Skip to content
BY 4.0 license Open Access Published online by De Gruyter April 17, 2023

Light-driven bioprocesses

  • Michael E. Runda and Sandy Schmidt ORCID logo EMAIL logo
From the journal Physical Sciences Reviews

Abstract

Enzyme catalysis and photocatalysis are two research areas that have become of major interest in organic synthesis. This is mainly because both represent attractive strategies for making chemical synthesis more efficient and sustainable. Because enzyme catalysis offers several inherent advantages, such as high substrate specificity, regio-, and stereoselectivity, and activity under environmentally benign reaction conditions, biocatalysts are increasingly being adopted by the pharmaceutical and chemical industries. In addition, photocatalysis has proven to be a powerful approach for accessing unique reactivities upon light irradiation and performing reactions with an extended substrate range under milder conditions compared to light-independent alternatives. It is therefore not surprising that bio- and photocatalytic approaches are now often combined to exploit the exquisite selectivity of enzymes and the unique chemical transformations accessible to photocatalysis. In this chapter, we provide an overview of the wide variety of light-driven bioprocesses, ranging from photochemical delivery of reducing equivalents to redox enzymes, photochemical cofactor regeneration, to direct photoactivation of enzymes. We also highlight the possibility of catalyzing non-natural reactions via photoinduced enzyme promiscuity and the combination of photo- and biocatalytic reactions used to create new synthetic methodologies.

1 Introduction

Light represents an ideal ‘reagent’ in chemical synthesis as it is safe to use and readily available [1]. Inspired by the way plants use sunlight to build complex molecules, the chemist Ciamician already speculated in 1912 that chemicals could be synthesized by utilizing sunlight as an abundant and renewable energy source [2]. While only a vision in the beginning, significant progress has been achieved in the past decade to implement light-driven catalytic processes in sustainable organic synthesis.

Enzymes are natural catalysts in the living world, catalyzing reactions with high chemo-, regio-, and stereoselectivity under mild reaction conditions. Thus, over the last two decades, much effort has been made to further extend the range of biocatalytic reactions available to synthetic chemists, in particular by increasing the available toolbox of biocatalysts, leading to novel catalytic functionalities [3]. Accelerated by the rapid development of genetic engineering tools in combination with protein engineering and advanced screening methods, biocatalytic processes are increasingly implemented as an alternative synthetic strategy in both academia and industry [4], [5], [6]. Here, the intrinsic advantages of enzymes for accessing enantiopure products are especially exploited in the pharmaceutical industries, further emphasizing the powerfulness of biocatalysts for organic synthesis [4]. Among the range of enzyme candidates that are highly desired for industrial implementation, redox enzymes or oxidoreductases have gained significant attention [7]. This is because of their outstanding ability to catalyze complex redox reactions under mild conditions, which cannot be performed by classical organic synthesis approaches or only by using harmful and toxic chemicals.

Photocatalysis, on the other hand, has proven itself as a truly powerful approach to perform reactions in the presence of light that would be very difficult or even impossible to conduct under dark conditions [8]. While earlier ultraviolet (UV) light was predominantly used for the direct activation of organic molecules, modern photochemical activation techniques are predominantly based on selective excitation of photocatalysts with visible light. As such, visible light-mediated activation of organic molecules represents a mild, sustainable and clean method for chemical activation and has been applied to various challenging reactions such as C–H bond activation, C–C bond cross-coupling, C–N bond formation, cycloadditions, and halogenations [9], [10], [11].

While nature invented the intriguing concept of combining enzyme catalysis with photocatalysis, chemists nowadays strive to mimic this approach by coupling the catalytic reactivity and outstanding selectivity of enzymes with photocatalytic reactions. In such a dual system, referred to as ‘photobiocatalysis’ (Figure 1), the exquisite selectivity of enzymes is combined with the unique chemical transformations accessible to photocatalysis. In such a dual system, light is absorbed by a photoactive compound to activate an organic substrate, and the subsequent enzymatic reaction catalyzes a selective transformation, synergistically leading to the synthesis of complex molecules that are difficult to obtain by conventional methods. While early efforts in the field predominantly focused on developing light-driven systems to provide reduced cofactors (NADPH, FAD) or oxidants (e.g. H2O2) to redox enzymes, researchers intensively explore now the possibility of catalyzing non-natural reactions via photoinduced enzyme promiscuity, and the combination of photocatalysis and biocatalysis that can be applied to create novel synthetic methodologies.

Figure 1: 
Photobiocatalysis, possessing advantages of the reactivity of photocatalysts and the selectivity of enzymes, is increasingly realized for the green synthesis of value-added chemicals. Adapted with permission from Ref. [12]. Copyright 2022 American Chemical Society.
Figure 1:

Photobiocatalysis, possessing advantages of the reactivity of photocatalysts and the selectivity of enzymes, is increasingly realized for the green synthesis of value-added chemicals. Adapted with permission from Ref. [12]. Copyright 2022 American Chemical Society.

This book chapter provides an overview of the recent progress and application of light-driven bioprocesses (photobiocatalysis) in organic synthesis, ranging from photobiocatalytic cofactor regeneration, natural and artificial photoenzymes, non-natural reactions via photoinduced enzyme promiscuity, and cascades involving photobiocatalytic transformations.

2 Basic principles and terms in light-driven bioprocesses

Photobiocatalysis adopts natural or artificial light for exploiting the advantages of photocatalyst reactivity and enzyme selectivity (Figure 1) [13], [14], [15], [16], [17]. In the search for efficient strategies that exploit the potential of light in biocatalysis, promising basic concepts have emerged that are fundamental to the meaningful implementation of both whole-cell and in vitro photobiocatalytic applications [15, 18, 19]. These concepts can be divided according to the following classification: (1) photocatalytic regeneration cascades, (2) photoenzymatic reactions utilizing natural or artificial photoenzymes or photoinduced enzyme promiscuity and (3) photoenzymatic systems (Figure 2) [20]. In photocatalytic regeneration cascades, light can be used to generate photoexcited electrons to fuel enzyme-catalyzed redox reactions, i.e. photocatalytic regeneration of redox enzymes (in vitro and in vivo, Figure 2A and B) [21]. In photoenzymatic reactions, natural or artificial cofactor-dependent photoenzymes are applied that directly need light to perform their catalytic reaction (Figure 2C) [12]. On the other hand, photoenzymatic systems refer to the combination of photo(organo)catalytic reactions that use light to directly drive small molecule conversions in combination further enzymatic functionalization steps in one-pot (tandem) photobiocatalytic reactions (Figure 2D) [16].

Figure 2: 
The main design principles in photobiocatalysis. A) Photocatalytic regeneration cascades via direct or indirect activation of redox enzymes (in vitro); B) light-driven cofactor supply using photoautotrophs (in vivo); C); natural and artificial photoenzymes for (non-natural) photo-biocatalytic reactions; D) photoenzymatic cascades combining photo- and biocatalytic transformations. Adapted with permission from Refs. [12, 14, 16]. Copyright 2022 American Chemical Society.
Figure 2:

The main design principles in photobiocatalysis. A) Photocatalytic regeneration cascades via direct or indirect activation of redox enzymes (in vitro); B) light-driven cofactor supply using photoautotrophs (in vivo); C); natural and artificial photoenzymes for (non-natural) photo-biocatalytic reactions; D) photoenzymatic cascades combining photo- and biocatalytic transformations. Adapted with permission from Refs. [12, 14, 16]. Copyright 2022 American Chemical Society.

Photobiocatalytic approaches aiming at the direct or indirect photoactivation of redox enzymes gained increasing attention in the past decade, in particular to tackle common bottlenecks related to the application of oxidoreductases in organic synthesis. Independent on the reaction system i.e. in vitro, or whole cells, light can be used to generate photoexcited electrons to fuel enzyme-catalyzed redox reactions. Dependent on the reaction setup, these electrons can be subsequently utilized for regenerating or even circumventing the need for nicotinamide cofactors. The basic concept behind this light-induced electron transfer relies on a photochemical effect defined as charge separation (Box 1).

In such light-induced electron transfer processes, the difference between the redox potentials between an electron donor and an oxidized photoactive compound is crucial. In the context of photobiocatalysis, these photoactive compounds are commonly defined as photosensitizers and determine the feasibility of a light-driven reaction system. Photosensitizers can absorb light at specific wavelengths in their excited states [22]. The non-absorbed visible light portion is reflected, which we discern as colors. The reason plants or algae appear green is because they contain chlorophyll. For instance, chlorophyll b within the P680 reaction center of photosystem II (PSII) shows absorption maxima at around 460 and 650 nm, which overlaps with the visible spectrum of blue and red light, respectively [23]. Light-irradiation of P680 and subsequent charge separation yield in P680+ and the reduction of pheophytin functioning as an electron acceptor. The formed P680●+ proton radical acts as a strong oxidant with a redox potential of about 1.2–1.4 V [24]. Emphasized as the strongest oxidant generated in a natural reaction [25], P680●+ can initiate the oxidation of H2O via the oxygen-evolving complex [26]. Via multiple transition states, four electrons and protons are subsequently subtracted from two water molecules releasing O2 and four protons, a reaction defined as water splitting. Analog to the described light-driven water-splitting reaction in PSII, current photobiocatalytic approaches often rely on the exploitation of the photo-excitability of a photosensitizer to realize light-induced electron transfers. Especially, organic molecules such as dyes or pigments with delocalized conjugated π-systems have revealed suitable charge separation properties in various photobiocatalytic applications [27, 28]. Their redox potentials during photoexcitation and charge separation determine their compatibility within the electron donor/acceptor framework.

The overall aim of implementing photosensitizers in enzyme-catalyzed reaction systems is either the injection or withdrawal of photoexcited electrons. In that sense, photosensitizers act as ‘converters’ of light into redox potential. This change in redox potential is the fundamental driving force for the transfer of electrons from an electron donor to an electron acceptor. As indicated, this mechanism follows a linear reaction mode rather than a closed reaction cycle. Thus, the supply of electrons must be maintained over time to avoid implementing an undesired bottleneck or rate-limiting step within the reaction system. As the initial source of electrons is continuously consumed, it is conventional to indicate its consumption using the term ‘sacrificial’ electron donor. Compounds exhibiting low redox potentials at operational pH, including organic acids (formic acid or ascorbic acid) or tertiary amines (triethanolamine (TEOA), 2-ethanesulfonic acid (MES), 3-(N-morpholino)propanesulfonic acid (MOPS) or ethylenediaminetetraacetic acid (EDTA)), are typically considered as suitable sacrificial electron donors [15, 29]. However, choosing a suitable sacrificial electron donor/photosensitizer combination for a desired light-driven reaction setup is often based on the preceded empirical evaluation and optimization studies [27].

BOX 1: Charge separation.

Photoinduced charge separation describes the light-driven transfer of electrons between donor and acceptor molecules resulting in the formation of respective charge transfer states. The driving force behind this mechanism is the excitability of electrons in a ground state to higher energy levels upon light irradiation. In homogeneous photocatalysis, the electrons within the Highest Occupied Molecular Orbital (HOMO) of the photocatalyst undergo excitation to the Lowest Unoccupied Molecular Orbital (LUMO). In the presence of an acceptor molecule, the excited electron can be subsequently mediated which results in photoinduced charge separation:



Light irradiation of the photocatalyst (PC) results in its excited state (PC*). Thereby an electron from the Highest Occupied Molecular Orbital (HOMO) is excited to a higher energy level, the Lowest Unoccupied Molecular Orbital (LUMO). The excited electron can be subsequently mediated to the LUMO of an electron acceptor (A) which results in photoinduced charge separation and the formation of PC+ and A.

In heterogeneous catalysis, electrons in the valence band can be exited into the conduction band upon light irradiation. Unlike in homogeneous catalysis, photoinduced charge separation occurs within the semiconductor material without the need for an acceptor or donor molecule. The charge separation results in the formation of negative charges electrons and positively charged holes within the semiconductor. Mediated to the surface of the photocatalyst, positively charged holes or electrons can participate in oxidation and reduction reactions, respectively.

3 Photocatalytic regeneration cascades via direct or indirect activation of redox enzymes

The reactivity of most oxidoreductases (EC1, over 80 %) involves the transfer of electrons between different atoms or molecules, while during the catalyzed redox reaction one compound is being oxidized, another one is being reduced. Due to that, oxidoreductases require additional electron donor or acceptor molecules that have to be supplied in stoichiometric amounts. Typical electron donors used by oxidoreductases are the nicotinamide cofactors NADH and NADPH. Since NAD(P)H plays a central role in biocatalytic processes utilizing oxidoreductases, researchers have developed various in situ regeneration strategies, allowing the use of the costly cofactor in catalytic amounts (Box 2). Overall, this contributes to a significant cost reduction of a biocatalytic process [30]. While enzymatic regeneration systems nowadays prevail in synthetic applications [31], several photochemical regeneration systems have been developed in the past decade. Reductive regeneration of redox enzymes can be achieved either directly, i.e. by direct reduction of the active sites of the enzymes, or indirectly, i.e. involving the nicotinamide cofactors (Scheme 1).

BOX 2: Conventional cofactor regeneration in biotechnology.

Conventional in vitro applications aiming at the regeneration of cofactors commonly rely on the coupled substrate approach [134], or the in situ regeneration via enzyme coupling [30]. In the substrate-coupled approach, the enzyme of interest accepts the desired substrate as well as a cosubstrate. As applied in the reduction of ketones catalyzed by alcohol dehydrogenases, isopropanol can be used as a cosubstrate for the oxidative back reaction generating reduced nicotinamide cofactor and acetone as a by-product [134]. By following the enzyme-coupled approach, the reaction system is extended by an additional enzymatic conversion, which catalyzes the in situ cofactor regeneration. For instance, in the enzyme-catalyzed reduction of ketones, NAD(P)H could be successfully regenerated by using glucose dehydrogenase or formate dehydrogenase in combination with glucose or formic acid as cosubstrates [135], [136], [137].
Scheme 1: 
Artificial light-induced electron transfer for the direct or indirect photoactivation of oxidoreductases. Adapted with permission from Ref. [16].
Scheme 1:

Artificial light-induced electron transfer for the direct or indirect photoactivation of oxidoreductases. Adapted with permission from Ref. [16].

3.1 Indirect photoactivation of redox enzymes

In photobiocatalysis, several concepts have been reported that demonstrate the feasibility of reducing NAD(P)+ by exploiting the reduction potential of photosensitizers in their excited state [32]. Analogous to oxygenic photosynthesis, light is converted into chemical energy stored in NAD(P)H. Unlike the oxidation of NAD(P)H to NAD(P)+, the back-reaction to regenerate reduced nicotinamide has to be catalyzed regioselectively, as the direct reduction of NAD(P)+ can yield inactive isomers or undesired dimerization [3334]. Other studies suggest the use of additional organometallic compounds as electron mediators, which facilitate the formation of the active 1,4-NAD(P)H, subsequently consumed in enzymatic reactions [35, 36]. Therein, RhIII and IrIII complexes turned out to be highly potent in performing the mediation of electrons between an electrode and the cofactor, while a hydride ion (H) serves as a reducing equivalent in the selective reduction of NAD(P)+ [36]. Based on these observations, indirect photochemical regeneration strategies of reduced NAD(P) have been developed (Scheme 2A), wherein the photochemical nicotinamide cofactor recycling using electron mediators is for instance coupled to the activity of cytochrome P450s [37] or dehydrogenases [28].

Scheme 2: 
General photobiocatalytic concept for indirect activation of redox enzymes. A) General scheme of the indirect photoactivation redox enzymes. B) A thermophilic ene-reductase (TOYE) catalyzes the enzymatic reduction of an α,β-unsaturated substrate. The reaction is driven by photoexcited electrons deriving from the sacrificial electron donor triethanolamine (TEA). [Ru/bpz)3]2+ is used as photosensitizer and methyl viologen (MV2+) as mediator [39]. C) Light-driven activation of cytochrome P450 BM3 catalyzing the O-dealkylation of 7-ethoxycoumarin [37]. Eosin Y is used as photosensitizer transferring electrons from the sacrificial electron donor TEOA to an organometallic mediator for regenerating NADPH. Adapted with permission from Ref. [16].
Scheme 2:

General photobiocatalytic concept for indirect activation of redox enzymes. A) General scheme of the indirect photoactivation redox enzymes. B) A thermophilic ene-reductase (TOYE) catalyzes the enzymatic reduction of an α,β-unsaturated substrate. The reaction is driven by photoexcited electrons deriving from the sacrificial electron donor triethanolamine (TEA). [Ru/bpz)3]2+ is used as photosensitizer and methyl viologen (MV2+) as mediator [39]. C) Light-driven activation of cytochrome P450 BM3 catalyzing the O-dealkylation of 7-ethoxycoumarin [37]. Eosin Y is used as photosensitizer transferring electrons from the sacrificial electron donor TEOA to an organometallic mediator for regenerating NADPH. Adapted with permission from Ref. [16].

Early studies on photochemical NAD(P)H regeneration systems rely on the use of Ru[(bpy)3]2+ or ZnTMPyP4 as photosensitizers in combination with methyl viologen (MV) as a primary electron acceptor [38]. In the respective photoinduced charge separation at the photosensitizer, electrons are initially transferred to MV2+, using either EDTA, 2-mercaptoethanol, triethylamine (TEA) or TEOA as sacrificial electron donors. The generated MV+● radical acts as the electron donor for the subsequent enzymatic reduction of NAD(P)+ catalyzed by an NAD(P)+-dependent dehydrogenase (Scheme 2B) [39]. By coupling the light-driven cofactor regeneration system to a second enzymatic conversion, CO2 conversion into keto acids [40], reduction of ketones [41], and reduction and amination of keto acids was demonstrated [38, 42]. However, the need for an additional biocatalyst to catalyze the reduction of NAD(P)+ does not offer a significant advantage over conventional enzymatic cofactor regeneration systems [31].

In searching for more safe and straightforward solutions for the photochemical regeneration of NAD(P)H, reaction systems implementing organic dyes [43], flavins [44], quantum dots [45], graphene-based photocatalysts [46] and porphyrins [47] as photosensitizers in combination with [Cp*Rh(bpy)H2O]2+ as electron mediator were developed. [Cp*Rh(bpy)H2O]2+ represents a versatile organometallic mediator exhibiting high activities and stability for the highly selective for generating the active 1,4-cofactor variants [15, 19, 44]. For example, Lee et al. used a light-driven approach to catalyze the O-dealkylation of 7-ethoxycoumarin with cytochrome P450 BM3 (Scheme 2C) [37]. The xanthene dye eosin Y acts as a photosensitizer to transfer photoexcited electrons from TEOA to an organometallic mediator molecule for NADPH regeneration. Lee et al. also proposed the light-driven regeneration of NADH by using the sacrificial electron donor TEOA and eosin Y as photosensitizer [28]. Eosin Y, an organic dye previously used for cell staining in histological applications, revealed its potential as a photosensitizer in various photochemical applications [48]. In a photobiocatalytic setup for the conversion of α-ketoglutarate to L-glutamate catalyzed by L-glutamate dehydrogenase, [Cp*Rh(bpy)H2O]2+ was used as an electron mediator between photoexcited eosin Y and NAD+ yielding NADH [28]. Thereby, a turnover number of about 25 NAD(P)+ h−1 was estimated for the photosensitizer-mediator driven cofactor regeneration system, yielding about 90 % substrate conversion. Driven by these findings, the scope of potential photosensitizers was further explored by screening different xanthene dyes [49]. As a result, rose bengal, fluorescein, erythrosine B and phloxine B show similar or even better performance in the light-driven cofactor supply for the GDH compared to the previously reported eosin Y.

Other examples report the use of graphene-based photosensitizers harboring covalently bound porphyrin derivates [46], or difluoro-boraindacene (BODIPY) [50], as chromophores that can donate electrons to [Cp*Rh(bpy)H2O]2+ upon visible light exposure. Analogous to previously described examples, [Cp*Rh(bpy)H2O]2+ catalyzes the reduction of NAD+ to NADH, the final step in the photocatalytic cofactor recycling mechanism. Also quantum dots have been used as photosensitizers in photochemical NADH cofactor regeneration systems with [Cp*Rh(bpy)H2O]2+ [51].

The previous examples emphasize the feasibility of light-driven NAD(P)H regeneration combined with the enzymatic reduction of organic compounds. Similarly, light-induced charge separation and electron transfer can be utilized for the in situ generation of oxidized NAD(P)+. In this case, NAD(P)H generated during enzyme-catalyzed substrate oxidation serves as an initial electron donor quenching the oxidized photosensitizer. Consequently, an electron acceptor, often O2, represents the final component in the light-driven electron transfer chain. Gargiulo et al. successfully coupled the photochemical generation of NAD(P)+ to the oxidation of alcohols into ketones and lactones catalyzed by alcohol dehydrogenase [52]. In this case, flavins were applied as photosensitizers for the light-driven reoxidation of NAD(P)H to NAD(P)+. Since oxygen serves as the terminal electron acceptor for the regeneration of the reduced flavin, the strong oxidant H2O2 is formed as a byproduct. While Gargiulo et al. used catalase to avoid possible side effects caused by H2O2, the use of dithionite as a photocatalytic component in the regeneration of NADP+ yields in the final formation of H2O as proposed by Rickus et al. [53]. The generation of NADP+ was determined by the enzymatic conversion of isocitrate to α-ketoglutarate. The reaction catalyzed by an NADP+-specific isocitrate dehydrogenase revealed turnover frequencies of about 35 h−1 by providing the photosensitizer in solution.

3.2 Direct photoactivation of redox enzymes

Reaction systems with the purpose of direct photoactivation of oxidoreductases have the following aims: activating cofactor-independent enzymes, circumventing the need for cofactors and decreasing the complexity of multicomponent reaction systems (Scheme 3A). Therefore, the focus is on the redox-active prosthetic groups, including flavins, heme moieties, and non-heme bond metal ions within the polypeptide framework of oxidoreductases. In nature, these enzymes receive electrons directly from reduced cofactors or indirectly via redox partners. These mechanisms provide a site for injecting photoexcited electrons resulting in light-driven enzyme activation.

Scheme 3: 
Photobiocatalytic concepts for the direct activation of redox enzymes. A) General scheme for the direct photoactivation of redox enzymes via electron transfer from a sacrificial electron donor. B) In vitro direct photoactivation of an OYE variant from Thermus scotoductus (TsOYE) catalyzing the asymmetric reduction of C=C bonds [54]. Rose bengal as photosensitizer transfers photoexcited electrons to the prosthetic FMN group of TsOYE. C) The combination of a photosensitizer (e.g. 5(6)-carboxyeosin) with sacrificial electron donors (e.g. MES buffer) fuels Rieske oxygenase (RO)-catalyzed hydroxylations in whole cells [27]. In this example, photoexcited electrons are transferred to the active site (non-heme iron center) of the RO [27]. Adapted with permission from Ref. [16].
Scheme 3:

Photobiocatalytic concepts for the direct activation of redox enzymes. A) General scheme for the direct photoactivation of redox enzymes via electron transfer from a sacrificial electron donor. B) In vitro direct photoactivation of an OYE variant from Thermus scotoductus (TsOYE) catalyzing the asymmetric reduction of C=C bonds [54]. Rose bengal as photosensitizer transfers photoexcited electrons to the prosthetic FMN group of TsOYE. C) The combination of a photosensitizer (e.g. 5(6)-carboxyeosin) with sacrificial electron donors (e.g. MES buffer) fuels Rieske oxygenase (RO)-catalyzed hydroxylations in whole cells [27]. In this example, photoexcited electrons are transferred to the active site (non-heme iron center) of the RO [27]. Adapted with permission from Ref. [16].

In this context, members of the ene-reductase (ERED) enzyme family have been studied for regeneration by chromophores that are being utilized as photosensitizers, which, upon light excitation, can transfer electrons from a sacrificial electron donor such as TEOA to the oxidized FMN within the active site of the EREDs (Scheme 3B) [54, 55]. Instead of indirectly regenerating the ERED via regeneration of reduced nicotinamide cofactors, reducing equivalents from simple sacrificial electron donors such as TEOA, EDTA, formate, or phosphite via were provided via photocatalytic oxidation. The photoenzymatic conversion of, for instance, 2-methylcyclohexenone by TsOYE in combination with rose bengal (RB) resulted in enantiopure (R)-2-methylcyclohexanone (ee >99 %) with a yield of up to 53 % and total turnover numbers of 235 for TsOYE (Scheme 3B) [56]. The substitution of halogen atoms in xanthene dyes significantly affected the turnover frequency.

Also heme-containing enzymes such as cytochrome P540s (CYPs) have been utilized for direct photoactivations. For instance, Park et al. demonstrated the light-induced electron transfer from TEOA to the heme-prosthetic group of various CYPs [57]. Emphasized as a cofactor-independent reaction system, the hydroxylation of various drugs and steroid molecules by using resting Escherichia coli cells containing overexpressed CYP variants and eosin Y as photosensitizer has been shown. In reactions performed for over 20 h under visible light irradiation, total turnover numbers of up to 180 of various CYPs were achieved. Analogously, a study conducted by Özgen et al. indicates the versatility of the described whole-cell approach using dye derivates as organic photosensitizers in combination with EDTA, MES or MOPS buffer as sacrificial electron donors (Scheme 3C) [27]. The asymmetric hydroxylation of olefins catalyzed by four Rieske non-heme iron-dependent oxygenase (RO) variants has been performed to confirm the applicability of the proposed reaction system to further multi-component oxygenases. By that, the costly supply for reduced NAD(P)H cofactors and limitations arising from the assumed instability of ROs in cell-free systems were circumvented.

3.3 Light-driven cofactor supply using photoautotrophs

The application of whole-cells as biocatalysts offers unique advantages and has been widely used in the biosynthesis of high-value fine and bulk chemicals, as well as active pharmaceutical ingredients [58, 59]. The continued identification of new microorganisms as well as recent advances in synthetic biology and metabolic engineering, together with the rapid development of molecular genetic tools, have led to a renaissance of whole-cell biocatalysis.

Besides upregulating the production of endogenous metabolites in industrial-scale fermentation processes [60], recombinant whole cells provide an interesting platform for biotransformations catalyzed by heterologous enzymes. By exploiting the cell’s metabolism, one-pot multi-step conversions of simple substrates into value-added chemicals without additional cofactor supply systems represent an overall goal in industrial biocatalysis [61]. However, the use of heterotrophic microorganisms such as Corynebacterium glutamicum in the industrial production of amino acids, or recombinant E. coli as a model organism for establishing e.g. enzyme cascades, is dependent on the sufficient supply of sugars (e.g. glucose) as carbon fuel for their metabolism. From an environmental perspective, glucose represents a renewable carbon source obtained from biomass and is thus considered a green alternative to petrochemical raw materials for producing fuels, bulk or fine chemicals. From a socio-economical perspective, however, glucose production is mainly based on the enzymatic hydrolysis of edible biomass, which is questioned as it overlaps with the global feedstock and food supply [62]. To counteract this debate, the development of alternative routes to increase the accessibility of fermentable carbon sources from non-edible biomass such as lignocellulose-rich materials has gained increased attention [63].

Unlike heterotrophs as whole-cell platforms for bioconversions, photoautotrophs can utilize CO2 as a sole carbon source to fuel their metabolism. The energy needed to assimilate CO2 is derived from natural light during photosynthesis [36]. The light-induced oxidation of water (referred to as water-splitting) yields hydrogen protons (H+) and electrons, and the light-generated electrons are directly used to reduce NADP+ to NADPH. This makes photosynthetic organisms a particularly interesting chassis for coupling heterologously expressed oxidoreductases to the rich NADPH pool, thereby driving whole-cell light-driven biotransformations [64]. Recently, cyanobacteria have emerged as highly interesting whole-cell biocatalysts in such light-driven redox reactions [64], and have been applied for instance in the asymmetric reduction of ketones [65] or aldehydes [66]. The similarity of cyanobacteria to other Gram-negative bacteria such as E. coli enables the generation of recombinant variants by applying available genetic engineering techniques [67]. By that, biotransformations can be conducted that go beyond the endogeneous reaction scope of cyanobacteria [68]. Considering the enhanced atom economy enabling water as a sacrificial electron donor to fuel enzymatic redox reactions, the use of cyanobacteria appears to have a significant advantage over recombinant heterotrophs.

Following these considerations, Köninger et al. coupled the photosynthetic apparatus of recombinant cyanobacteria to the activity of oxidoreductases (Figure 3A) [69]. Thereby, the asymmetric reduction of activated C=C bonds catalyzed by the NADPH-dependent ERED YqjM from Bacillus subtilis expressed in recombinant Synechocystis sp. PCC 6803 cells was successfully shown. On a semi-preparative scale using 100 mg substrate, an isolated product yield of up to 80 % of enantiopure (R)-2-methylsuccinimide was achieved, which is comparable to the efficiency of typical whole-cell biotransformations in E. coli [39]. Büchsenschütz et al. reported that recombinant Synechocystis cells can also be applied for the asymmetric reduction of cyclic amines [70]. Thereby, reduced NADPH cofactors, directly or indirectly provided by photosynthesis, can fuel the catalytic activity of imine reductases (IREDs) by supplying reducing equivalents. By optimizing reaction conditions, full conversions of prochiral substrates could be achieved, yielding optically pure secondary amines.

Figure 3: 
The cyanobacterium Synechocystis sp. PCC 6803 is used in whole-cell biotransformations for A) the C=C double bond reduction catalyzed by the ene reductase YqjM from Bacillus subtilis [69, 74]. B) The hydroxylation of nonanoic acid methyl ester catalyzed by the monooxygenase AlkB from Pseudomonas putida Gpo1 [71, 72]. Adapted with permission from Refs. [69, 71] .
Figure 3:

The cyanobacterium Synechocystis sp. PCC 6803 is used in whole-cell biotransformations for A) the C=C double bond reduction catalyzed by the ene reductase YqjM from Bacillus subtilis [69, 74]. B) The hydroxylation of nonanoic acid methyl ester catalyzed by the monooxygenase AlkB from Pseudomonas putida Gpo1 [71, 72]. Adapted with permission from Refs. [69, 71] .

Hoschek et al. coupled the formation of O2, driven by the initial water-splitting reaction of oxygenic photosynthesis, to enzyme-catalyzed oxyfunctionalization reactions (Figure 3B) [71]. It is assumed that the light-driven in situ generations of O2 has the potential to counteract mass-transfer limitations between the gas and liquid phase, a crucial factor in the upscaling of O2-dependent processes. In a corresponding study, engineered Synechocystis sp. PCC 6803 was used as a host organism for the heterologous expression of an NADH-dependent alkane monooxygenase (AlkB) from Pseudomonas putida GPo1 [72]. As revealed in whole-cell reactions performed under anaerobic conditions, the hydroxylation of nonanoic acid methyl ester into the corresponding ω-hydroxylated product strictly depends on the light-driven water-oxidation yielding a sufficient amount of intracellular O2. Driven by these findings, the same research group evaluated the scalability of hydroxylation reactions without requiring external aeration [73]. The in vivo conversion of cyclohexane into the corresponding mono-hydroxylated alcohol catalyzed by a CYP from Acidovorax sp. in a two-liquid phase setup resulted in 2.6 g of cyclohexanol after 52 h.

3.4 Photocatalytic generation of hydrogen peroxide in situ

While the obtained turnover numbers for photobiocatalytic cofactor regeneration (e.g. NADPH, FAD) are in many cases too low for synthetic applications, the developed examples of combining a photosensitizer with a sacrificial electron donor can also be used for the in situ generation of H2O2. This is particularly interesting as many oxidoreductases such as CYPs and ROs rely on a complex electron-transfer chain and are cofactor-dependent, limiting their application in organic synthesis. In contrast, peroxygenases use H2O2 as oxidant, however they usually show poor robustness against high H2O2 concentrations. As such, the concentration of H2O2 to be carefully controlled to avoid rapid deactivation of the enzymes while maintaining high reactivity. In this context, several photocatalytic strategies for in situ H2O2 generation have been developed. For instance, Zhang et al. used gold–titanium dioxide (Au–TiO2) to generate H2O2 through the methanol-driven reductive activation of ambient O2 to drive an unspecific peroxygenase from Agrocybe aegerita (rAaeUPO) [75]. Using this approach, the stereoselective hydroxylation of ethylbenzene to (R)-1-phenylethanol was achieved with high enantioselectivity (>98 % ee) and high turnover numbers for the biocatalyst (>71,000). This reaction system was further explored by using water instead of methanol as electron donor to produce H2O2 in situ to drive rAaeUPO [76]. By using this system, the stereoselective hydroxylation of ethylbenzene to (R)-1-phenylethanol has also been achieved, resulting in 110 mg product with an ee of >97 %. This system was later on further optimized.

In addition to using TiO2-based photocatalysts or even flavin [77] for the in situ H2O2 generation, further photocatalysts have been explored for the same purpose and enzyme class. In this regard, a water-soluble sodium anthraquinone sulfonate (SAS) was used to generate H2O2 to drive a vanadium-dependent chloroperoxidase from Curvularia inaequalis (CiV-CPO)-catalyzed halogenation reaction [78]. With this system, a yield of 91 % and turnover number of 318,000 was achieved. Interestingly, this approach eliminates the diffusion limitations of the heterogeneous photocatalyst and protects the enzyme from long-lived radicals. Other examples using nitrogen-doped carbon nanodots [79, 80] also highlight that these photocatalysts are promising to drive peroxygenase-catalyzed hydroxylation reactions. In particular, it was shown that the spatial separation of the photocatalyst from the enzyme avoids the inactivation of the enzyme, resulting in promising turnover numbers of the biocatalyst of more than 60,000 [79]. In addition, these nitrogen-doped carbon nanodots can also be applied in neat reaction media. For instance, it was shown that the immobilization of rAaeUPO increased enzyme stability and thus facilitated the reaction in neat substrate such as cyclohexane [80].

4 Natural and artificial photoenzymes for (non-natural) photo-biocatalytic reactions

While the last decade has seen a strong focus on the use of light-regenerated cofactors to achieve native enzymatic activity, recent developments indicate that the combination of biocatalysis and photocatalysis is even more powerful and can unlock non-natural enzyme reactivities [12]. In particular, the discovery and application of natural and artificial photoenzymes capable of directly converting light into chemical energy are well placed to further expand the applications of photobiocatalysis.

4.1 Natural photoenzymes

Yet, the scope of photoenzymes discovered in nature is very limited and the investigation of their biocatalytic applicability is often restricted to their natural substrate and reaction scope. One of the first natural photoenzymes was discovered in the 1950s by the physicist Claud S. Rupert [81]. This enzyme, a so-called photolyase, is a flavoprotein containing two light-harvesting cofactors, that is responsible for repairing UV light induced DNA damage (Scheme 4A) [82]. Although photolyases can catalyze DNA photoreactivation with considerable high efficiency, their implementation in biocatalytic applications has not yet been described in the literature [83].

Another class of natural photoenzymes are the light-dependent protochlorophyllide reductases (LPORs, Scheme 4B). These enzymes are involved in the biosynthesis of chlorophyll in both oxygenic and anoxygenic phototrophs, in which they catalyze the reduction of protochlorophyllide (pchlide) to chlorophyllide (chlide) [84]. Yet, the biocatalytic application of LPORs is limited, although it could be shown that in vitro they catalyze the conversion of various pchlide derivatives to the corresponding reduced products [85]. In addition, several LPOR homologs were successfully identified from different origins and biochemically characterized, and their activity was studied for the conversion of pchlide to chlide under different conditions and furthermore analyzed for their cofactor flexibility [86].

In 2017, Beisson et al. identified another natural photoenzyme from the microalgae Chlorella variabilis [87], and ever since its discovery, it is intensively studied for its biocatalytic potential. This enzyme is a fatty acid decarboxylase (FAP) catalyzing a light-mediated decarboxylation of fatty acids via a flavin cofactor (Scheme 4C). Upon light irradiation, the flavin cofactor mediates the first single electron transfer from the enzyme-bound carboxylic acid, initiating the sequence of CO2 extrusion and back-transfer of the initially abstracted electron to the newly formed C-centered radical [87, 88]. Under light illumination at around 450 nm (blue light), CvFAP efficiently converts long-chain fatty acids into the corresponding alkanes, while achieving high turnover numbers and almost full conversion [89]. An ever growing number of examples highlight the potential of this photoenzyme (and its engineered variants) for various biocatalytic applications, particularly for the synthesis of chiral compounds via kinetic resolution of racemic hydroxyl carboxylic acids [90], or implementation in cascade approaches for the synthesis of secondary alcohols [91] or polymer building blocks [92] from unsaturated fatty acids.

Scheme 4: 
Overview on natural photoenzymes. A) Photoreactivation of pyrimidine dimers in UV-damage DNA catalyzed by a photolyase [82]. B) Light-dependent protochlorophyllide reductase (LPOR)-catalyzed reduction of protochlorophyllide to chlorophyllide [84]. C) The fatty acid photodecarboxylase from Chlorella variabilis (CvFAP) catalyzes a decarboxylation reaction upon light illumination, resulting in the formation of hydrocarbons [87]. Adapted with permission from Refs. [12, 16]. Copyright 2022 American Chemical Society.
Scheme 4:

Overview on natural photoenzymes. A) Photoreactivation of pyrimidine dimers in UV-damage DNA catalyzed by a photolyase [82]. B) Light-dependent protochlorophyllide reductase (LPOR)-catalyzed reduction of protochlorophyllide to chlorophyllide [84]. C) The fatty acid photodecarboxylase from Chlorella variabilis (CvFAP) catalyzes a decarboxylation reaction upon light illumination, resulting in the formation of hydrocarbons [87]. Adapted with permission from Refs. [12, 16]. Copyright 2022 American Chemical Society.

4.2 Artificial photoenzymes

While the chemistry that can be unlocked from natural photoenzymes is still limited, the combination of biocatalysis and photocatalysis opens up a golden window of opportunities to unlock abiological transformations. Although the concept of using light to harness abiological reaction chemistries from an existing and often engineered enzyme scaffold is a rather new development in photobiocatalysis, a handful of examples already highlight the powerfulness of this approach. The construction of these so-called artificial photoenzymes typically exploits two strategies: (1) the covalent linking of a photocatalysts via an (unnatural) amino acid and using the photocatalyst for direct single-electron transfer; and (2) the conjugation of a photosensitizer nearby an immobilized organometallic complex within a protein scaffold [10, 12, 22]. While artificial enzymes also have been constructed to supply electrons to naturally occurring enzyme cofactors to increase native activity [13, 15, 93], we herein focus on artificial enzymes that have been constructed for non-natural reactivities. Already in 2015, Lewis et al. used a click chemistry approach to conjugate a modified acridinium 9-mesityl-10-methyl (Acr+-Mes) cofactor to an unnatural amino acid within a prolyl oligopeptidase to construct an artificial photoenzyme for the sulfoxidation of thioanisoles [94, 95].

Wang et al. constructed an artificial photoenzyme by conjugating a catalytically active organometallic complex in close vicinity to a genetically encoded photoactive chromophore in the superfolder yellow fluorescent protein scaffold [96]. They genetically encoded benzophenone-alanine into the protein scaffold and conjugated a nickel-terpyridine complex for the photocatalytic reduction of CO2. With the help of protein engineering, the positions of the chromophore and catalyst were tuned, the photosensitizers’ photochemical properties were modulated, and the microchemical environment of the protein was adjusted to enable efficient CO2 reduction [97]. This photoactive protein scaffold was further modified to enable accomplish cross-couplings of aryl halides via genetically encoded benzophenone chromophore and an adjacent artificial NiII(bpy) cofactor [98].

The incorporation of a photosensitizer into a protein scaffold via genetic code expansion for enantioselective [2+2]-cycloadditions via triplet energy transfer has been explored by the groups of Green as well as Chen, Zhong, and Wu [99, 100]. Both groups developed the artificial photoenzyme for [2+2]-cycloadditions based on a different protein scaffold, however, both artificial photobiocatalysts were constructed by genetic code expansion to incorporate a photosensitizer. Upon light irradiation of the photosensitizer, a [2+2]-cycloaddition is promoted. The chosen protein scaffold thereby delivers the needed selectivity, which was further optimized via protein engineering. For instance, the obtained first generation photoenzyme developed by Trimble et al. catalyzed the intramolecular [2+2]-cycloaddition of 4-(but-3-en-1-yloxy)quinolin-2(1H)-one to two regioisomeric products (straight and crossed, Figure 4) [99]. Via directed evolution, an efficient and enantioselective enzyme (up to 99 % ee) promoting selective cycloadditions with >300 turnovers under aerobic conditions was obtained.

Figure 4: 
Design of an artificial photenzyme promoting enantioselective [2+2]-cycloadditions via triplet energy transfer (EnT) [99]. The photosensitizer 4-benzoylphenylalanine (BpA) is genetically encoded into the protein scaffold of a computationally designed Diels Alderase, followed by directed evolution, affording an enantioselective photoenzyme for the conversion of 4-(but-3-en-1-yloxy)quinolin-2(1H)-one to the straight or crossed product. Adapted from Ref. [99].
Figure 4:

Design of an artificial photenzyme promoting enantioselective [2+2]-cycloadditions via triplet energy transfer (EnT) [99]. The photosensitizer 4-benzoylphenylalanine (BpA) is genetically encoded into the protein scaffold of a computationally designed Diels Alderase, followed by directed evolution, affording an enantioselective photoenzyme for the conversion of 4-(but-3-en-1-yloxy)quinolin-2(1H)-one to the straight or crossed product. Adapted from Ref. [99].

These examples highlight the power of combining photocatalysis with biocatalysis by introducing photocatalysts via bioconjugation with unnatural amino acids or genetically encoding a photosensitizer into a protein scaffold. As such, novel chemical reactivities can be imparted to enzymes, while the design parameters of such biohybrid systems can be modulated to achieve more efficient chemical conversions. However, difficulties arising in the design and construction of artificial photoenzymes impede the chemical diversity that is yet accessible by this approach [12].

4.3 Photoinduced enzyme promiscuity

The protein machineries of living systems are often ‘promiscuous’ – that is, capable of catalyzing reactions other than its biological function. These promiscuous functions can be used to generate catalytic novelty, and promiscuous activities can even be induced or significantly improved by protein engineering [101], [102], [103], [104], [105], [106], [107]. Besides tailoring the substrate specificity or enzyme stability, changing the reaction mechanism of an enzyme is an integral approach that provides access to biocatalysts with a non-biological reaction scope [108, 109].

Recent studies have reported that light can induce catalytic promiscuity in oxidoreductases, which is driven by the excitation of flavin or nicotinamide cofactors as well as organic photosensitizers, leading to the generation of radical intermediates within the enzyme active sites, resulting in various biological transformations. This approach holds great promise for extending the synthetic capabilities of biocatalysts and, when combined with protein engineering, for addressing long-standing selectivity and reactivity challenges in chemical synthesis. This is particularly important as the stereochemical outcome of reactions involving radical intermediates is often difficult to control using existing small molecule catalysts [110].

Recently, Emmanuel et al. have shown that a ketoreductase (KRED) can catalyze the enantioselective dehalogenation of halolactones, a very different reaction from that they were evolved for (Scheme 5A) [111]. In this example, the authors use light to excite the cofactor NADPH bound to the active site of the KRED. Upon light irradiation, the photoexcited cofactor and halolactone substrate generates a charge-transfer complex, leading to the formation of the corresponding cofactor and substrate radicals. The KRED-catalyzed removal of a halogen atom from the halolactone results then in the formation of the dehalogenated chiral lactone product [111].

Scheme 5: 
Overview of different light-driven biocatalytic approaches that use light to induce promiscuity in NAD(P)H-dependent enzymes and flavoenzymes. A) Upon light irradiation, a ketoreductase (KRED) catalyzes an enantioselective dehalogenation resulting in the formation of chiral lactones [111]. B) A photoexcited ene-reductase (ERED) catalyzes the radical cyclization of α-chloroamides into γ-lactams [112]. C) Double-bond reductases catalyzing the deacetoxylation of α-acetoxyketones using photoexcited Rose Bengal (RB) [113]. Adapted with permission from Ref. [16].
Scheme 5:

Overview of different light-driven biocatalytic approaches that use light to induce promiscuity in NAD(P)H-dependent enzymes and flavoenzymes. A) Upon light irradiation, a ketoreductase (KRED) catalyzes an enantioselective dehalogenation resulting in the formation of chiral lactones [111]. B) A photoexcited ene-reductase (ERED) catalyzes the radical cyclization of α-chloroamides into γ-lactams [112]. C) Double-bond reductases catalyzing the deacetoxylation of α-acetoxyketones using photoexcited Rose Bengal (RB) [113]. Adapted with permission from Ref. [16].

In follow-up studies, this concept was transferred to flavin-dependent ene-reductases (Scheme 5B) [112]. While intrinsically, these NAD(P)H-dependent enzymes catalyze the two-electron reduction of activated alkenes, the light-induced radical species formation enabled a hydroalkylation used as driving force for non-natural intramolecular cyclization [112], or an intermolecular radical hydroalkylation of terminal alkenes [114]. Unlike to the examples with KREDs, in which NAD(P)H served as the photoactive component, the light-induced single-electron transfer responsible for the promiscuity of ene-reductases is presumably driven by the photoexcitation of the electron-donor complex composed of reduced FMN- and substrate.

These examples rely on the photoinduced electron transfer from NADPH or ground state electron transfer from a flavin hydroquinone. Another approach is to use an exogenous reductant to facilitate electron transfer, while a protein scaffold serves as the chiral catalyst (Scheme 5C) [110, 113]. However, this strategy requires the development of gating strategies to ensure that the formation of the radical species occurs exclusively within the active site of the enzyme. This hypothesis was investigated on a reductive radical deacetoxylation reaction with a NADPH-dependent double bond reductase (DBR) as an enzymatic scaffold. Indeed, the deacetoxylation was observed with good levels of enantioselectivity upon addition of commonly used transition-metal and organic photocatalysts. Moreover, it was proposed that green-light irradiation of RB in the presence of NAD(P)H forms radical RB●−, an intermediate capable of reducing the enzyme-bond substrate yielding in the formation of a deacetoxylated α-acyl radical. The hydrogen transfer from reduced NADPH yields then the respective product in the final step [113].

5 Photoenzymatic cascades combining photo- and biocatalytic transformations

Next to photobiocatalytic applications that are designed to provide redox enzymes with reducing equivalents for catalysis, the combination of photo- with biocatalytic reactions steps is an emerging field of interest. These so-called photochemoenzymatic (PCE) cascades use light to directly drive small molecule interconversions and combine them with further enzymatic functionalization steps. Cascade reactions (Box 3) in general are of growing importance in biotechnological applications, thus it is not surprising that the development of cascades implementing photocatalytic reaction steps is also getting momentum. The increasing number of examples that have been reported recently further highlight these developments, particularly because the use of light can significantly extend the available reaction chemistries to synthesize various pharmaceuticals and fine chemicals [16].

While one-pot linear cascades (referred to as concurrent or tandem reactions) are much more preferable as no intermediate work-up is required [115], the possibility of performing a PCE cascade in a simultaneous reaction mode typically requires a careful evaluation of photo- and biocatalyst compatibility. Thereby, the type of photocatalyst, possible side reactions and the stability of the enzyme in the presence of the photocatalyst are crucial factors determining the efficiency of the simultaneous reaction mode [16]. Here, the combination of a photocatalyst with an enzyme can allow the design of powerful reactions that exploit the unique reactivity of photocatalysts and the effectiveness of enzymes [116]. However, critical challenges need to be overcome in order to operate such systems with high efficiency. Below are some examples of PCE cascades and the approaches used to overcome incompatibility issues.

BOX 3: Cascade biocatalysis – classifications.

As biocatalytic cascade, we define a reaction system that combines two or more chemical steps in one reaction vessel without isolation of intermediates. Thereby, enzymatic, chemical (metal or organo) or spontaneous catalytic reaction steps can occur. Moreover, these multi-step reactions can be either performed simultaneously or sequentially. Simultaneous cascades are also known as concurrent cascades or tandem reactions. In this case, all reaction steps take place at the same time, meaning that all enzymes and reagents are present from the beginning in the reaction vessel. In case of sequential cascades, the catalysis usually proceeds step-wise, while additional enzymes/reagents are added once another reaction step has been completed. According to Schrittwieser et al., cascades are generally classified according to their topology: (1) in a linear cascade the product of one chemical step serves as the substrate of the subsequent chemical step; (2) in an orthogonal cascade the conversion of a substrate to the product is coupled with a second reaction to remove one or more by-products; (3) in a cyclic cascade one enantiomer of a racemic substrate mixture is converted to an intermediate product which is then converted back to the racemic starting material, leaving the unreacted substrate enantiomer as the final product; and (4) a parallel linked cascade in which two separate biocatalytic reactions are linked by complementary cofactor requirements of the two enzymes [115].

5.1 Simultaneous and sequential photochemoenzymatic cascades

The realization of PCE cascades in simultaneous mode is often difficult, mainly because of the often opposite reaction conditions required by the photocatalyst and the enzyme. In addition, stability problems can arise due to the presence of the enzyme, substrate and product in the reaction mixture. Hartwig and colleagues reported a concurrent PCE cascade in which a photocatalyst is used for the isomerization of alkenes, which are further converted by an ERED to generate valuable enantio-enriched products (Scheme 6) [116]. Initially, a number of organometallic and organic photocatalysts were screened for their ability to isomerize the (Z)-configured substrates, and different photocatalysts such as FMN and Ir(III) were identified as most suitable. In the following, the enzymatic reduction of the (Z)-configured substrates was investigated. The cooperative reduction catalyzed by different EREDs and by using different photocatalysts such as FMN and Ir(III) under blue light illumination resulted in the formation of several products with high yields (up to 87 %) and ee’s of >99 %. In this example, two features of the photocatalyst were crucial for performing the cascade in a concurrent reaction mode: (1) the photocatalytic reaction takes place at room temperature to match the enzyme requirements, and (2) the mechanism underlying the photocatalytic reaction involves intermediates that are stable towards water and the functional group in proteins.

Scheme 6: 
Concurrent photochemoenzymatic cascade comprising a photocatalytic isomerization with the enzymatic reduction of alkenes to the corresponding enantio-enriched products under blue light illumination [116]. Adapted with permission from Ref. [16].
Scheme 6:

Concurrent photochemoenzymatic cascade comprising a photocatalytic isomerization with the enzymatic reduction of alkenes to the corresponding enantio-enriched products under blue light illumination [116]. Adapted with permission from Ref. [16].

A second example emphasized the power of combining photocatalysis with biocatalysis for the selective functionalization of (non)-activated C–H bonds, which is an ongoing challenge in classical organic synthesis (Scheme 7) [117]. The applied photocatalyst, sodium anthraquinone sulfonate (SAS), was used for the oxidation of alkanes to the corresponding aldehydes or ketones, followed by the enzymatic transformation of the intermediary aldehyde/ketone to various high-value-added (chiral) products, such as formic esters, lactones, chiral cyanohydrins, chiral acyloins, carboxylic acids, chiral cyclohexanones and amines [117]. With this system, a range of chiral products featuring different functional groups were obtained. Additionally, (R)-benzoin and (R)-mandelonitrile were synthesized at gram-scale with high isolated yields and excellent ee (>99 %), highlighting the applicability of the developed PCE cascade. While the photochemoenzymatic synthesis of these two products was successfully demonstrated at gram-scale, the remaining cascades require further optimization in order to overcome limitations related to catalyst inhibition/deactivation due to the formation of reactive oxygen or radical species and cross-reactivity in the presence of light.

Scheme 7: 
Concurrent and step-wise photochemoenzymatic cascades comprising a photocatalytic oxyfunctionalization catalyzed by sodium anthraquinone sulfonate (SAS), followed by further functionalization of the aldehyde or ketone intermediate by different enzymes, such as amine transaminase (ATA), keto reductase (KRED), ene-reductase (ERED), hydroxynitrile lyase (HNL) and benzaldehyde lyase (BAL) [117]. Adapted with permisison from Ref. [16].
Scheme 7:

Concurrent and step-wise photochemoenzymatic cascades comprising a photocatalytic oxyfunctionalization catalyzed by sodium anthraquinone sulfonate (SAS), followed by further functionalization of the aldehyde or ketone intermediate by different enzymes, such as amine transaminase (ATA), keto reductase (KRED), ene-reductase (ERED), hydroxynitrile lyase (HNL) and benzaldehyde lyase (BAL) [117]. Adapted with permisison from Ref. [16].

In addition to the highly desired functionalization of molecule scaffolds, building more complex molecules from cheap and readily available starting materials is a desirable feature in every synthetic endeavor [118]. The combination of enzymes with photocatalysts thereby offers an attractive approach for implementing multi-step synthetic processes to build molecular complexity. Ravelli and Schmidt recently showed that the merging of photocatalytic C–C bond formation with enzymatic asymmetric reduction enables the direct conversion of simple aldehydes and acrylates or unsaturated carboxylic acids into chiral γ-lactones (Scheme 8) [119]. In this example, the photocatalyst tetrabutylammonium decatungstate (TBADT) is used to catalyze the hydroacylation of the starting olefins, yielding the corresponding keto esters/acids. In the subsequent step, an alcohol dehydrogenase (ADH) converts the keto ester/acid to the corresponding chiral alcohol, which undergoes lactonization to the desired γ-lactone. It was demonstrated that the synthesis of several aliphatic and aromatic γ-lactones was thereby achieved with up to >99 % ee and >99 % yield. This example shows that building molecular complexity from simple, cheap and largely available starting materials is possible by merging photocatalysis with biocatalysis to access high-value added chiral compounds.

Scheme 8: 
Step-wise photochemoenzymatic synthesis of chiral γ-lactones comprising decatungstate photocatalyzed hydroacylation of simple olefins with asymmetric reduction of the intermediate 1,4-keto esters/acids catalyzed by ADH. Adapted with permisison from Ref. [119].
Scheme 8:

Step-wise photochemoenzymatic synthesis of chiral γ-lactones comprising decatungstate photocatalyzed hydroacylation of simple olefins with asymmetric reduction of the intermediate 1,4-keto esters/acids catalyzed by ADH. Adapted with permisison from Ref. [119].

Intriguingly, the combination of a photocatalyst with an enzyme can also be used to control the stereochemical outcome of a reaction. For instance, Schmermund et al. showed that the redox potential of a carbon nitride photocatalyst (CN-OA-m) can be tuned by changing the irradiation wavelength to generate electron holes with different oxidation potentials [120]. In combination with an unspecific peroxygenase from A. aegerita, the illumination with of CN-OA-m with green light led to the enantioselective hydroxylation of ethylbenzene to (R)-phenylethanol, whereas blue light irradiation led to the formation of acetophenone, which was further converted with an ADH to the desired (S)-phenylethanol.

5.2 Photoenzymes in biocatalytic cascades

While the majority of PCE cascades rely on the activation of a photocatalytic reaction step by light, also natural and artificial photoenzymes that only perform catalysis upon light irradiation have been implemented in multi-step processes. For instance, a hybrid P450 BM3 biocatalyst containing a covalently attached Ru(II)-diimine photosensitizer has been used in a PCE cascade reaction for the synthesis of trifluoro methylated/hydroxylated substituted arenes [121]. The photochemical properties of the Ru(II) imine photosensitizer allow for the initiation of single electron transfer events, which facilitate the addition of a CF3 radical to arenes. In the hybrid enzyme, the Ru(II)-diimine photosensitizer provides the necessary electrons to carry out hydroxylation reactions on trifluoromethylated substrates upon visible light activation. While a range of different substrates could be converted, lower yields were obtained in most cases. Beneficially, the regio- and stereoselectivity of the hybrid P450 catalyst enables the differentiation between the trifluoromethylated isomers.

Also ‘true’ photoenzymes have been implemented in cascade approaches (Scheme 9) [89, 91, 122]. For example, Huijbers et al. demonstrated the use of CvFAP in a cascade reaction for the conversion of saturated and unsaturated fatty acids to the corresponding long-chain alkanes or alkenes [89]. This approach for synthesizing long-chain alkenes directly from triglycerides via a two-step cascade reaction represents an interesting alternative strategy to the typical transesterification used in biofuel production. The two-step cascade thereby combines enzymatic hydrolysis of triglycerides with subsequent light-driven decarboxylation catalyzed by CvFAP. In addition to the production of alkanes from triglycerides, another bienzymatic cascade was developed to convert castor oil to (R,Z)-octadec-9-en-7-ol [123]. Therein, non-edible castor oil is hydrolyzed with the help of a lipase, yielding of free ricinoleic acid. Subsequently, CvFAP is catalyzing the decarboxylation under blue light illumination, yielding product concentrations of up to 60 mM [123]. The authors suggested that further optimization of the cascade is needed to compensate for the drop in pH caused by the accumulation of fatty acids in the reaction, and thus achieve even higher product yields [89, 123].

Scheme 9: 
Different enzymatic cascades comprising the photodecarboxylase CvFAP [89, 91, 122]. Adapted with permisison from Ref. [16].
Scheme 9:

Different enzymatic cascades comprising the photodecarboxylase CvFAP [89, 91, 122]. Adapted with permisison from Ref. [16].

6 Current challenges of applying light in bioprocesses

The use of light to drive photobiocatalytic reactions is an exciting development in applied biocatalysis, and holds great potential for overcoming current challenges in implementing enzyme-catalyzed processes in organic synthesis. However, as promising and revolutionary it might sound, the desire for the rapid development of novel photobiocatalytic reaction systems drives the distraction from the fundamental concept of reproducibility and implementation in an industrial setting.

This is a particular challenge when it comes to the equipment setup used for photocatalytic reactions. As recently emphasized by Edwards et al., ensuring reproducibility of light-driven bioprocesses and therewith standardization of photochemistry platforms is crucial to ensure a broad applicability of photobiocatalytic reactions [124]. Therefore, a deeper characterization of the used and often home-made light reactor setups is needed to enable a better understanding of the underlying photochemical principles and thus to enable the scale up photobiocatalytic approaches. The comprehensive characterization of the applied light system, such as light source used, photon stoichiometry, internal reaction temperature, light intensity, distance between the light source and reaction mixture, and path length, is crucial to avoid batch-to-batch variability that is often caused by differences of the light source. Encouragingly, a trend towards standardization of photochemistry platforms is there, which is expected to ease a broader implementation of photochemistry in both, academia and industry [124].

Next to the reproducibility challenge, the upscaling and implementation of photobiocatalytic platforms in an industrial setting remains troublesome. For instance, when a photosensitizer, absorbing light at specific wavelengths, is used, charge separation in the presence of an electron acceptor occurs. By implication, the light portion absorbed by a single chromophore is not accessible to another and restricts photoexcitation. This photo limitation not only has crucial impact on photobiocatalytic in vitro reactions, but also on establishing high-cell density cultures of cyanobacteria for respective in vivo applications. As the growth of photoautotrophs is strongly dependent on the availability and intensity of light, circumventing or restricting so-called self-shading effects by optimizing the cultivation conditions has a significant advantage for their implementation as whole-cell biocatalysts [125].

Recent developments in (scalable) photo(bio)reactors (PBRs) are already promising. They could overcome some of the limitations mentioned above. Recent reviews, for example, have highlighted various photobioreactor concepts that can be used to efficiently grow microalgae and perform biocatalytic processes, sometimes even on a large scale [126, 127]. The design and optimization of photobioreactors is an important aspect of photobiochemical processes, with the supply of light to the reactors being one of the major challenges, as it decreases exponentially with the distance from the light source in batch. In this context, continuous flow processes have an advantage as the light penetration is uniform regardless of the reactor scale [126]. Although there are examples of novel reactor designs for photocatalytic reactions in flow, photobiocatalytic applications using flow reactors are still limited. Next to continuous flow processes that are considered in photobiocatalysis, the use of an internal instead of external light source via wireless light emitters (WLEs) provides a promising approach to optimize light absorption in a PBR for cultures of phototrophic microorganisms [128]. A recent study by Hobisch et al. used WLEs to improve the light distribution and product formation in a bubble column reactor using cyanobacteria as whole-cell biocatalysts [129]. In addition, the internal illumination strategy also provides an interesting strategy for continuous flow applications, e.g., in a packed bed reactor.

When it comes to the robustness and catalytic performance of a photobiocatalytic reaction system, the choice of photosensitizer and electron donor is essential. Here, the stability of the photosensitizer often represents a bottleneck as photodegradation or photobleaching of organic dye photosensitizers is frequently observed [130, 131]. Due to that, the application of quantum dots in combination with metal oxides such as TiO2 is emphasized for establishing more robust light-driven reaction setups [132]. On the other hand, the choice of sacrificial electron donor should be made under the consideration of atom economy, availability and compatibility with the reaction conditions. In this context, the use of H2O as sacrificial electron donor is of utmost interest. However, only a few proof of concept studies reported the feasibility of using H2O as direct electron donor in combination with inorganic photocatalysts such as titanium dioxide TiO2 to perform light-driven reactions [133]. The problem with using H2O as electron donor is the high stability and low oxidation potential, hampering general applicability in photobiocatalytic approaches [133].

7 Conclusions

Photobiocatalysis has emerged into an exciting research area with an ever growing impact in applied biocatalysis. Next to the large number of examples that report photobiocatalytic cofactor regeneration via direct or indirect activation of redox enzymes, it is expected that cyanobacteria as chassis for in vivo photobiocatalysis will be further explored and optimized in the near future. Furthermore, the development and application of photochemoenzymatic cascades to combine the outstanding reactivities accessible to photocatalysts with the high selectivity of enzyme-catalyzed reactions greatly highlights the diversity of photobiocatalysis. It is also anticipated that the discovery of additional natural photoenzymes or the design of new artificial photoenzymes, together with the further exploitation of light-induced enzyme promiscuity, will open up unexpected opportunities to extend the catalytic scope currently accessible to biocatalysis. With the recent developments of novel photobioreactor concepts for up-scaling, we expect that in the near future the potential of these newly developed photobiocatalytic strategies for organic synthesis can be fully assessed.


Corresponding author: Sandy Schmidt, Department of Chemical and Pharmaceutical Biology, Groningen Research Institute of Pharmacy, University of Groningen, Antonius Deusinglaan 1, 9713AV Groningen, The Netherlands, E-mail:

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: None declared.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

1. Yoon, TP, Ischay, MA, Du, JN. Visible light photocatalysis as a greener approach to photochemical synthesis. Nat Chem 2010;2:527–32. https://doi.org/10.1038/Nchem.687.Search in Google Scholar PubMed

2. Ciamician, G. The photochemistry of the future. Science 1912;36:385–94. https://doi.org/10.1126/science.36.926.385.Search in Google Scholar PubMed

3. Pyser, JB, Chakrabarty, S, Romero, EO, Narayan, ARH. State-of-the-art biocatalysis. ACS Cent Sci 2021;7:1105–16. https://doi.org/10.1021/acscentsci.1c00273.Search in Google Scholar PubMed PubMed Central

4. Wu, S, Snajdrova, R, Moore, JC, Baldenius, K, Bornscheuer, UT. Biocatalysis: enzymatic synthesis for industrial applications. Angew Chem Int Ed 2021;60:88–119. https://doi.org/10.1002/anie.202006648.Search in Google Scholar PubMed PubMed Central

5. Sun, H, Zhang, H, Ang, EL, Zhao, H. Biocatalysis for the synthesis of pharmaceuticals and pharmaceutical intermediates. Bioorg Med Chem 2018;26:1275–84. https://doi.org/10.1016/j.bmc.2017.06.043.Search in Google Scholar PubMed

6. Bell, EL, Finnigan, W, France, SP, Green, AP, Hayes, MA, Hepworth, LJ, et al.. Biocatalysis. Nat Rev Methods Prim 2021;1:46. https://doi.org/10.1038/s43586-021-00044-z.Search in Google Scholar

7. Dong, JJ, Fernández-Fueyo, E, Hollmann, F, Paul, CE, Pesic, M, Schmidt, S, et al.. Biocatalytic oxidation reactions: a chemist’s perspective. Angew Chem Int Ed 2018;57:9238–61. https://doi.org/10.1002/anie.201800343.Search in Google Scholar PubMed PubMed Central

8. Yang, X, Wang, D. Photocatalysis: from fundamental principles to materials and applications. ACS Appl Energy Mater 2018;1:6657–93. https://doi.org/10.1021/acsaem.8b01345.Search in Google Scholar

9. Stephenson, CRJ, Yoon, TP, MacMillan, DWC. Visible light photocatalysis in organic chemistry. Weinheim: Wiley-VCH; 2017.10.1002/9783527674145Search in Google Scholar

10. Shaw, MH, Twilton, J, MacMillan, DWC. Photoredox catalysis in organic chemistry. J Org Chem 2016;81:6898–926. https://doi.org/10.1021/acs.joc.6b01449.Search in Google Scholar PubMed PubMed Central

11. Marzo, L, Pagire, SK, Reiser, O, König, B. Visible-light photocatalysis: does it make a difference in organic synthesis? Angew Chem Int Ed 2018;57:10034–72. https://doi.org/10.1002/anie.201709766.Search in Google Scholar PubMed

12. Harrison, W, Huang, X, Zhao, H. Photobiocatalysis for abiological transformations. Acc Chem Res 2022;55:1087–96. https://doi.org/10.1021/acs.accounts.1c00719.Search in Google Scholar PubMed

13. Schmermund, L, Jurkaš, V, Özgen, FF, Barone, GD, Büchsenschütz, HC, Winkler, CK, et al.. Photo-biocatalysis: biotransformations in the presence of light. ACS Catal 2019;9:4115–44. https://doi.org/10.1021/acscatal.9b00656.Search in Google Scholar

14. Seel, CJ, Gulder, T. Biocatalysis fueled by light: on the versatile combination of photocatalysis and enzymes. ChemBioChem 2019;20:1871–97. https://doi.org/10.1002/cbic.201800806.Search in Google Scholar PubMed

15. Lee, SH, Choi, DS, Kuk, SK, Park, CB. Photobiocatalysis: activating redox enzymes by direct or indirect transfer of photoinduced electrons. Angew Chem Int Ed 2018;57:7958–85. https://doi.org/10.1002/anie.201710070.Search in Google Scholar PubMed

16. Özgen, FF, Runda, ME, Schmidt, S. Photo-biocatalytic cascades: combining chemical and enzymatic transformations fueled by light. ChemBioChem 2021;22:790–806. https://doi.org/10.1002/cbic.202000587.Search in Google Scholar PubMed PubMed Central

17. Peng, Y, Chen, Z, Xu, J, Wu, Q. Recent advances in photobiocatalysis for selective organic synthesis. Org Process Res Dev 2022;26:1900–13. https://doi.org/10.1021/acs.oprd.1c00413.Search in Google Scholar

18. Seel, CJ, Gulder, T. Biocatalysis fueled by light: on the versatile combination of photocatalysis and enzymes. ChemBioChem 2019;20:1871–97. https://doi.org/10.1002/cbic.201800806.Search in Google Scholar

19. Maciá-Agulló, JA, Corma, A, Garcia, H. Photobiocatalysis: the power of combining photocatalysis and enzymes. Chem – Eur J 2015;21:10940–59. https://doi.org/10.1002/chem.201406437.Search in Google Scholar PubMed

20. Höfler, G, Hollmann, F, Paul, CE, Rauch, M, van Schie, M, Willot, S. Chapter 9 Photocatalysis to promote cell-free biocatalytic reactions. In: Kourist, R, Schmidt, S, editors. The autotrophic biorefinery raw materials from biotechnology. Berlin, Boston: De Gruyter; 2021:247–76 pp.10.1515/9783110550603-009Search in Google Scholar

21. Zhang, W, Hollmann, F. Nonconventional regeneration of redox enzymes – a practical approach for organic synthesis? Chem Commun 2018;54:7281–9. https://doi.org/10.1039/C8CC02219D.Search in Google Scholar

22. Romero, NA, Nicewicz, DA. Organic photoredox catalysis. Chem Rev 2016;116:10075–166. https://doi.org/10.1021/acs.chemrev.6b00057.Search in Google Scholar PubMed

23. Chappelle, EW, Kim, MS, McMurtrey, JE. Ratio analysis of reflectance spectra (RARS): an algorithm for the remote estimation of the concentrations of chlorophyll A, chlorophyll B, and carotenoids in soybean leaves. Remote Sens Environ 1992;39:239–47. https://doi.org/10.1016/0034-4257(92)90089-3.Search in Google Scholar

24. Cardona, T, Sedoud, A, Cox, N, Rutherford, AW. Charge separation in photosystem II: a comparative and evolutionary overview. Biochim Biophys Acta 2012;1817:26–43. https://doi.org/10.1016/j.bbabio.2011.07.012.Search in Google Scholar PubMed

25. Shevela, D, Kern, JF, Govindjee, G, Whitmarsh, J, Messinger, J. Photosystem II. ELS 2021;2:1–16. https://doi.org/10.1002/9780470015902.a0000669.pub2.Search in Google Scholar

26. Mandal, M, Kawashima, K, Saito, K, Ishikita, H. Redox potential of the oxygen-evolving complex in the electron transfer cascade of photosystem II. J Phys Chem Lett 2019;11:249–55. https://doi.org/10.1021/acs.jpclett.9b02831.Search in Google Scholar PubMed

27. Feyza Özgen, F, Runda, ME, Burek, BO, Wied, P, Bloh, JZ, Kourist, R, et al.. Artificial light-harvesting complexes enable Rieske oxygenase catalyzed hydroxylations in non-photosynthetic cells. Angew Chem Int Ed 2020;59:3982–7. https://doi.org/10.1002/anie.201914519.Search in Google Scholar PubMed PubMed Central

28. Lee, SH, Nam, DH, Kim, JH, Baeg, JO, Park, CB. Eosin Y-sensitized artificial photosynthesis by highly efficient visible-light-driven regeneration of nicotinamide cofactor. ChemBioChem 2009;10:1621–4. https://doi.org/10.1002/cbic.200900156.Search in Google Scholar PubMed

29. Pellegrin, Y, Odobel, F. Sacrificial electron donor reagents for solar fuel production. Compt Rendus Chim 2017;20:283–95. https://doi.org/10.1016/j.crci.2015.11.026.Search in Google Scholar

30. Chenault, HK, Simon, ES, Whitesides, GM. Cofactor regeneration for enzyme-catalysed synthesis. Biotechnol Genet Eng Rev 1988;6:221–70. https://doi.org/10.1080/02648725.1988.10647849.Search in Google Scholar PubMed

31. Mordhorst, S, Andexer, JN. Round, round we go – strategies for enzymatic cofactor regeneration. Nat Prod Rep 2020;37:1316–33. https://doi.org/10.1039/D0NP00004C.Search in Google Scholar PubMed

32. Ni, Y, Hollmann, F. Artificial photosynthesis: hybrid systems. In: Jeuken, LJC, editor. Biophotoelectrochemistry from bioelectrochemistry to biophotovoltaics. Cham: Springer International Publishing; 2016:137–58 pp.10.1007/10_2015_5010Search in Google Scholar PubMed

33. Ke, B. Electrolytic reduction of diphosphopyridine nucleotide at some solid metal electrodes. J Am Chem Soc 1956;78:3649–51. https://doi.org/10.1021/ja01596a025.Search in Google Scholar

34. Cunningham, AJ, Underwood, AL. Electrochemical reduction of triphosphopyridine nucleotide. Arch Biochem Biophys 1966;117:88–92. https://doi.org/10.1016/0003-9861(66)90129-9.Search in Google Scholar PubMed

35. Steckhan, E, Herrmann, S, Ruppert, R, Dietz, E, Frede, M, Spika, E. Analytical study of a series of substituted (2,2′-bipyridyl)(pentamethylcyclopentadienyl)rhodium and -iridium complexes with regard to their effectiveness as redox catalysts for the indirect electrochemical and chemical reduction of NAD(P)+. Organometallics 1991;10:1568–77. https://doi.org/10.1021/om00051a056.Search in Google Scholar

36. Hollmann, F, Schmid, A, Steckhan, E. The first synthetic application of a monooxygenase employing indirect electrochemical NADH regeneration. Angew Chem Int Ed 2001;40:169–71. https://doi.org/10.1002/1521-3773(20010105)40:1<169::AID-ANIE169>3.0.CO;2-T.10.1002/1521-3773(20010105)40:1<169::AID-ANIE169>3.0.CO;2-TSearch in Google Scholar

37. Lee, SH, Kwon, YC, Kim, DM, Park, CB. Cytochrome P450-catalyzed O-dealkylation coupled with photochemical NADPH regeneration. Biotechnol Bioeng 2013;110:383–90. https://doi.org/10.1002/bit.24729.Search in Google Scholar

38. Mandler, D, Willner, I. Photosensitized NAD (P) H regeneration systems; application in the reduction of butan-2-one, pyruvic, and acetoacetic acids and in the reductive amination of pyruvic and oxoglutaric acid to amino acid. J Chem Soc, Perkin Trans 1986;2:805–11. https://doi.org/10.1039/P29860000805.Search in Google Scholar

39. Peers, MK, Toogood, HS, Heyes, DJ, Mansell, D, Coe, BJ, Scrutton, NS. Light-driven biocatalytic reduction of α,β-unsaturated compounds by ene reductases employing transition metal complexes as photosensitizers. Catal Sci Technol 2016;6:169–77. https://doi.org/10.1039/c5cy01642h.Search in Google Scholar

40. Willner, I, Mandler, D, Riklin, A. Photoinduced carbon dioxide fixation forming malic and isocitric acid. J Chem Soc Chem Commun 1986:1022–4. https://doi.org/10.1039/C39860001022.Search in Google Scholar

41. Mandler, D, Willner, I. Solar light induced formation of chiral 2-butanol in an enzyme-catalyzed chemical system. J Am Chem Soc 1984;106:5352–3. https://doi.org/10.1021/ja00330a053.Search in Google Scholar

42. Mandler, D, Willner, I. Photoinduced enzyme-catalysed synthesis of amino acids by visible light. J Chem Soc Chem Commun 1986:851–3. https://doi.org/10.1039/C39860000851.Search in Google Scholar

43. Lee, SH, Kwon, YC, Kim, DM, Park, CB. Cytochrome P450-catalyzed O-dealkylation coupled with photochemical NADPH regeneration. Biotechnol Bioeng 2013;110:383–90. https://doi.org/10.1002/bit.24729.Search in Google Scholar PubMed

44. Nam, DH, Park, CB. Visible light-driven NADH regeneration sensitized by proflavine for biocatalysis. ChemBioChem 2012;13:1278–82. https://doi.org/10.1002/cbic.201200115.Search in Google Scholar PubMed

45. Nam, DH, Lee, SH, Park, CB. CDTE, CDSE, and CDS nanocrystals for highly efficient regeneration of nicotinamide cofactor under visible light. Small 2010;6:922–6. https://doi.org/10.1002/smll.201000077.Search in Google Scholar PubMed

46. Yadav, RK, Baeg, JO, Oh, GH, Park, NJ, Kong, KJ, Kim, J, et al.. A photocatalyst-enzyme coupled artificial photosynthesis system for solar energy in production of formic acid from CO2. J Am Chem Soc 2012;134:11455–61. https://doi.org/10.1021/ja3009902.Search in Google Scholar PubMed

47. van Esch, JH, Hoffmann, MAM, Nolte, RJM. Reduction of nicotinamides, flavins, and manganese porphyrins by formate, catalyzed by membrane-bound rhodium complexes. J Org Chem 1995;60:1599–610. https://doi.org/10.1021/jo00111a018.Search in Google Scholar

48. Hari, DP, König, B. Synthetic applications of eosin Y in photoredox catalysis. Chem Commun 2014;50:6688–99. https://doi.org/10.1039/c4cc00751d.Search in Google Scholar PubMed

49. Lee, SH, Nam, DH, Park, CB. Screening xanthene dyes for visible light-driven nicotinamide adenine dinucleotide regeneration and photoenzymatic synthesis. Adv Synth Catal 2009;351:2589–94. https://doi.org/10.1002/adsc.200900547.Search in Google Scholar

50. Yadav, RK, Baeg, J-O, Kumar, A, Kong, K, Oh, GH, Park, N-J. Graphene–BODIPY as a photocatalyst in the photocatalytic–biocatalytic coupled system for solar fuel production from CO2. J Mater Chem A 2014;2:5068–76. https://doi.org/10.1039/C3TA14442A.Search in Google Scholar

51. Ryu, J, Lee, SH, Nam, DH, Park, CB. Rational design and engineering of quantum-dot-sensitized TiO2 nanotube arrays for artificial photosynthesis. Adv Mater 2011;23:1883–8. https://doi.org/10.1002/adma.201004576.Search in Google Scholar PubMed

52. Gargiulo, S, Arends, IWCE, Hollmann, F. A photoenzymatic system for alcohol oxidation. ChemCatChem 2011;3:338–42. https://doi.org/10.1002/cctc.201000317.Search in Google Scholar

53. Rickus, JL, Chang, PL, Tobin, AJ, Zink, JI, Dunn, B. Photochemical coenzyme regeneration in an enzymatically active optical material. J Phys Chem B 2004;108:9325–32. https://doi.org/10.1021/jp038051g.Search in Google Scholar

54. Lee, SH, Choi, DS, Pesic, M, Lee, YW, Paul, CE, Hollmann, F, et al.. Cofactor-free, direct photoactivation of enoate reductases for the asymmetric reduction of C=C bonds. Angew Chem 2017;129:8807–11. https://doi.org/10.1002/ange.201702461.Search in Google Scholar

55. Grau, MM, Van Der Toorn, JC, Otten, LG, Macheroux, P, Taglieber, A, Zilly, FE, et al.. Photoenzymatic reduction of C=C double bonds. Adv Synth Catal 2009;351:3279–86. https://doi.org/10.1002/adsc.200900560.Search in Google Scholar

56. Lee, SH, Choi, DS, Pesic, M, Lee, YW, Paul, CE, Hollmann, F, et al.. Cofactor-free, direct photoactivation of enoate reductases for the asymmetric reduction of C=C bonds. Angew Chem Int Ed 2017;56:8681–5. https://doi.org/10.1002/anie.201702461.Search in Google Scholar PubMed PubMed Central

57. Park, JH, Lee, SH, Cha, GS, Choi, DS, Nam, DH, Lee, JH, et al.. Cofactor-free light-driven whole-cell cytochrome P450 catalysis. Angew Chem Int Ed 2015;54:969–73. https://doi.org/10.1002/anie.201410059.Search in Google Scholar PubMed

58. Lin, B, Tao, Y. Whole-cell biocatalysts by design. Microb Cell Factories 2017;16:1–12. https://doi.org/10.1186/s12934-017-0724-7.Search in Google Scholar PubMed PubMed Central

59. Wachtmeister, J, Rother, D. Recent advances in whole cell biocatalysis techniques bridging from investigative to industrial scale. Curr Opin Biotechnol 2016;42:169–77. https://doi.org/10.1016/j.copbio.2016.05.005.Search in Google Scholar PubMed

60. De Carvalho, CCCR. Whole cell biocatalysts: essential workers from nature to the industry. Microb Biotechnol 2017;10:250–63. https://doi.org/10.1111/1751-7915.12363.Search in Google Scholar PubMed PubMed Central

61. Wu, S, Li, Z. Whole-cell cascade biotransformations for one-pot multistep organic synthesis. ChemCatChem 2018;10:2164–78. https://doi.org/10.1002/cctc.201701669.Search in Google Scholar

62. Kumar, B, Bhardwaj, N, Agrawal, K, Chaturvedi, V, Verma, P. Current perspective on pretreatment technologies using lignocellulosic biomass: an emerging biorefinery concept. Fuel Process Technol 2020;199:106244. https://doi.org/10.1016/j.fuproc.2019.106244.Search in Google Scholar

63. Zhou, Z, Liu, D, Zhao, X. Conversion of lignocellulose to biofuels and chemicals via sugar platform: an updated review on chemistry and mechanisms of acid hydrolysis of lignocellulose. Renew Sustain Energy Rev 2021;146:111169. https://doi.org/10.1016/j.rser.2021.111169.Search in Google Scholar

64. Jodlbauer, J, Rohr, T, Spadiut, O, Mihovilovic, MD, Rudroff, F. Biocatalysis in green and blue: cyanobacteria. Trends Biotechnol 2021;39:875–89. https://doi.org/10.1016/j.tibtech.2020.12.009.Search in Google Scholar PubMed

65. Nakamura, K, Yamanaka, R, Tohi, K, Hamada, H. Cyanobacterium-catalyzed asymmetric reduction of ketones. Tetrahedron Lett 2000;41:6799–802. https://doi.org/10.1016/S0040-4039(00)01132-1.Search in Google Scholar

66. Yamanaka, R, Nakamura, K, Murakami, M, Murakami, A. Selective synthesis of cinnamyl alcohol by cyanobacterial photobiocatalysts. Tetrahedron Lett 2015;56:1089–91. https://doi.org/10.1016/j.tetlet.2015.01.092.Search in Google Scholar

67. Pacheco, CC, Ferreira, EA, Oliveira, P, Tamagnini, P. Chapter 6 synthetic biology of cyanobacteria. In: Kourist, R, Schmidt, S, editors. The autotrophic biorefinery: raw materials from biotechnology. Berlin, Boston: De Gruyter; 2021:131–72 pp.10.1515/9783110550603-006Search in Google Scholar

68. Grimm, HC, Erdem, E, Kourist, R. Chapter 8 Biocatalytic applications of autotrophic organisms. In: Kourist, R, Schmidt, S, editors. The autotrophic biorefinery: raw materials from biotechnology. Berlin, Boston: De Gruyter; 2021:207–46 pp.10.1515/9783110550603-008Search in Google Scholar

69. Köninger, K, Gómez Baraibar, Á, Mügge, C, Paul, CE, Hollmann, F, Nowaczyk, MM, et al.. Recombinant cyanobacteria for the asymmetric reduction of C=C bonds fueled by the biocatalytic oxidation of water. Angew Chem Int Ed 2016;55:5582–5. https://doi.org/10.1002/anie.201601200.Search in Google Scholar PubMed

70. Büchsenschütz, HC, Vidimce-Risteski, V, Eggbauer, B, Schmidt, S, Winkler, CK, Schrittwieser, JH, et al.. Stereoselective biotransformations of cyclic imines in recombinant cells of Synechocystis sp. PCC 6803. ChemCatChem 2020;12:726–30. https://doi.org/10.1002/cctc.201901592.Search in Google Scholar

71. Hoschek, A, Bühler, B, Schmid, A. Overcoming the gas–liquid mass transfer of oxygen by coupling photosynthetic water oxidation with biocatalytic oxyfunctionalization. Angew Chem Int Ed 2017;56:15146–9. https://doi.org/10.1002/anie.201706886.Search in Google Scholar PubMed PubMed Central

72. Hoschek, A, Bühler, B, Schmid, A. Stabilization and scale-up of photosynthesis-driven ω-hydroxylation of nonanoic acid methyl ester by two-liquid phase whole-cell biocatalysis. Biotechnol Bioeng 2019;116:1887–900. https://doi.org/10.1002/bit.27006.Search in Google Scholar PubMed

73. Hoschek, A, Toepel, J, Hochkeppel, A, Karande, R, Bühler, B, Schmid, A. Light-dependent and aeration-independent gram-scale hydroxylation of cyclohexane to cyclohexanol by CYP450 harboring Synechocystis sp. PCC 6803. Biotechnol J 2019;14:1–10. https://doi.org/10.1002/biot.201800724.Search in Google Scholar PubMed

74. Assil-Companioni, L, Büchsenschütz, HC, Solymosi, D, Dyczmons-Nowaczyk, NG, Bauer, KKF, Wallner, S, et al.. Engineering of NADPH supply boosts photosynthesis-driven biotransformations. ACS Catal 2020;10:11864–77. https://doi.org/10.1021/acscatal.0c02601.Search in Google Scholar PubMed PubMed Central

75. Zhang, W, Burek, BO, Fernández-Fueyo, E, Alcalde, M, Bloh, JZ, Hollmann, F. Selective activation of C–H bonds in a cascade process combining photochemistry and biocatalysis. Angew Chem Int Ed 2017;56:15451–5. https://doi.org/10.1002/anie.201708668.Search in Google Scholar PubMed PubMed Central

76. Zhang, W, Fernández-Fueyo, E, Ni, Y, Van Schie, M, Gacs, J, Renirie, R, et al.. Selective aerobic oxidation reactions using a combination of photocatalytic water oxidation and enzymatic oxyfunctionalizations. Nat Catal 2018;1:55–62. https://doi.org/10.1038/s41929-017-0001-5.Search in Google Scholar PubMed PubMed Central

77. Churakova, E, Kluge, M, Ullrich, R, Arends, I, Hofrichter, M, Hollmann, F. Specific photobiocatalytic oxyfunctionalization reactions. Angew Chem Int Ed 2011;50:10716–9. https://doi.org/10.1002/anie.201105308.Search in Google Scholar PubMed

78. Yuan, B, Mahor, D, Fei, Q, Wever, R, Alcalde, M, Zhang, W, et al.. Water-soluble anthraquinone photocatalysts enable methanol-driven enzymatic halogenation and hydroxylation reactions. ACS Catal 2020;10:8277–84. https://doi.org/10.1021/acscatal.0c01958.Search in Google Scholar PubMed PubMed Central

79. Van Schie, MMCH, Zhang, W, Tieves, F, Choi, DS, Park, CB, Burek, BO, et al.. Cascading g-C3N4 and peroxygenases for selective oxyfunctionalization reactions. ACS Catal 2019;9:7409–17. https://doi.org/10.1021/acscatal.9b01341.Search in Google Scholar

80. Hobisch, M, van Schie, MMCH, Kim, J, Røjkjær Andersen, K, Alcalde, M, Kourist, R, et al.. Solvent-free photobiocatalytic hydroxylation of cyclohexane. ChemCatChem 2020;12:1–6. https://doi.org/10.1002/cctc.202000512.Search in Google Scholar

81. Rupert, CS, Goodgal, SH, Herriott, RM. Photoreactivation in vitro of ultraviolet-inactivated Hemophilus influenzae transforming factor. J Gen Physiol 1958;41:451–71. https://doi.org/10.1085/jgp.41.3.451.Search in Google Scholar PubMed PubMed Central

82. Dikbas, UM, Tardu, M, Canturk, A, Gul, S, Ozcelik, G, Baris, I, et al.. Identification and characterization of a new class of (6-4) photolyase from Vibrio cholerae. Biochemistry 2019;58:4352–60. https://doi.org/10.1021/acs.biochem.9b00766.Search in Google Scholar PubMed

83. Zhong, D. Ultrafast catalytic processes in enzymes. Curr Opin Chem Biol 2007;11:174–81. https://doi.org/10.1016/j.cbpa.2007.02.034.Search in Google Scholar PubMed

84. Kaschner, M, Loeschcke, A, Krause, J, Minh, BQ, Heck, A, Endres, S, et al.. Discovery of the first light-dependent protochlorophyllide oxidoreductase in anoxygenic phototrophic bacteria. Mol Microbiol 2014;93:1066–78. https://doi.org/10.1111/mmi.12719.Search in Google Scholar PubMed

85. Klement, H, Helfrich, M, Oster, U, Schoch, S, Rüdiger, W. Pigment-free NADPH: protochlorophyllide oxidoreductase from Avena sativa L. Purification and substrate specificity. Eur J Biochem 1999;265:862–74. https://doi.org/10.1046/j.1432-1327.1999.00627.x.Search in Google Scholar PubMed

86. Schmermund, L, Bierbaumer, S, Schein, VK, Winkler, CK, Kara, S, Kroutil, W. Extending the library of light-dependent protochlorophyllide oxidoreductases and their solvent tolerance, stability in light and cofactor flexibility. ChemCatChem 2020;12:4044–51. https://doi.org/10.1002/cctc.202000561.Search in Google Scholar

87. Sorigué, D, Légeret, B, Cuiné, S, Blangy, S, Moulin, S, Billon, E, et al.. An algal photoenzyme converts fatty acids to hydrocarbons. Science 2017;357:903–7. https://doi.org/10.1126/science.aan6349.Search in Google Scholar PubMed

88. Heyes, DJ, Lakavath, B, Hardman, SJO, Sakuma, M, Hedison, TM, Scrutton, NS. Photochemical mechanism of light-driven fatty acid photodecarboxylase. ACS Catal 2020;10:6691–6. https://doi.org/10.1021/acscatal.0c01684.Search in Google Scholar PubMed PubMed Central

89. Huijbers, MME, Zhang, W, Tonin, F, Hollmann, F. Light-driven enzymatic decarboxylation of fatty acids. Angew Chem Int Ed 2018;57:13648–51. https://doi.org/10.1002/anie.201807119.Search in Google Scholar PubMed PubMed Central

90. Xu, J, Hu, Y, Fan, J, Arkin, M, Li, D, Peng, Y, et al.. Light-driven kinetic resolution of α-functionalized carboxylic acids enabled by an engineered fatty acid photodecarboxylase. Angew Chem 2019;310027:ange.201903165. https://doi.org/10.1002/ange.201903165.Search in Google Scholar

91. Zhang, W, Lee, JH, Younes, SHH, Tonin, F, Hagedoorn, PL, Pichler, H, et al.. Photobiocatalytic synthesis of chiral secondary fatty alcohols from renewable unsaturated fatty acids. Nat Commun 2020;11:1–8. https://doi.org/10.1038/s41467-020-16099-7.Search in Google Scholar PubMed PubMed Central

92. Cha, H-J, Hwang, S-Y, Lee, D-S, Akula, RK, Kwon, Y-U, Voß, M, et al.. Whole-cell photoenzymatic cascades to synthesize long chain aliphatic amines and esters from renewable fatty acids. Angew Chem Int Ed 2020;59:7024–28. https://doi.org/10.1002/anie.201915108.Search in Google Scholar PubMed

93. Zhang, Y, Zhao, Y, Li, R, Liu, J. Bioinspired NADH regeneration based on conjugated photocatalytic systems. Sol RRL 2021;5:2000339. https://doi.org/10.1002/solr.202000339.Search in Google Scholar

94. Gu, Y, Ellis-Guardiola, K, Srivastava, P, Lewis, JC. Preparation, characterization, and oxygenase activity of a photocatalytic artificial enzyme. ChemBioChem 2015;16:1880–3. https://doi.org/10.1002/cbic.201500165.Search in Google Scholar PubMed PubMed Central

95. Zubi, YS, Liu, B, Gu, Y, Sahoo, D, Lewis, JC. Controlling the optical and catalytic properties of artificial metalloenzyme photocatalysts using chemogenetic engineering. Chem Sci 2022;13:1459–68. https://doi.org/10.1039/D1SC05792H.Search in Google Scholar

96. Liu, X, Kang, F, Hu, C, Wang, L, Xu, Z, Zheng, D, et al.. A genetically encoded photosensitizer protein facilitates the rational design of a miniature photocatalytic CO2-reducing enzyme. Nat Chem 2018;10:1201–6. https://doi.org/10.1038/s41557-018-0150-4.Search in Google Scholar PubMed

97. Kang, F, Yu, L, Xia, Y, Yu, M, Xia, L, Wang, Y, et al.. Rational design of a miniature photocatalytic CO2-reducing enzyme. ACS Catal 2021;11:5628–35. https://doi.org/10.1021/acscatal.1c00287.Search in Google Scholar

98. Fu, Y, Huang, J, Wu, Y, Liu, X, Zhong, F, Wang, J. Biocatalytic cross-coupling of aryl halides with a genetically engineered photosensitizer artificial dehalogenase. J Am Chem Soc 2021;143:617–22. https://doi.org/10.1021/jacs.0c10882.Search in Google Scholar PubMed

99. Trimble, JS, Crawshaw, R, Hardy, FJ, Levy, CW, Brown, MJB, Fuerst, DE, et al.. A designed photoenzyme promotes enantioselective [2+2]-cycloadditions via triplet energy transfer. Nature 2022;611:709–14. https://doi.org/10.1038/s41586-022-05335-3.Search in Google Scholar PubMed

100. Sun, N, Huang, J, Qian, J, Zhou, T-P, Guo, J, Tang, L, et al.. Enantioselective [2+2]-cycloadditions with triplet photoenzymes. Nature 2022;611:715–20. https://doi.org/10.1038/s41586-022-05342-4.Search in Google Scholar PubMed

101. Van Der Meer, JY, Poddar, H, Baas, BJ, Miao, Y, Rahimi, M, Kunzendorf, A, et al.. Using mutability landscapes of a promiscuous tautomerase to guide the engineering of enantioselective Michaelases. Nat Commun 2016;7:10911. https://doi.org/10.1038/ncomms10911.Search in Google Scholar PubMed PubMed Central

102. Renata, H, Wang, ZJ, Arnold, FH. Expanding the enzyme universe: accessing non-natural reactions by mechanism-guided directed evolution. Angew Chem Int Ed 2015;54:3351–67. https://doi.org/10.1002/anie.201409470.Search in Google Scholar PubMed PubMed Central

103. Arnold, FH. Directed evolution: bringing new chemistry to life. Angew Chem Int Ed 2018;57:4143–8. https://doi.org/10.1002/anie.201708408.Search in Google Scholar PubMed PubMed Central

104. Brandenberg, OF, Fasan, R, Arnold, FH. Exploiting and engineering hemoproteins for abiological carbene and nitrene transfer reactions. Curr Opin Biotechnol 2017;47:102–11. https://doi.org/10.1016/j.copbio.2017.06.005.Search in Google Scholar PubMed PubMed Central

105. Coelho, PS, Brustad, EM, Kannan, A, Arnold, FH. Olefin cyclopropanation via carbene transfer catalyzed by engineered cytochrome P450 enzymes. Science 2013;339:307–10. https://doi.org/10.1126/science.1231434.Search in Google Scholar PubMed

106. Zhang, RK, Chen, K, Huang, X, Wohlschlager, L, Renata, H, Arnold, FH. Enzymatic assembly of carbon–carbon bonds via iron-catalysed sp3 C–H functionalization. Nature 2019;565:67–72. https://doi.org/10.1038/s41586-018-0808-5.Search in Google Scholar PubMed PubMed Central

107. Yang, Y, Arnold, FH. Navigating the unnatural reaction space: directed evolution of heme proteins for selective carbene and nitrene transfer. Acc Chem Res 2021;54:1209–25. https://doi.org/10.1021/acs.accounts.0c00591.Search in Google Scholar PubMed PubMed Central

108. Dunham, NP, Arnold, FH. Nature’s machinery, repurposed: expanding the repertoire of iron-dependent oxygenases. ACS Catal 2020;10:12239–55. https://doi.org/10.1021/acscatal.0c03606.Search in Google Scholar PubMed PubMed Central

109. Ye, Y, Fu, H, Hyster, TK. Activation modes in biocatalytic radical cyclization reactions. J Ind Microbiol Biotechnol 2021;48:1–18. https://doi.org/10.1093/jimb/kuab021.Search in Google Scholar PubMed PubMed Central

110. Hyster, TK. Radical biocatalysis: using non-natural single electron transfer mechanisms to access new enzymatic functions. Synlett 2020;31:248–54. https://doi.org/10.1055/s-0037-1611818.Search in Google Scholar

111. Emmanuel, MA, Greenberg, NR, Oblinsky, DG, Hyster, TK. Accessing non-natural reactivity by irradiating nicotinamide-dependent enzymes with light. Nature 2016;540:414–7. https://doi.org/10.1038/nature20569.Search in Google Scholar PubMed

112. Biegasiewicz, KF, Cooper, SJ, Gao, X, Oblinsky, DG, Kim, JH, Garfinkle, SE, et al.. Photoexcitation of flavoenzymes enables a stereoselective radical cyclization. Science 2019;364:1166–9. https://doi.org/10.1126/science.aaw1143.Search in Google Scholar PubMed PubMed Central

113. Biegasiewicz, KF, Cooper, SJ, Emmanuel, MA, Miller, DC, Hyster, TK. Catalytic promiscuity enabled by photoredox catalysis in nicotinamide-dependent oxidoreductases. Nat Chem 2018;10:770–5. https://doi.org/10.1038/s41557-018-0059-y.Search in Google Scholar PubMed

114. Huang, X, Wang, B, Wang, Y, Jiang, G, Feng, J, Zhao, H. Photoenzymatic enantioselective intermolecular radical hydroalkylation. Nature 2020;584:69–74. https://doi.org/10.1038/s41586-020-2406-6.Search in Google Scholar PubMed

115. Schrittwieser, JH, Velikogne, S, Hall, M, Kroutil, W. Artificial biocatalytic linear cascades for preparation of organic molecules. Chem Rev 2018;118:270–348. https://doi.org/10.1021/acs.chemrev.7b00033.Search in Google Scholar PubMed

116. Litman, ZC, Wang, Y, Zhao, H, Hartwig, JF. Cooperative asymmetric reactions combining photocatalysis and enzymatic catalysis. Nature 2018;560:355–9. https://doi.org/10.1038/s41586-018-0413-7.Search in Google Scholar PubMed

117. Zhang, W, Fueyo, EF, Hollmann, F, Martin, LL, Pesic, M, Wardenga, R, et al.. Combining photo-organo redox- and enzyme catalysis facilitates asymmetric C–H bond functionalization. Eur J Org Chem 2019;2019:80–4. https://doi.org/10.1002/ejoc.201801692.Search in Google Scholar PubMed PubMed Central

118. Nicolaou, KC, Hale, CRH, Nilewski, C, Ioannidou, HA. Constructing molecular complexity and diversity: total synthesis of natural products of biological and medicinal importance. Chem Soc Rev 2012;41:5185–238. https://doi.org/10.1039/c2cs35116a.Search in Google Scholar PubMed PubMed Central

119. Özgen, FF, Jorea, A, Capaldo, L, Kourist, R, Ravelli, D, Schmidt, S. The synthesis of chiral γ-lactones by merging decatungstate photocatalysis with biocatalysis. ChemCatChem 2022;14:e202200855. https://doi.org/10.1002/cctc.202200855.Search in Google Scholar

120. Schmermund, L, Reischauer, S, Bierbaumer, S, Winkler, CK, Diaz-Rodriguez, A, Edwards, LJ, et al.. Chromoselective photocatalysis enables stereocomplementary biocatalytic pathways**. Angew Chem 2021;133:7041–5. https://doi.org/10.1002/ange.202100164.Search in Google Scholar

121. Sosa, V, Melkie, M, Sulca, C, Li, JJ, Tang, L, Li, JJ, et al.. Selective light-driven chemoenzymatic trifluoromethylation/hydroxylation of substituted arenes. ACS Catal 2018;8:2225–9. https://doi.org/10.1021/acscatal.7b04160.Search in Google Scholar

122. Ma, Y, Zhang, X, Zhang, W, Li, P, Li, Y, Hollmann, F, et al.. Photoenzymatic production of next generation biofuels from natural triglycerides combining a hydrolase and a photodecarboxylase. ChemPhotoChem 2020;4:39–44. https://doi.org/10.1002/cptc.201900205.Search in Google Scholar

123. Ma, Y, Zhang, X, Li, Y, Li, P, Hollmann, F, Wang, Y. Production of fatty alcohols from non-edible oils by enzymatic cascade reactions. Sustain Energy Fuels 2020;4:1–6. https://doi.org/10.1039/d0se00848f.Search in Google Scholar

124. Bonfield, HE, Knauber, T, Lévesque, F, Moschetta, EG, Susanne, F, Edwards, LJ. Photons as a 21st century reagent. Nat Commun 2020;11:2–5. https://doi.org/10.1038/s41467-019-13988-4.Search in Google Scholar PubMed PubMed Central

125. Alagesan, S, Gaudana, SB, Krishnakumar, S, Wangikar, PP. Model based optimization of high cell density cultivation of nitrogen-fixing cyanobacteria. Bioresour Technol 2013;148:228–33. https://doi.org/10.1016/j.biortech.2013.08.144.Search in Google Scholar PubMed

126. Chanquia, SN, Valotta, A, Gruber-Woelfler, H, Kara, S. Photobiocatalysis in continuous flow. Front Catal 2022;1:1–15. https://doi.org/10.3389/fctls.2021.816538.Search in Google Scholar

127. Chanquia, SN, Vernet, G, Kara, S. Photobioreactors for cultivation and synthesis: specifications, challenges, and perspectives. Eng Life Sci 2022;22:712–24. https://doi.org/10.1002/elsc.202100070.Search in Google Scholar PubMed PubMed Central

128. Heining, M, Sutor, A, Stute, SC, Lindenberger, CP, Buchholz, R. Internal illumination of photobioreactors via wireless light emitters: a proof of concept. J Appl Phycol 2015;27:59–66. https://doi.org/10.1007/s10811-014-0290-x.Search in Google Scholar

129. Hobisch, M, Spasic, J, Malihan-Yap, L, Barone, GD, Castiglione, K, Tamagnini, P, et al.. Internal illumination to overcome the cell density limitation in the scale-up of whole-cell photobiocatalysis. ChemSusChem 2021;14:3219–25. https://doi.org/10.1002/cssc.202100832.Search in Google Scholar PubMed PubMed Central

130. Herculano, LS, Malacarne, LC, Zanuto, VS, Lukasievicz, GVB, Capeloto, OA, Astrath, NGC. Investigation of the photobleaching process of eosin Y in aqueous solution by thermal lens spectroscopy. J Phys Chem B 2013;117:1932–7. https://doi.org/10.1088/2050-6120/ab7365.Search in Google Scholar PubMed

131. Demchenko, AP. Photobleaching of organic fluorophores: quantitative characterization, mechanisms, protection. Methods Appl Fluoresc 2020;8:22001. https://doi.org/10.1016/j.jcis.2013.03.049.Search in Google Scholar PubMed

132. Kalanur, SS, Hwang, YJ, Joo, O-S. Construction of efficient CdS–TiO2 heterojunction for enhanced photocurrent, photostability, and photoelectron lifetimes. J Colloid Interface Sci 2013;402:94–9. https://doi.org/10.1016/j.jcis.2013.03.049.Search in Google Scholar

133. Mifsud, M, Gargiulo, S, Iborra, S, Arends, IWCE, Hollmann, F, Corma, A. Photobiocatalytic chemistry of oxidoreductases using water as the electron donor. Nat Commun 2014;5:1–6. https://doi.org/10.1038/ncomms4145.Search in Google Scholar PubMed

134. Ju, X, Tang, Y, Liang, X, Hou, M, Wan, Z, Tao, J. Development of a biocatalytic process to prepare (S)-N-Boc-3-hydroxypiperidine. Org Process Res Dev 2014;18:827–30. https://doi.org/10.1021/op500022y.Search in Google Scholar

135. Seelbach, K, Riebel, B, Hummel, W, Kula, M-R, Tishkov, VI, Egorov, AM, et al.. A novel, efficient regenerating method of NADPH using a new formate dehydrogenase. Tetrahedron Lett 1996;37:1377–80. https://doi.org/10.1016/0040-4039(96)00010-X.Search in Google Scholar

136. Xu, Z, Jing, K, Liu, Y, Cen, P. High-level expression of recombinant glucose dehydrogenase and its application in NADPH regeneration. J Ind Microbiol Biotechnol 2007;34:83–90. https://doi.org/10.1007/s10295-006-0168-2.Search in Google Scholar PubMed

137. Shaked, Z, Whitesides, GM. Enzyme-catalyzed organic synthesis: NADH regeneration by using formate dehydrogenase. J Am Chem Soc 1980;102:7104–5. https://doi.org/10.1021/ja00543a038.Search in Google Scholar

Received: 2022-07-23
Accepted: 2023-03-10
Published Online: 2023-04-17

© 2023 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 30.4.2024 from https://www.degruyter.com/document/doi/10.1515/psr-2022-0109/html
Scroll to top button