic acid | CAR | Resting cells | 1.5 g L−1 h−1 | 222 |
| 3,4‐Dimethylbenzaldehyde | Pseudocumene | AOx | Living cells | 1.6 g L−1 h−1 | 223 |
| Acetaldehyde | Pyruvate | PDC | Living cells | 1.6 g L−1 (10 h) | 224 |
| Acetaldehyde | Ethanol | AOx | Resting cells | 44 g L−1 (24 h) | 225 |
| Isobutyraldehyde | 2‐ketoisovalerate | KID | Living cells | 35 g L−1 (5 days) | 226 |
| Hexenal | 13‐Fatty Acid Hydroperoxides | HPL | Cell lysate | 8 g L−1 h−1 | 191 |
| Octanal | Octanoic acid | CAR | Living cells | 0.458 g L−1 h−1 | 228 |
| Phenylacetaldehyde | Styrene oxide | SOI | Resting cells | 405 g L−1 (2 h) | 195 |
| Cinnamaldehyde | Cinnamic acid | CAR | Living cells | 2.2 g L−1 h−1 | 229 |
| Pyrrole carbaldehyde | Pyrrole‐2‐carboxylic acid | CAR | Living cells | 0.2 g L−1 (18 h) | 233 |
CONFLICT OF INTEREST STATEMENT
The authors declare no financial or commercial conflict of interest.
ACKNOWLEDGEMENTS
This work has been financially supported by the Austrian Science Fund (FWF Project P 33687). The COMET centre acib: Next Generation Bioproduction is funded by BMK, BMDW, SFG, Standortagentur Tirol, Government of Lower Austria and Vienna Business Agency in the framework of COMET—Competence Centers for Excellent Technologies. The COMET‐funding program is managed by the Austrian Research Promotion Agency FFG.
Schober L, Dobiašová H, Jurkaš V, Parmeggiani F, Rudroff F, Winkler M. Enzymatic reactions towards aldehydes: An overview. Flavour Fragr J. 2023;38:221‐242. doi: 10.1002/ffj.3739
DATA AVAILABILITY STATEMENT
Data sharing is not applicable to this article as no new data were created or analyzed in this study.
REFERENCES
- 1. Kohlpaintner C, Schulte M, Falbe J, Lappe P, Weber J, Frey GD. Ullmann's Encyclopedia of Industrial Chemistry. Wiley‐VCH Verlag GmbH & Co. KGaA; 2013. [Google Scholar]
- 2. Zhou J, Chen Z, Wang Y. Bioaldehydes and beyond: expanding the realm of bioderived chemicals using biogenic aldehydes as platforms. Curr Opin Chem Biol. 2020;59:37‐46. [DOI] [PubMed] [Google Scholar]
- 3. Pickett J, Anderson D, Bowles D, et al. Sustainable Biofuels: Prospects and Challenges. The Royal Society; 2008. [Google Scholar]
- 4. Renders T, Van Den Bosch S, Koelewijn SF, Schutyser W, Sels BF. Lignin‐first biomass fractionation: the advent of active stabilisation strategies. Energ Environ Sci. 2017;10:1551‐1557. [Google Scholar]
- 5. Schrader J, Etschmann MMW, Sell D, Hilmer J‐M, Rabenhorst J. Applied biocatalysis for the synthesis of natural flavour compounds – current industrial processes and future prospects. Biotechnol Lett. 2004;26:463‐472. [DOI] [PubMed] [Google Scholar]
- 6. Lecourt M, Antoniotti S. Chemistry, sustainability and naturality of perfumery biotech ingredients. ChemSusChem. 2020;13:5600‐5610. [DOI] [PubMed] [Google Scholar]
- 7. Kazimírová V, Rebroš M. Production of aldehydes by biocatalysis. Int J Mol Sci. 2021;22:4949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Winiarska A, Hege D, Gemmecker Y, et al. Tungsten enzyme using hydrogen as an electron donor to reduce carboxylic acids and NAD+ . ACS Catal. 2022;12:8707‐8717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Napora‐Wijata K, Strohmeier GA, Winkler M. Biocatalytic reduction of carboxylic acids. Biotechnol J. 2014;9:822‐843. [DOI] [PubMed] [Google Scholar]
- 10. Hollmann F, Arends IWCE, Holtmann D. Enzymatic reductions for the chemist. Green Chem. 2011;13:2285. [Google Scholar]
- 11. Nissen LS, Basen M. The emerging role of aldehyde: ferredoxin oxidoreductases in microbially‐catalyzed alcohol production. J Biotechnol. 2019;306:105‐117. [DOI] [PubMed] [Google Scholar]
- 12. Ni Y, Hagedoorn P, Xu J, Arends IWCE, Hollmann F. A biocatalytic hydrogenation of carboxylic acids. Chem Commun (Camb). 2012;48:12056‐12058. [DOI] [PubMed] [Google Scholar]
- 13. Winkler M, Ling JG. Biocatalytic carboxylate reduction–recent advances and new enzymes. ChemCatChem. 2022;14:e202200441. [Google Scholar]
- 14. Tavanti M, Hosford J, Lloyd RC, Brown MJB. Recent developments and challenges for the industrial implementation of polyphosphate kinases. ChemCatChem. 2021;13:3565‐3580. [Google Scholar]
- 15. Strohmeier GA, Eiteljörg IC, Schwarz A, Winkler M. Enzymatic one‐step reduction of carboxylates to aldehydes with cell‐free regeneration of ATP and NADPH. Chem Eur J. 2019;25:6119‐6123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Jaroensuk J, Chuaboon L, Chaiyen P. Biochemical reactions for in vitro ATP production and their applications. Mol Catal. 2023;537:112937. [Google Scholar]
- 17. Tavanti M, Hosford J, Lloyd RC, Brown MJB. ATP regeneration by a single polyphosphate kinase powers multigram‐scale aldehyde synthesis in vitro . Green Chem. 2021;23:828‐837. [Google Scholar]
- 18. Schwarz A, Hecko S, Rudroff F, Kohrt JT, Howard RM, Winkler M. Cell‐free in vitro reduction of carboxylates to aldehydes: with crude enzyme preparations to a key pharmaceutical building block. Biotechnol J. 2021;16:2000315. [DOI] [PubMed] [Google Scholar]
- 19. Park J, Lee HS, Oh J, Joo JC, Yeon YJ. A highly active carboxylic acid reductase from Mycobacterium abscessus for biocatalytic reduction of vanillic acid to vanillin. Biochem Eng J. 2020;161:107683. [Google Scholar]
- 20. Hansen J, Hansen EH, Sompalli HP, Sheridan J, Heal J, Hamilton W. Compositions and methods for the biosynthesis of vanillin or vanillin β‐D‐glucoside. 2013:WO2013022881A8.
- 21. Awodi UR, Ronan JL, Masschelein J, De Los Santos ELC, Challis GL. Thioester reduction and aldehyde transamination are universal steps in actinobacterial polyketide alkaloid biosynthesis. Chem Sci. 2017;8:411‐415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Mullowney M, McClure RA, Robey MT, Kelleher NL, Thomson RJ. Natural products from thioester reductase containing biosynthetic pathways. Nat Prod Rep. 2018;35:847‐878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Yang C, Xu Y, Xu K, Tan G, Yu X. Preparation of new halogenated diphenyl pyrazine analogs in Escherichia coli by a mono‐module fungal nonribosomal peptide synthetase from Penicillium herquei . Tetrahedron Lett. 2018;59:3084‐3087. [Google Scholar]
- 24. Hai Y, Huang AM, Tang Y. Structure‐guided function discovery of an NRPS‐like glycine betaine reductase for choline biosynthesis in fungi. Proc Natl Acad Sci USA. 2019;116:10348‐10353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Gahloth D, Dunstan MS, Quaglia D, et al. Structures of carboxylic acid reductase reveal domain dynamics underlying catalysis. Nat Chem Biol. 2017;13:975‐981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Schultz JC, Mishra S, Gaither E, et al. Metabolic engineering of Rhodotorula toruloides IFO0880 improves C16 and C18 fatty alcohol production from synthetic media. Microb Cell Fact. 2022;21:26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Bang HB, Lee YH, Kim SC, Sung CK, Jeong KJ. Metabolic engineering of Escherichia coli for the production of cinnamaldehyde. Microb Cell Fact. 2016;15:16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Chen M, Pang B, Ding W, Zhao Q, Tang Z, Liu W. Investigation of 2,2′‐bipyridine biosynthesis reveals a common two‐component system for aldehydes production by carboxylate reduction. Org Lett. 2022;24:897‐902. [DOI] [PubMed] [Google Scholar]
- 29. Krishnan A, McNeil BA, Stuart DT. Biosynthesis of fatty alcohols in engi‑neered microbial cell factories: advances and limitations. Front Bioeng Biotechnol. 2020;8:610936. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Liu J, Wu S, Li Z. Recent advances in enzymatic oxidation of alcohols. Curr Opin Chem Biol. 2018;43:77‐86. [DOI] [PubMed] [Google Scholar]
- 31. Ribeaucourt D, Bissaro B, Lambert F, Lafond M, Berrin J‐G. Biocatalytic oxidation of fatty alcohols into aldehydes for the flavors and fragrances industry. Biotechnol Adv. 2022;56:107787. [DOI] [PubMed] [Google Scholar]
- 32. Pickl M, Fuchs M, Glueck SM, Faber K. The substrate tolerance of alcohol oxidases. Appl Microbiol Biotechnol. 2015;99:6617‐6642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Kroutil W, Mang H, Edegger K, Faber K. Biocatalytic oxidation of primary and secondary alcohols. Adv Synth Catal. 2004;346:125‐142. [Google Scholar]
- 34. Whittaker JW. Free radical catalysis by galactose oxidase. Chem Rev. 2003;103:2347‐2364. [DOI] [PubMed] [Google Scholar]
- 35. Parikka K, Master E, Tenkanen M. Oxidation with galactose oxidase: multifunctional enzymatic catalysis. J Mol Catal B Enzym. 2015;120:47‐59. [Google Scholar]
- 36. McKenna SM, Leimkühler S, Herter S, Turner NJ, Carnell AJ. Enzyme cascade reactions: synthesis of furandicarboxylic acid (FDCA) and carboxylic acids using oxidases in tandem. Green Chem. 2015;17:3271‐3275. [Google Scholar]
- 37. Karich A, Kleeberg SB, Ullrich R, Hofrichter M. Enzymatic preparation of 2, 5‐furandicarboxylic acid (FDCA)—a substitute of terephthalic acid—by the joined action of three fungal enzymes. Microorganisms. 2018;6:5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Cajnko MM, Novak U, Grilc M, Likozar B. Enzymatic conversion reactions of 5‐hydroxymethylfurfural (HMF) to bio‐based 2, 5‐diformylfuran (DFF) and 2, 5‐furandicarboxylic acid (FDCA) with air: mechanisms, pathways and synthesis selectivity. Biotechnol Biofuels. 2020;13:66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Birmingham WR, Turner NJ. A single enzyme oxidative “cascade” via a dual‐functional galactose oxidase. ACS Catal. 2018;8:4025‐4032. [Google Scholar]
- 40. Lippow SM, Moon TS, Basu S, et al. Engineering enzyme specificity using computational design of a defined‐sequence library. Chem Biol. 2010;17:1306‐1315. [DOI] [PubMed] [Google Scholar]
- 41. Deacon SE, Mahmoud K, Spooner RK, et al. Enhanced fructose oxidase activity in a galactose oxidase variant. ChemBioChem. 2004;5:972‐979. [DOI] [PubMed] [Google Scholar]
- 42. Rannes JB, Ioannou A, Willies SC, et al. Glycoprotein labeling using engineered variants of galactose oxidase obtained by directed evolution. J Am Chem Soc. 2011;133:8436‐8439. [DOI] [PubMed] [Google Scholar]
- 43. Escalettes F, Turner NJ. Directed evolution of galactose oxidase: generation of enantioselective secondary alcohol oxidases. ChemBioChem. 2008;9:857‐860. [DOI] [PubMed] [Google Scholar]
- 44. Herter S, McKenna SM, Frazer AR, Leimkühler S, Carnell AJ, Turner NJ. Galactose oxidase variants for the oxidation of amino alcohols in enzyme cascade synthesis. ChemCatChem. 2015;7:2313‐2317. [Google Scholar]
- 45. Huffman MA, Fryszkowska A, Alvizo O, et al. Design of an in vitro biocatalytic cascade for the manufacture of islatravir. Science. 2019;366:1255‐1259. [DOI] [PubMed] [Google Scholar]
- 46. (Tyler) Yin D, Urresti S, Lafond M, et al. Structure‐function characterization reveals new catalytic diversity in the galactose oxidase and glyoxal oxidase family. Nat Commun. 2015;6:10197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Mathieu Y, Offen WA, Forget SM, et al. Discovery of a fungal copper radical oxidase with high catalytic efficiency toward 5‐hydroxymethylfurfural and benzyl alcohols for bioprocessing. ACS Catal. 2020;10:3042‐3058. [Google Scholar]
- 48. Cleveland M, Lafond M, Xia FR, et al. Two Fusarium copper radical oxidases with high activity on aryl alcohols. Biotechnol Biofuels. 2021;14:138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Cleveland ME, Mathieu Y, Ribeaucourt D, et al. A survey of substrate specificity among auxiliary activity family 5 copper radical oxidases. Cell Mol Life Sci. 2021;78:8187‐8208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Ribeaucourt D, Saker S, Navarro D, et al. Identification of copper‐containing oxidoreductases in the secretomes of three Colletotrichum species with a focus on copper radical oxidases for the biocatalytic production of fatty aldehydes. Appl Environ Microbiol. 2021;87:e01526‐e01521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Ribeaucourt D, Bissaro B, Guallar V, et al. Comprehensive insights into the production of long chain aliphatic aldehydes using a copper‐radical alcohol oxidase as biocatalyst. Chem Eng. 2021;9:4411‐4421. [Google Scholar]
- 52. Ribeaucourt D, Höfler GT, Yemloul M, et al. Tunable production of (R)‐ or (S)‐citronellal from geraniol via a bienzymatic cascade using a copper radical alcohol oxidase and old yellow enzyme. ACS Catal. 2022;12:1111‐1116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Savino S, Fraaije MW. The vast repertoire of carbohydrate oxidases: an overview. Biotechnol Adv. 2021;51:107634. [DOI] [PubMed] [Google Scholar]
- 54. Murray WD, Duff SJB. Bio‐oxidation of aliphatic and aromatic high molecular weight alcohols by Pichia pastoris alcohol oxidase. Appl Microbiol Biotechnol. 1990;33:202‐205. [Google Scholar]
- 55. Dienys G, Jarmalavičius S, Budrien S, Čitavičius D, Sereikait J. Alcohol oxidase from the yeast Pichia pastoris—a potential catalyst for organic synthesis. J Mol Catal B Enzym. 2003;21:47‐49. [Google Scholar]
- 56. Dijkman WP, Fraaije MW. Discovery and characterization of a 5‐hydroxymethylfurfural oxidase from Methylovorus sp. strain MP688. Appl Environ Microbiol. 2014;80:1082‐1090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Dijkman WP, Binda C, Fraaije MW, Mattevi A. Structure‐based enzyme tailoring of 5‐hydroxymethylfurfural oxidase. ACS Catal. 2015;5:1833‐1839. [Google Scholar]
- 58. Heath RS, Birmingham WR, Thompson MP, Taglieber A, Daviet L, Turner NJ. An engineered alcohol oxidase for the oxidation of primary alcohols. ChemBioChem. 2019;20:276‐281. [DOI] [PubMed] [Google Scholar]
- 59. Thompson MP, Derrington SR, Heath RS, et al. A generic platform for the immobilisation of engineered biocatalysts. Tetrahedron. 2019;75:327‐334. [Google Scholar]
- 60. van Schie MMCH, de Almeida TP, Laudadio G, et al. Biocatalytic synthesis of the Green Note trans‐2‐hexenal in a continuous‐flow microreactor. Beilstein J Org Chem. 2018;14:697‐703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. de Almeida TP, van Schie MMCH, Ma A, et al. Efficient aerobic oxidation of trans‐2‐hexen‐1‐ol using the aryl alcohol oxidase from Pleurotus eryngii . Adv Synth Catal. 2019;361:2668‐2672. [Google Scholar]
- 62. Jankowski N, Koschorreck K, Urlacher VB. High‐level expression of aryl‐alcohol oxidase 2 from Pleurotus eryngii in Pichia pastoris for production of fragrances and bioactive precursors. Appl Microbiol Biotechnol. 2020;104:9205‐9218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Jankowski N, Koschorreck K, Urlacher VB. Aryl‐alcohol‐oxidase‐mediated synthesis of piperonal and other valuable aldehydes. Adv Synth Catal. 2022;364:2364‐2372. [Google Scholar]
- 64. Lappe A, Jankowski N, Albrecht A, Koschorreck K. Characterization of a thermotolerant aryl‐alcohol oxidase from Moesziomyces antarcticus oxidizing 5‐hydroxymethyl‐2‐furancarboxylic acid. Appl Microbiol Biotechnol. 2021;105:8313‐8327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Rembeza E, Boverio A, Fraaije MW, Engqvist MKM. Discovery of two novel oxidases using a high‐throughput activity screen. ChemBioChem. 2022;23:e202100510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Gygli G, de Vries RP, van Berkel WJH. On the origin of vanillyl alcohol oxidases. Fungal Genet Biol. 2018;116:24‐32. [DOI] [PubMed] [Google Scholar]
- 67. Dijkman WP, de Gonzalo G, Mattevi A, Fraaije MW. Flavoprotein oxidases: classification and applications. Appl Microbiol Biotechnol. 2013;97:5177‐5188. [DOI] [PubMed] [Google Scholar]
- 68. Jin J, Mazon H, van den Heuvel RHH, Janssen DB, Fraaije MW. Discovery of a eugenol oxidase from Rhodococcus sp. strain RHA1. FEBS J. 2007;274:2311‐2321. [DOI] [PubMed] [Google Scholar]
- 69. García‐Bofill M, Sutton PW, Straatman H, et al. Biocatalytic synthesis of vanillin by an immobilised eugenol oxidase: high biocatalyst yield by enzyme recycling. Appl Catal Gen. 2021;610:117934. [Google Scholar]
- 70. Mordhorst S, Andexer JN. Round, round we go – strategies for enzymatic cofactor regeneration. Nat Prod Rep. 2020;37:1316‐1333. [DOI] [PubMed] [Google Scholar]
- 71. Dong J, Fernández‐Fueyo E, Hollmann F, et al. Biocatalytic oxidation reactions: a chemist's perspective. Angew Chem Int Ed. 2018;57:9238‐9261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Cea G, Wilson L, Bolívar JM, Markovits A, Illanes A. Effect of chain length on the activity of free and immobilized alcohol dehydrogenase towards aliphatic alcohols. Enzyme Microb Technol. 2009;44:135‐138. [Google Scholar]
- 73. Orbegozo T, Lavandera I, Fabian WMF, Mautner B, de Vries JG, Kroutil W. Biocatalytic oxidation of benzyl alcohol to benzaldehyde via hydrogen transfer. Tetrahedron. 2009;65:6805‐6809. [Google Scholar]
- 74. Cassimjee KE, Marín SR, Berglund P. Synthesis of cyclic polyamines by enzymatic generation of an amino aldehyde in situ . Macromol Rapid Commun. 2012;33:1580‐1583. [DOI] [PubMed] [Google Scholar]
- 75. Leskovac V, Trivić S, Peričin D. The three zinc‐containing alcohol dehydrogenases from baker's yeast, Saccharomyces cerevisiae . FEMS Yeast Res. 2002;2:481‐494. [DOI] [PubMed] [Google Scholar]
- 76. Šalić A, Pindrić K, Zelić B. Bioproduction of food additives hexanal and hexanoic acid in a microreactor. Appl Biochem Biotechnol. 2013;171:2273‐2284. [DOI] [PubMed] [Google Scholar]
- 77. Šalić A, Zelić B. ADH‐catalysed hexanol oxidation with fully integrated NADH regeneration performed in microreactors connected in series. RSC Adv. 2014;4:41714‐41721. [Google Scholar]
- 78. Kirmair L, Seiler DL, Skerra A. Stability engineering of the Geobacillus stearothermophilus alcohol dehydrogenase and application for the synthesis of a polyamide 12 precursor. Appl Microbiol Biotechnol. 2015;99:10501‐10513. [DOI] [PubMed] [Google Scholar]
- 79. Orbegozo T, de Vries JG, Kroutil W. Biooxidation of primary alcohols to aldehydes through hydrogen transfer employing Janibacter terrae . Eur J Org Chem. 2010;2010:3445‐3448. [Google Scholar]
- 80. Blikstad C, Dahlström KM, Salminen TA, Widersten M. Stereoselective oxidation of aryl‐substituted vicinal diols into chiral α‐hydroxy aldehydes by re‐engineered propanediol oxidoreductase. ACS Catal. 2013;3:3016‐3025. [Google Scholar]
- 81. Slabu I, Galman JL, Lloyd RC, Turner NJ. Discovery, engineering and synthetic application of transaminase biocatalysts. ACS Catal. 2017;7:8263‐8284. [Google Scholar]
- 82. Guo F, Berglund P. Transaminase biocatalysis: optimization and application. Green Chem. 2017;19:333‐360. [Google Scholar]
- 83. Mehta PK, Hale TI, Christen P. Aminotransferases: demonstration of homology and division into evolutionary subgroups. Eur J Biochem. 1993;214:549‐561. [DOI] [PubMed] [Google Scholar]
- 84. Cellini B, Montioli R, Paiardini A, et al. Molecular defects of the glycine 41 variants of alanine glyoxylate aminotransferase associated with primary hyperoxaluria type I. Proc Natl Acad Sci USA. 2010;107:2896‐2901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Börner T, Rämisch S, Reddem ER, et al. Explaining operational instability of amine transaminases: substrate‐induced inactivation mechanism and influence of quaternary structure on enzyme–cofactor intermediate stability. ACS Catal. 2017;7:1259‐1269. [Google Scholar]
- 86. Lichman BR, Lamming ED, Pesnot T, Smith JM, Hailes HC, Ward JM. One‐pot triangular chemoenzymatic cascades for the syntheses of chiral alkaloids from dopamine. Green Chem. 2015;17:852‐855. [Google Scholar]
- 87. Wang Y, Tappertzhofen N, Méndez‐Sánchez D, et al. Design and use of de novo cascades for the biosynthesis of new benzylisoquinoline alkaloids. Angew Chem Int Ed. 2019;58:10120‐10125. [DOI] [PubMed] [Google Scholar]
- 88. Contente ML, Dall'Oglio F, Tamborini L, Molinari F, Paradisi F. Highly efficient oxidation of amines to aldehydes with flow‐based biocatalysis. ChemCatChem. 2017;9:3843‐3848. [Google Scholar]
- 89. Contente ML, Paradisi F. Transaminase‐catalyzed continuous synthesis of biogenic aldehydes. ChemBioChem. 2019;20:2830‐2833. [DOI] [PubMed] [Google Scholar]
- 90. Peralta GH, Bergamini CV, Hynes ER. Aminotransferase and glutamate dehydrogenase activities in lactobacilli and streptococci. Braz J Microbiol. 2016;47:741‐748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Yvon M, Thirouin S, Rijnen L, Fromentier D, Gripon JC. An aminotransferase from Lactococcus lactis initiates conversion of amino acids to cheese flavor compounds. Appl Environ Microbiol. 1997;63:414‐419. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Batista VF, Galman JL, Pinto DC, Silva AMS, Turner NJ. Monoamine oxidase: tunable activity for amine resolution and functionalization. ACS Catal. 2018;8:11889‐11907. [Google Scholar]
- 93. Finberg JPM, Youdim MBH, Riederer P, Tipton KF. MAO‐the Mother of All Amine Oxidases. Springer; 2013. [Google Scholar]
- 94. Hare MLC. Tyramine oxidase: a new enzyme system in liver. Biochem J. 1928;22:968‐979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Atkin KE, Reiss R, Koehler V, et al. The structure of monoamine oxidase from Aspergillus niger provides a molecular context for improvements in activity obtained by directed evolution. J Mol Biol. 2008;384:1218‐1231. [DOI] [PubMed] [Google Scholar]
- 96. Roh JH, Wouters J, Depiereux E, et al. Purification, cloning, and three‐dimensional structure prediction of Micrococcus luteus FAD‐containing tyramine oxidase. Biochem Biophys Res Commun. 2000;268:293‐297. [DOI] [PubMed] [Google Scholar]
- 97. Vianello R, Repič M, Mavri J. How are biogenic amines metabolized by monoamine oxidases? Eur J Org Chem. 2012;2012:7057‐7065. [Google Scholar]
- 98. Chajkowski‐Scarry S, Rimoldi JM. Monoamine oxidase A and B substrates: probing the pathway for drug development. Future Med Chem. 2014;6:697‐717. [DOI] [PubMed] [Google Scholar]
- 99. Gaweska H, Fitzpatrick PF. Structures and mechanism of the monoamine oxidase family. Biomol Concepts. 2011;2:365‐377. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Renson J, Weissbach H, Udenfriend S. Studies on the biological activities of the aldehydes derived from norepinephrine, serotonin, tryptamine and histamine. J Pharmacol Exp Ther. 1964;143:326‐331. [PubMed] [Google Scholar]
- 101. Hoover LK, Moo‐Young M, Legge RL. Biotransformation of dopamine to norlaudanosoline by Aspergillus niger . Biotechnol Bioeng. 1991;38:1029‐1033. [DOI] [PubMed] [Google Scholar]
- 102. Ghislieri D, Green AP, Pontini M, et al. Engineering an enantioselective amine oxidase for the synthesis of pharmaceutical building blocks and alkaloid natural products. J Am Chem Soc. 2013;135:10863‐10869. [DOI] [PubMed] [Google Scholar]
- 103. Carr R, Alexeeva M, Enright A, Eve TSC, Dawson MJ, Turner NJ. Directed evolution of an amine oxidase possessing both broad substrate specificity and high enantioselectivity. Angew Chem Int Ed Engl. 2003;42:4807‐4810. [DOI] [PubMed] [Google Scholar]
- 104. Herter S, Medina F, Wagschal S, Benhaïm C, Leipold F, Turner NJ. Mapping the substrate scope of monoamine oxidase (MAO‐N) as a synthetic tool for the enantioselective synthesis of chiral amines. Bioorg Med Chem. 2018;26:1338‐1346. [DOI] [PubMed] [Google Scholar]
- 105. Toscani A, Risi C, Black GW, et al. Monoamine oxidase (MAO‐N) whole cell biocatalyzed aromatization of 1, 2, 5, 6‐tetrahydropyridines into pyridines. ACS Catal. 2018;8:8781‐8787. [Google Scholar]
- 106. Huang G, Zhu F, Chen Y, et al. A spectrophotometric assay for monoamine oxidase activity with 2, 4‐dinitrophenylhydrazine as a derivatized reagent. Anal Biochem. 2016;512:18‐25. [DOI] [PubMed] [Google Scholar]
- 107. Janes SM, Mure M, Klinman JP, et al. Identification of topaquinone and its consensus sequence in copper amine oxidases. Biochemistry. 1992;31:12147‐12154. [DOI] [PubMed] [Google Scholar]
- 108. Mure M, Mills SA, Klinman JP. Catalytic mechanism of the topa quinone containing copper amine oxidases. Biochemistry. 2002;41:9269‐9278. [DOI] [PubMed] [Google Scholar]
- 109. Elmore BO, Bollinger JA, Dooley DM. Human kidney diamine oxidase: heterologous expression, purification, and characterization. J Biol Inorg Chem. 2002;7:565‐579. [DOI] [PubMed] [Google Scholar]
- 110. McGrath AP, Hilmer KM, Collyer CA, et al. Structure and inhibition of human diamine oxidase. Biochemistry. 2009;48:9810‐9822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. Cooper A, Equi AM, Ner SK, Watson AB, Robins DJ. Oxidation of N‐alkylputrescines by diamine oxidases. Tetrahedron Lett. 1989;30:5167‐5170. [Google Scholar]
- 112. Mergemeier K, Lehr M. HPLC‐UV assays for evaluation of inhibitors of mono and diamine oxidases using novel phenyltetrazolylalkanamine substrates. Anal Biochem. 2018;549:29‐38. [DOI] [PubMed] [Google Scholar]
- 113. Bonamore A, Calisti L, Calcaterra A, et al. A novel enzymatic strategy for the synthesis of substituted tetrahydroisoquinolines. ChemistrySelect. 2016;1:1525‐1528. [Google Scholar]
- 114. Isobe K, Sasaki T, Aigami Y, Yamada M, Kishino S, Ogawa J. Characterization of a new enzyme oxidizing ω‐amino group of aminocarboxyric acid, aminoalcohols and amines from Phialemonium sp. AIU 274. J Mol Catal B: Enzym. 2013;96:89‐95. [Google Scholar]
- 115. Di Fabio E, Incocciati A, Boffi A, Bonamore A, Macone A. Biocatalytic production of aldehydes: exploring the potential of Lathyrus cicera amine oxidase. Biomolecules. 2021;11:1540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Di Fusco M, Federico R, Boffi A, MacOne A, Favero G, Mazzei F. Characterization and application of a diamine oxidase from Lathyrus sativus as component of an electrochemical biosensor for the determination of biogenic amines in wine and beer. Anal Bioanal Chem. 2011;401:707‐716. [DOI] [PubMed] [Google Scholar]
- 117. Rosini E, Tonin F, Vasylieva N, Marinesco S, Pollegioni L. Evolution of histamine oxidase activity for biotechnological applications. Appl Microbiol Biotechnol. 2014;98:739‐748. [DOI] [PubMed] [Google Scholar]
- 118. Sadrameli SM. Thermal/catalytic cracking of hydrocarbons for the production of olefins: a state‐of‐the‐art review I: thermal cracking review. Fuel. 2015;140:102‐115. [Google Scholar]
- 119. Sadrameli SM. Thermal/catalytic cracking of liquid hydrocarbons for the production of olefins: a state‐of‐the‐art review II: catalytic cracking review. Fuel. 2016;173:285‐297. [Google Scholar]
- 120. Kumawat J, Gupta VK. Fundamental aspects of heterogeneous Ziegler–Natta olefin polymerization catalysis: an experimental and computational overview. Polym Chem. 2020;11:6107‐6128. [Google Scholar]
- 121. Mol JC. Industrial applications of olefin metathesis. J Mol Catal. 2004;213:39‐45. [Google Scholar]
- 122. Reuben B, Wittcoff H. The SHOP process: an example of industrial creativity. J Chem Educ. 1988;65:605‐607. [Google Scholar]
- 123. Rajagopalan A, Lara M, Kroutil W. Oxidative alkene cleavage by chemical and enzymatic methods. Adv Synth Catal. 2013;355:3321‐3335. [Google Scholar]
- 124. Tuynman A, Spelberg JL, Kooter IM, Schoemaker HE, Wever R. Enantioselective epoxidation and carbon–carbon bond cleavage catalyzed by Coprinus cinereus peroxidase and myeloperoxidase. J Biol Chem. 2000;275:3025‐3030. [DOI] [PubMed] [Google Scholar]
- 125. Marasco EK, Vay K, Schmidt‐Dannert C. Identification of carotenoid cleavage dioxygenases from Nostoc sp. PCC 7120 with different cleavage activities. J Biol Chem. 2006;281:31583‐31593. [DOI] [PubMed] [Google Scholar]
- 126. Schmidt H, Kurtzer R, Eisenreich W, Schwab W. The carotenase AtCCD1 from Arabidopsis thaliana is a dioxygenase. J Biol Chem. 2006;281:9845‐9851. [DOI] [PubMed] [Google Scholar]
- 127. Marasco EK, Schmidt‐Dannert C. Identification of bacterial carotenoid cleavage dioxygenase homologues that cleave the interphenyl α, β double bond of stilbene derivatives via a monooxygenase reaction. ChemBioChem. 2008;9:1450‐1461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Ryu JY, Seo J, Park S, et al. Characterization of an isoeugenol monooxygenase (Iem) from Pseudomonas nitroreducens Jin1 that transforms isoeugenol to vanillin. Biosci Biotechnol Biochem. 2013;77:289‐294. [DOI] [PubMed] [Google Scholar]
- 129. Yamada M, Okada Y, Yoshida T, Nagasawa T. Purification, characterization and gene cloning of isoeugenol‐degrading enzyme from Pseudomonas putida IE27. Arch Microbiol. 2007;187:511‐517. [DOI] [PubMed] [Google Scholar]
- 130. Han Z, Long L, Ding S. Expression and characterization of carotenoid cleavage oxygenases from Herbaspirillum seropedicae and Rhodobacteraceae bacterium capable of biotransforming isoeugenol and 4‐vinylguaiacol to vanillin. Front Microbiol. 2019;10:1869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Ni J, Wu Y‐T, Tao F, Peng Y, Xu P. A coenzyme‐free biocatalyst for the value‐added utilization of lignin‐derived aromatics. J Am Chem Soc. 2018;140:16001‐16005. [DOI] [PubMed] [Google Scholar]
- 132. Furuya T, Miura M, Kino K. A coenzyme‐independent decarboxylase/oxygenase cascade for the efficient synthesis of vanillin. ChemBioChem. 2014;15:2248‐2254. [DOI] [PubMed] [Google Scholar]
- 133. Tang J, Shi L, Li L, Long L, Ding S. Expression and characterization of a 9‐cis‐epoxycarotenoid dioxygenase from Serratia sp. ATCC 39006 capable of biotransforming isoeugenol and 4‐vinylguaiacol to vanillin. Biotechnol Rep. 2018;18:e00253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134. Yamada M, Okada Y, Yoshida T, Nagasawa T. Biotransformation of isoeugenol to vanillin by Pseudomonas putida IE27 cells. Appl Microbiol Biotechnol. 2007;73:1025‐1030. [DOI] [PubMed] [Google Scholar]
- 135. Qi Z, Tong X, Zhang X, Lin H, Bu S, Zhao L. One‐pot synthesis of dihydro‐β‐ionone from carotenoids using carotenoid cleavage dioxygenase and enoate reductase. Bioprocess Biosyst Eng. 2022;45:891‐900. [DOI] [PubMed] [Google Scholar]
- 136. Fahlbusch K‐G, Hammerschmidt F‐J, Panten J, et al. Flavors and Fragrances. Wiley‐VCH Verlag GmbH & Co. KGaA; 2003. [Google Scholar]
- 137. Schmitt G, Seiffert G, Kroneck PMH, Braaz R, Jendrossek D. Spectroscopic properties of rubber oxygenase RoxA from Xanthomonas sp., a new type of dihaem dioxygenase. Microbiology. 2010;156:2537‐2548. [DOI] [PubMed] [Google Scholar]
- 138. Nacke C, Schrader J. Micelle based delivery of carotenoid substrates for enzymatic conversion in aqueous media. J Mol Catal B Enzym. 2012;77:67‐73. [Google Scholar]
- 139. Mutti FG. Alkene cleavage catalysed by heme and nonheme enzymes: reaction mechanisms and biocatalytic applications. Bioinorg Chem Appl. 2012;2012:626909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Zhao L, Xie Y, Chen L, Xu X, Zhao CX, Cheng F. Efficient biotransformation of isoeugenol to vanillin in recombinant strains of Escherichia coli by using engineered isoeugenol monooxygenase and sol‐gel chitosan membrane. Process Biochem. 2018;71:76‐81. [Google Scholar]
- 141. Long Z, Duan N, Xue Y, et al. Characterization of a novel lutein cleavage dioxygenase, EhLCD, from Enterobacter hormaechei YT‐3 for the enzymatic synthesis of 3‐hydroxy‐β‐ionone from lutein. Catalysts. 2021;11:1257. [Google Scholar]
- 142. Zhao Y, Zhong GF, Yang XP, Hu XM, Bin Mao D, Ma YP. Bioconversion of lutein to form aroma compounds by Pantoea dispersa . Biotechnol Lett. 2015;37:1687‐1692. [DOI] [PubMed] [Google Scholar]
- 143. Zhong G, Wang F, Sun J, et al. Bioconversion of lutein by Enterobacter hormaechei to form a new compound, 8‐methyl‐α‐ionone. Biotechnol Lett. 2017;39:1019‐1024. [DOI] [PubMed] [Google Scholar]
- 144. Dunwell JM, Culham A, Carter CE, Sosa‐Aguirre CR, Goodenough PW. Evolution of functional diversity in the cupin superfamily. Trends Biochem Sci. 2001;26:740‐746. [DOI] [PubMed] [Google Scholar]
- 145. Dunwell JM, Purvis A, Khuri S. Cupins: the most functionally diverse protein superfamily? Phytochemistry. 2004;65:7‐17. [DOI] [PubMed] [Google Scholar]
- 146. Grill B. Engineering of an Alkene Cleavage Enzyme. Master Thesis. Graz University of Technology; 2017. [Google Scholar]
- 147. Hajnal I, Faber K, Schwab H, Hall M, Steiner K. Oxidative alkene cleavage catalysed by manganese‐dependent cupin TM1459 from Thermotoga maritima . Adv Synth Catal. 2015;357:3309‐3316. [Google Scholar]
- 148. Fahlbusch KG, Hammerschmidt FJ, Panten J, et al. Flavors and fragrances. In: Ullmann's Encyclopedia of Industrial Chemistry. Wiley‐VCH Verlag GmbH & Co. KGaA; 2003:73‐198. [Google Scholar]
- 149. Bugg TDH. Dioxygenase enzymes and oxidative cleavage pahtways. Chem Mol Sci Chem Eng. 2010;8:583‐623. [Google Scholar]
- 150. Spence EL, Langley GJ, Bugg TDH. Cis‐trans isomerisation of a cyclopropyl radical trap catalyzed by extradiol catechol dioxygenases: evidence for a semiquinone intermediate. J Am Chem Soc. 1996;118:8336‐8343. [Google Scholar]
- 151. Vaillancourt FH, Barbosa CJ, Spiro TG, et al. Definitive evidence for monoanionic binding of 2, 3‐dihydroxybiphenyl to 2, 3‐dihydroxybiphenyl 1, 2‐dioxygenase from UV resonance Raman spectroscopy, UV/Vis absorption spectroscopy, and crystallography. J Am Chem Soc. 2002;124:2485‐2496. [DOI] [PubMed] [Google Scholar]
- 152. Sanvoisin J, Langley GJ, Bugg TDH. Mechanism of extradiol catechol dioxygenases: evidence for a lactone intermediate in the 2,3‐dihydroxyphenylpropionate 1,2‐dioxygenase reaction. J Am Chem Soc. 1995;117:7836‐7837. [Google Scholar]
- 153. Bai X, Nie M, Diwu Z, et al. Preparation of 2‐hydroxymuconic semialdehyde from catechol by combining enzymatic catalysis with bisulfite nucleophilic addition. J Environ Chem Eng. 2021;9:105970. [Google Scholar]
- 154. Straganz GD, Glieder A, Brecker L, Ribbons DW, Steiner W. Acetylacetone‐cleaving enzyme Dke1: a novel CC‐bond‐cleaving enzyme from Acinetobacter johnsonii . Biochem J. 2003;369:573‐581. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Straganz G, Brecker L, Weber HJ, Steiner W, Ribbons DW. A novel β‐diketone‐cleaving enzyme from Acinetobacter johnsonii: acetylacetone 2, 3‐oxygenase. Biochem Biophys Res Commun. 2002;297:232‐236. [DOI] [PubMed] [Google Scholar]
- 156. Diebold AR, Neidig ML, Moran GR, Straganz GD, Solomon EI. The three‐his triad in Dke1: comparisons to the classical facial triad. Biochemistry. 2010;49:6945‐6952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157. Straganz GD, Nidetzky B. Reaction coordinate analysis for β‐diketone cleavage by the non‐heme Fe2+‐dependent dioxygenase Dke1. J Am Chem Soc. 2005;127:12306‐12314. [DOI] [PubMed] [Google Scholar]
- 158. Straganz GD, Hofer H, Steiner W, Nidetzky B. Electronic substituent effects on the cleavage specificity of a non‐heme Fe2+‐dependent β‐diketone dioxygenase and their mechanistic implications. J Am Chem Soc. 2004;126:12202‐12203. [DOI] [PubMed] [Google Scholar]
- 159. Zhang Y, Kang SA, Mukherjee T, et al. Crystal structure and mechanism of tryptophan 2, 3‐dioxygenase, a heme enzyme involved in tryptophan catabolism and in quinolinate biosynthesis. Biochemistry. 2007;46:145‐155. [DOI] [PubMed] [Google Scholar]
- 160. Sugimoto H, Oda SI, Otsuki T, Hino T, Yoshida T, Shiro Y. Crystal structure of human indoleamine 2, 3‐dioxygenase: catalytic mechanism of O2 incorporation by a heme‐containing dioxygenase. Proc Natl Acad Sci USA. 2006;103:2611‐2616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161. Lewis‐Ballester A, Forouhar F, Kim SM, et al. Molecular basis for catalysis and substrate‐mediated cellular stabilization of human tryptophan 2, 3‐dioxygenase. Sci Rep. 2016;6:35169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162. Pilotte L, Larrieu P, Stroobant V, et al. Reversal of tumoral immune resistance by inhibition of tryptophan 2, 3‐dioxygenase. Proc Natl Acad Sci USA. 2012;109:2497‐2502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163. Routy JP, Routy B, Graziani GM, Mehraj V. The kynurenine pathway is a double‐edged sword in immune‐privileged sites and in cancer: implications for immunotherapy. Int J Tryptophan Res. 2016;9:67‐77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164. Thackray SJ, Mowat CG, Chapman SK. Exploring the mechanism of tryptophan 2, 3‐dioxygenase. Biochem Soc Trans. 2008;36:1120‐1123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165. Makino R, Obayashi E, Hori H, et al. Initial O2 insertion step of the tryptophan dioxygenase reaction proposed by a heme‐modification study. Biochemistry. 2015;54:3604‐3616. [DOI] [PubMed] [Google Scholar]
- 166. Yanagisawa S, Sugimoto H, Shiro Y, Ogura T. A specific interaction of L‐tryptophan with CO of CO‐bound indoleamine 2, 3‐dioxygenase identified by resonance Raman spectroscopy. Biochemistry. 2010;49:10081‐10088. [DOI] [PubMed] [Google Scholar]
- 167. Lewis‐Ballester A, Batabyal D, Egawa T, et al. Evidence for a ferryl intermediate in a heme‐based dioxygenase. Proc Natl Acad Sci USA. 2009;106:17371‐17376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Yanagisawa S, Yotsuya K, Hashiwaki Y, et al. Identification of the Fe–O2 and the Fe=O heme species for indoleamine 2, 3‐dioxygenase during catalytic turnover. Chem Lett. 2010;39:36‐37. [Google Scholar]
- 169. Braaz R, Armbruster W, Jendrossek D. Heme‐dependent rubber oxygenase RoxA of Xanthomonas sp. cleaves the carbon backbone of poly(cis‐1,4‐isoprene) by a dioxygenase mechanism. Appl Environ Microbiol. 2005;71:2473‐2478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170. Birke J, Röther W, Schmitt G, Jendrossek D. Functional identification of rubber oxygenase (RoxA) in soil and marine myxobacteria. Appl Environ Microbiol. 2013;79:6391‐6399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171. Birke J, Röther W, Jendrossek D. RoxB is a novel type of rubber oxygenase that combines properties of rubber oxygenase RoxA and latex clearing protein (Lcp). Appl Environ Microbiol. 2017;83:e00721‐e00717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172. Jendrossek D, Reinhardt S. Sequence analysis of a gene product synthesised by Xanthomonas sp. during growth on natural rubber latex. FEMS Microbiol Lett. 2003;224:61‐65. [DOI] [PubMed] [Google Scholar]
- 173. Cressey P, Reeve J. Metabolism of cyanogenic glycosides: a review. Food Chem Toxicol. 2019;125:225‐232. [DOI] [PubMed] [Google Scholar]
- 174. Bröckner A, Raspotnig G, Wehner K, Meusinger R, Norton RA, Heethoff M. Storage and release of hydrogen cyanide in a chelicerate (Oribatula tibialis). Proc Natl Acad Sci USA. 2017;114:3469‐3472. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175. Bhalla TC, Kumar V, Kumar V, Thakur N. Nitrile metabolizing enzymes in biocatalysis and biotransformation. Appl Biochem Biotechnol. 2018;185:925‐946. [DOI] [PubMed] [Google Scholar]
- 176. Lanfranchi E, Steiner K, Glieder A, et al. Mini‐review: recent developments in hydroxynitrile lyases for industrial biotechnology. Recent Pat Biotechnol. 2013;7:197‐206. [DOI] [PubMed] [Google Scholar]
- 177. Bracco P, Busch H, von Langermann J, Hanefeld U. Enantioselective synthesis of cyanohydrins catalysed by hydroxynitrile lyases – a review. Org Biomol Chem. 2016;14:6375‐6389. [DOI] [PubMed] [Google Scholar]
- 178. Fuchs G, Boll M, Heider J. Microbial degradation of aromatic compounds – from one strategy to four. Nat Rev Microbiol. 2011;9:803‐816. [DOI] [PubMed] [Google Scholar]
- 179. Gigot C, Ongena M, Fauconnier M‐L, Wathelet J‐P, du Jardin P, Thonart P. The lipoxygenase metabolic pathway in plants: potential for industrial production of natural green leaf volatiles. Biotechnol Agron Soc Environ. 2010;14:451‐460. [Google Scholar]
- 180. Stolterfoht H, Rinnofner C, Winkler M, Pichler H. Recombinant lipoxygenases and hydroperoxide lyases for the synthesis of green leaf volatiles. J Agric Food Chem. 2019;67:13367‐13392. [DOI] [PubMed] [Google Scholar]
- 181. Galliard T, Matthew JA. The enzymic formation of long chain aldehydes and alcohols by α‐oxidation of fatty acids in extracts of cucumber fruit (Cucumis sativus). Biochim Biophys Acta Lipids Lipid Metab. 1976;424:26‐35. [DOI] [PubMed] [Google Scholar]
- 182. Hammer AK, Albrecht F, Hahne F, et al. Biotechnological production of odor‐active methyl‐branched aldehydes by a novel α‐dioxygenase from Crocosphaera subtropica . J Agric Food Chem. 2020;68:10432‐10440. [DOI] [PubMed] [Google Scholar]
- 183. Koszelak‐Rosenblum M, Krol AC, Simmons DM, Goulah CC, Wroblewski L, Malkowski MG. His‐311 and Arg‐559 are key residues involved in fatty acid oxygenation in pathogen‐inducible oxygenase. J Biol Chem. 2008;283:24962‐24971. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184. Kim IJ, Brack Y, Bayer T, Bornscheuer UT. Two novel cyanobacterial α‐dioxygenases for the biosynthesis of fatty aldehydes. Appl Microbiol Biotechnol. 2022;106:197‐210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185. Kim IJ, Bayer T, Terholsen H, Bornscheuer UT. α‐Dioxygenases (α‐DOXs): promising biocatalysts for the environmentally friendly production of aroma compounds. ChemBioChem. 2022;23:e202100693. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186. Kanter J‐P, Honold PJ, Lüke D, et al. An enzymatic tandem reaction to produce odor‐active fatty aldehydes. Appl Microbiol Biotechnol. 2022;106:6095‐6107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187. Mukhtarova LS, Brühlmann F, Hamberg M, Khairutdinov BI, Grechkin AN. Plant hydroperoxide‐cleaving enzymes (CYP74 family) function as hemiacetal synthases: structural proof of hemiacetals by NMR spectroscopy. Biochim Biophys Acta Mol Cell Biol Lipids. 2018;1863:1316‐1322. [DOI] [PubMed] [Google Scholar]
- 188. Schneider C, Pratt DA, Porter NA, Brash AR. Control of oxygenation in lipoxygenase and cyclooxygenase catalysis. Chem Biol. 2007;14:473‐488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189. Kazimírová V, Zezulová V, Krasňan V, Štefuca V, Rebroš M. Optimization of hydroperoxide lyase production for recombinant lipoxygenase pathway cascade application. Catalysts. 2021;11:1201. [Google Scholar]
- 190. Kaur I, Korrapati N, Bonello J, Mukherjee A, Rishi V, Bendigiri C. Biosynthesis of natural aroma compounds using recombinant whole‐cell tomato hydroperoxide lyase biocatalyst. J Biosci. 2022;47:37. [PubMed] [Google Scholar]
- 191. Brühlmann F, Bosijokovic B. Efficient biochemical cascade for accessing green leaf alcohols. Org Process Res Dev. 2016;20:1974‐1978. [Google Scholar]
- 192. Hartmans S, Smits JP, Van der Werf MJ, Volkering F, De Bont JAM. Metabolism of styrene oxide and 2‐phenylethanol in the styrene‐degrading Xanthobacter strain 124X. Appl Environ Microbiol. 1989;55:2850‐2855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Xin R, See WWL, Yun H, Li X, Li Z. Enzyme‐catalyzed Meinwald rearrangement with an unusual regioselective and stereospecific 1, 2‐methyl shift. Angew Chemie Int Ed. 2022;61:e202204889. [DOI] [PubMed] [Google Scholar]
- 194. Choo JPS, Li Z. Styrene oxide isomerase catalyzed Meinwald rearrangement reaction: discovery and application in single‐step and one‐pot cascade reactions. Org Process Res Dev. 2022;26:1960‐1970. [Google Scholar]
- 195. Choo JPS, Kammerer RA, Li X, Li Z. High‐level production of phenylacetaldehyde using fusion‐tagged styrene oxide isomerase. Adv Synth Catal. 2021;363:1714‐1721. [Google Scholar]
- 196. Mo Q, Chen H, Fan C, et al. Utilization of a styrene‐derived pathway for 2‐phenylethanol production in budding yeast. Appl Microbiol Biotechnol. 2021;105:2333‐2340. [DOI] [PubMed] [Google Scholar]
- 197. Rather LJ, Knapp B, Haehnel W, Fuchs G. Coenzyme A‐dependent aerobic metabolism of benzoate via epoxide formation. J Biol Chem. 2010;285:20615‐20624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198. Luo Z, Yu S, Zeng W, Zhou J. Comparative analysis of the chemical and biochemical synthesis of keto acids. Biotechnol Adv. 2021;47:107706. [DOI] [PubMed] [Google Scholar]
- 199. Zhu Y, Eiteman MA, Altman R, Altman E. High glycolytic flux improves pyruvate production by a metabolically engineered Escherichia coli Strain. Appl Environ Microbiol. 2008;74:6649‐6655. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200. Song Y, Li J, dong Shin H, Liu L, Du G, Chen J. Biotechnological production of alpha‐keto acids: current status and perspectives. Bioresour Technol. 2016;219:716‐724. [DOI] [PubMed] [Google Scholar]
- 201. Li Y, Chen J, Lun SY. Biotechnological production of pyruvic acid. Appl Microbiol Biotechnol. 2001;57:451‐459. [DOI] [PubMed] [Google Scholar]
- 202. Li T, Huo L, Pulley C, Liu A. Decarboxylation mechanisms in biological system. Bioorg Chem. 2012;43:2‐14. [DOI] [PubMed] [Google Scholar]
- 203. Eram M, Ma K. Decarboxylation of pyruvate to acetaldehyde for ethanol production by hyperthermophiles. Biomolecules. 2013;3:578‐596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204. König S, Spinka M, Kutter S. Allosteric activation of pyruvate decarboxylases. A never‐ending story? J Mol Catal B Enzym. 2009;61:100‐110. [Google Scholar]
- 205. Bongers RS, Hoefnagel MHN, Kleerebezem M. High‐level acetaldehyde production in Lactococcus lactis by metabolic engineering. Appl Environ Microbiol. 2005;71:1109‐1113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206. Balagurunathan B, Tan L, Zhao H. Metabolic engineering of Escherichia coli for acetaldehyde overproduction using pyruvate decarboxylase from Zymomonas mobilis . Enzyme Microb Technol. 2018;109:58‐65. [DOI] [PubMed] [Google Scholar]
- 207. Tsou AY, Ransom SC, Gerlt JA, Buechter DD, Babbitt PC, Kenyon GL. Mandelate pathway of Pseudomonas putida: sequence relationships involving mandelate racemase, (S)‐mandelate dehydrogenase, and benzoylformate decarboxylase and expression of benzoylformate decarboxylase in Escherichia coli . Biochemistry. 1990;29:9856‐9862. [DOI] [PubMed] [Google Scholar]
- 208. Paula Dionísio A, Molina G, Souza de Carvalho D, dos Santos R, Bicas JL, Pastore GM. Natural Food Additives, Ingredients and Flavourings. Elsevier; 2012:231‐259. [Google Scholar]
- 209. Gocke D, Graf T, Brosi H, et al. Comparative characterisation of thiamin diphosphate‐dependent decarboxylases. J Mol Catal B Enzym. 2009;61:30‐35. [Google Scholar]
- 210. Wei J, Timler JG, Knutson CM, Barney BM. Branched‐chain 2‐keto acid decarboxylases derived from Psychrobacter . FEMS Microbiol Lett. 2013;346:105‐112. [DOI] [PubMed] [Google Scholar]
- 211. Boumba VA, Ziavrou KS, Vougiouklakis T. Biochemical pathways generating post‐mortem volatile compounds co‐detected during forensic ethanol analyses. Forensic Sci Int. 2008;174:133‐151. [DOI] [PubMed] [Google Scholar]
- 212. Smit BA, Van Hylckama Vlieg JET, Engels WJM, Meijer L, Wouters JTM, Smit G. Identification, cloning, and characterization of a Lactococcus lactis branched‐chain α‐keto acid decarboxylase involved in flavor formation. Appl Environ Microbiol. 2005;71:303‐311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213. Smit BA, Engels WJM, Smit G. Branched chain aldehydes: production and breakdown pathways and relevance for flavour in foods. Appl Microbiol Biotechnol. 2009;81:987‐999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214. De La Plaza M, Fernández de Palencia P, Peláez C, Requena T. Biochemical and molecular characterization of α‐ketoisovalerate decarboxylase, an enzyme involved in the formation of aldehydes from amino acids by Lactococcus lactis . FEMS Microbiol Lett. 2004;238:367‐374. [DOI] [PubMed] [Google Scholar]
- 215. Luo J, Jiang C, Zhao L, et al. Keto acid decarboxylase and keto acid dehydrogenase activity detected during the biosynthesis of flavor compound 3‐methylbutanal by the nondairy adjunct culture Lactococcus lactis ssp. lactis F9. J Dairy Sci. 2018;101:9725‐9735. [DOI] [PubMed] [Google Scholar]
- 216. Chen C, Yuan J, Yu H, et al. Cloning, purification, and characterization of branched‐chain α‐keto acid decarboxylases from Lactococcus lactis strains with different 3‐methylbutanal production abilities. Food Biosci. 2022;47:101713. [Google Scholar]
- 217. Chen C, Yuan J, Yu H, et al. Characterization of metabolic pathways for biosynthesis of the flavor compound 3‐methylbutanal by Lactococcus lactis . J Dairy Sci. 2022;105:97‐108. [DOI] [PubMed] [Google Scholar]
- 218. Zhao L, Chen Z, Lin S, Wu T, Yu S, Huo YX. In vitro biosynthesis of isobutyraldehyde through the establishment of a one‐step self‐assembly‐based immobilization strategy. J Agric Food Chem. 2021;69:14609‐14619. [DOI] [PubMed] [Google Scholar]
- 219. LeVieux JA, Baas BJ, Kaoud TS, et al. Kinetic and structural characterization of a cis‐3‐chloroacrylic acid dehalogenase homologue in Pseudomonas sp. UW4: a potential step between subgroups in the tautomerase superfamily. Arch Biochem Biophys. 2017;636:50‐56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220. Leffingwell JC, Leffingwell D. Flavours & fragrances: recent advances in biotechnology. Spec Chem Mag. 2015;35:32‐34. [Google Scholar]
- 221. Wang Q, Wu X, Lu X, He Y, Ma B, Xu Y. Efficient biosynthesis of vanillin from isoeugenol by recombinant isoeugenol monooxygenase from Pseudomonas nitroreducens Jin1. Appl Biochem Biotechnol. 2021;193:1116‐1128. [DOI] [PubMed] [Google Scholar]
- 222. Schwendenwein D, Fiume G, Weber H, Rudroff F, Winkler M. Selective enzymatic transformation to aldehydes in vivo by fungal carboxylate reductase from Neurospora crassa . Adv Synth Catal. 2016;358:3414‐3421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223. Bühler B, Bollhalder I, Hauer B, Witholt B, Schmid A. Use of the two‐liquid phase concept to exploit kinetically controlled multistep biocatalysis. Biotechnol Bioeng. 2003;81:683‐694. [DOI] [PubMed] [Google Scholar]
- 224. Balodite E, Strazdina I, Martynova J, et al. Translocation of Zymomonas mobilis pyruvate decarboxylase to periplasmic compartment for production of acetaldehyde outside the cytosol. Microbiol Open. 2019;8:e00809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225. Duff SJB, Murray WD. Production of flavor aldehydes using nongrowing whole cells of Pichia pastoris . Ann N Y Acad Sci. 1988;542:428‐433. [Google Scholar]
- 226. Rodriguez GM, Atsumi S. Isobutyraldehyde production from Escherichia coli by removing aldehyde reductase activity. Microb Cell Fact. 2012;11:90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227. Ben Akacha N, Gargouri M. Enzymatic synthesis of green notes with hydroperoxide‐lyase from olive leaves and alcohol‐dehydrogenase from yeast in liquid/gas reactor. Process Biochem. 2009;44:1122‐1127. [Google Scholar]
- 228. Horvat M, Winkler M. In vivo reduction of medium‐ to long‐chain fatty acids by carboxylic acid reductase (CAR) enzymes: limitations and solutions. ChemCatChem. 2020;12:5076‐5090. [Google Scholar]
- 229. Son J, Choi IH, Lim CG, et al. Production of cinnamaldehyde through whole‐cell bioconversion fromtrans‐cinnamic acid using engineered Corynebacterium glutamicum . J Agric Food Chem. 2022;70:2656‐2663. [DOI] [PubMed] [Google Scholar]
- 230. Bang HB, Son J, Kim SC, Jeong KJ. Systematic metabolic engineering of Escherichia coli for the enhanced production of cinnamaldehyde. Metab Eng. 2023;76:63‐74. [DOI] [PubMed] [Google Scholar]
- 231. Chang H, Bajaj I, Motagamwala AH, et al. Sustainable production of 5‐hydroxymethyl furfural from glucose for process integration with high fructose corn syrup infrastructure. Green Chem. 2021;23:3277‐3288. [Google Scholar]
- 232. Wang Y, Brown CA, Chen R. Industrial production, application, microbial biosynthesis and degradation of furanic compound, hydroxymethylfurfural (HMF). AIMS Microbiol. 2018;4:261‐273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233. Titchiner GR, Marshall SA, Miscikas H, Leys D. Biosynthesis of pyrrole‐2‐carbaldehyde via enzymatic CO2 fixation. Catalysts. 2022;12:538. [Google Scholar]
- 234. Gahloth D, Fisher K, Payne KAP, Cliff M, Levy C, Leys D. Structural and biochemical characterization of the prenylated flavin mononucleotide‐dependent indole‐3‐carboxylic acid decarboxylase. J Biol Chem. 2022;298:101771. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
Data sharing is not applicable to this article as no new data were created or analyzed in this study.
