Next Article in Journal
Paternal Obesity Induced by High-Fat Diet Impairs the Metabolic and Reproductive Health of Progeny in Rats
Previous Article in Journal
Hibernation-Induced microRNA Expression Promotes Signaling Pathways and Cell Cycle Dysregulation in Ictidomys tridecemlineatus Cardiac Tissue
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis of Metabolites and Metabolite-like Compounds Using Biocatalytic Systems

by
Roland Wohlgemuth
1,2,3
1
MITR, Institute of Applied Radiation Chemistry, Faculty of Chemistry, Lodz University of Technology, Zeromskiego Street 116, 90-924 Lodz, Poland
2
Swiss Coordination Committee Biotechnology (SKB), 8021 Zurich, Switzerland
3
European Society of Applied Biocatalysis (ESAB), 1000 Brussels, Belgium
Metabolites 2023, 13(10), 1097; https://doi.org/10.3390/metabo13101097
Submission received: 16 August 2023 / Revised: 13 October 2023 / Accepted: 15 October 2023 / Published: 19 October 2023
(This article belongs to the Section Thematic Reviews)

Abstract

:
Methodologies for the synthesis and purification of metabolites, which have been developed following their discovery, analysis, and structural identification, have been involved in numerous life science milestones. The renewed focus on the small molecule domain of biological cells has also created an increasing awareness of the rising gap between the metabolites identified and the metabolites which have been prepared as pure compounds. The design and engineering of resource-efficient and straightforward synthetic methodologies for the production of the diverse and numerous metabolites and metabolite-like compounds have attracted much interest. The variety of metabolic pathways in biological cells provides a wonderful blueprint for designing simplified and resource-efficient synthetic routes to desired metabolites. Therefore, biocatalytic systems have become key enabling tools for the synthesis of an increasing number of metabolites, which can then be utilized as standards, enzyme substrates, inhibitors, or other products, or for the discovery of novel biological functions.

Graphical Abstract

1. Introduction

Traditional knowledge of bioresources for microbial, plant and animal metabolites, as well as their processing and application, has been contributing tremendously to quality of life for thousands of years. Small molecular weight natural products have accompanied humankind and supported their quality of life in highly relevant areas, such as nutrition, diagnostics and therapy of diseases, dyes, cosmetics, and well-being, with an ever-increasing knowledge base. Numerous achievements have been made in terms of the isolation and purification of metabolites from natural bioresources and the elucidation of their molecular structures, demonstrating their large structural diversity. Renewed interest in the small molecule domain of biology [1] and the structures and functions of natural products and metabolites have brought the spotlight back onto metabolism [2]. With the great variety of small molecular weight compounds formed by the metabolism of biological cells, such as human and animal-derived metabolites, natural products from microbes and plants [3], as well as the large number of derivatives formed by the metabolism of biological cells from synthetic new molecular entities, a unified definition of ‘metabolite’ as a small molecular weight compound formed by the metabolism of a biological cell is used here. Manifold interactions between metabolites and other biomolecules, such as proteins, DNA, RNA, or other metabolites, within the same as well as between other biological cells, are of much fundamental interest for biological sensing, controlling and regulating processes at the genetic, epigenetic, transcriptomic, and proteomic levels. As precise experimental investigations of metabolites and their interactions with other biomolecules, such as their role as a substrate, inhibitor, or activator of an enzyme, are only possible by having available and pure metabolites, synthetic access is essential.
The concept of the metabolome [4], coined 25 years ago, has contributed to revitalizing interest in metabolites and metabolic pathways. Analyses of large numbers of metabolites have been facilitated by significant advances in powerful analytical methodologies with high information content [5,6,7], such as mass spectrometry (MS) [8] and nuclear magnetic resonance (NMR) [9], on which known and unknown metabolites can be based. A growing number of large databases are focusing on: (a) metabolite analysis using MS [8] or NMR [10]; (b) metabolites, metabolic pathways, natural products, and small molecules of biological interest [11,12,13]; and (c) species-specific metabolites and metabolic pathways (see Table 1 for a selection of species-specific metabolite databases) [14,15,16,17,18,19,20,21,22,23], providing fast access to information on the increasing number of identified metabolites and metabolic pathways.
The development of analytical methodologies for identifying biologically active metabolites and the interactions between metabolites and proteins and other biomolecules [24,25,26] offers great opportunities for delineating the molecular mechanisms of numerous biological processes. These include the activities of proteins and their modulation by activators, inhibitors, allosteric regulation, or post-translational modification; the sensing of metabolites by riboswitches or post-transcriptional modification of RNA; and controlling gene expression.
The use of isotopes for labelling small molecules continues to be essential for analytical methodologies, from measuring metabolite concentrations in biological matrices and determining metabolite fluxes in biological organisms to the discovery of metabolic pathways [27,28,29]. Biocatalytic systems have therefore been important for the synthesis of isotope-labelled metabolites, whether through whole cell systems or isolated enzymes [30,31]. Synthesis has also been key for proving the correct molecular structure [32] and for the production of larger amounts of the pure metabolite [33,34]. While synthetic organic chemistry provides a significant repertoire of well-established reactions, safety, and health, environment and sustainability aspects have become of increasing importance in the industrial manufacturing processes of metabolites [35]. Improving resource and energy efficiency, reducing risks and the extensive use of natural resources, and avoiding the use of toxic chemicals are major planetary issues for sustainable development. Catalytic reactions provide novel, green, and sustainable methodologies for powerful and resource-efficient synthetic chemistry. This has also been convincingly demonstrated by recent Nobel Prizes in chemistry which were awarded in 2018 to Frances Arnold for directed evolution of enzymes [36] and in 2021 to Benjamin List [37] and David McMillan [38] for the development of asymmetric organocatalysis. Fundamental and mutually beneficial inspirations can originate from the interface and interactions between biocatalysis and organocatalysis [39]. Metabolic pathways used by biological organisms to prepare valuable metabolites from the raw materials available in their environment have also been successfully utilized and developed into industrial bioprocesses for manufacturing of metabolites, metabolite-like compounds, non-natural chemical entities, and metabolites thereof [40,41,42]. The metabolic pathways, which start from highly functionalized biobased raw materials instead of hydrocarbons, also provide inspirations for biocatalytic defunctionalization reactions in transitioning towards raw materials from bioresources [43]. However, the desired use of biobased raw materials also needs to consider other goals, such as biodiversity, sustainability, and supply chain issues, if biologically endangered and rare biological species are required or the amount and quality of the biobased raw material is variable and subject to various environmental factors [44]. Therefore, the molecular and engineering fundamentals of how nature achieves the biosynthesis of metabolites using biocatalytic reactions in microbes, plants, animals, and humans have attracted much interest as a blueprint for optimized biocatalytic systems of metabolite production [45].
The purpose of this work is to provide an overview of design and engineering approaches for biocatalytic systems in metabolite production and their application in manufacturing processes. The significance of biocatalytic systems for metabolite production is connected with the general strategic advantages of using biocatalysis in synthesis, such as their high selectivity and shortened synthetic routes [46].

2. Design and Engineering of Biocatalytic Systems

Biocatalytic system design and engineering towards the synthesis of metabolites starts with route selection and includes the preparation of suitable biocatalysts and raw materials, reaction engineering, process integration, intensification, and scaling of the selected metabolite manufacturing processes [35]. The main aim of this work focuses on designing and engineering biocatalytic systems in order to produce metabolites. Great progress has been made in delineating natural and engineering synthetic metabolic pathways, advanced methodologies and tools for finding and applying biocatalysts, and last but not least product recovery and purification. These significant advances along a whole bioprocess and workflow have brought biocatalytic systems into a privileged position. Biocatalytic systems are being used for producing not only natural metabolites, but also metabolite-like compounds; metabolites derived from the transformation of new chemical entities by biological organisms and isotope-labelled metabolites.
Having synthetic access to a metabolite or metabolite-like compound, or even more convenient, having an already available pure product, has been and continues to be highly important and relevant [47,48] for a number of reasons, as shown in Figure 1A for fundamental and applied sciences, as well as for a great variety of applications in industry and medicine, as shown in Figure 1B.
In view of millions of protein sequences, genomic enzymology tools are highly valuable for generating clues and hypotheses to guide experiments towards the correct assignment of enzyme and metabolic functions, as the functions of a significant fraction of protein sequences are unknown, uncertain, or even misassigned [49]. The final proof of the assignments is not possible without experimental verification of the predicted enzyme functions. This requires the availability of the corresponding metabolites (a) as enzyme substrates to demonstrate their conversion into the predicted product and to perform the enzyme activity assays, and (b) for the discovery of novel biological functions. In the case of chiral enzyme substrates or products, or for investigations of enzymatic reaction mechanisms, enantiopure metabolites can resolve fundamental questions and provide more detailed insights. Applications in industry are growing and include food supplements, pharmaceuticals, flavors, fragrances, cosmetics, dyes, and agrochemicals. The excellent chemo-, regio-, and stereoselectivity and mild reaction conditions of biocatalytic systems, which avoid protection–deprotection schemes, are environment friendly and safe to use. This is advantageous for the use of reactions requiring toxic chemicals, heavy metals, or the introduction and removal of protecting groups in chemocatalytic or stoichiometric reactions.
There are also disadvantages to overcome for biocatalytic systems when substrate and product concentrations are limited by solubility in aqueous media or enzyme inhibition. In these cases, a reaction type corresponding to a well-established key functionalization reaction in organic chemistry requires intense elaboration of a suitable biocatalytic systems for a selected substrate to product conversion due to the narrow substrate scope of the biocatalyst. Stability issues, for biocatalysts as well as for substrates and products, need to be checked, and in case of fragile biologically active metabolites, suitable operating windows for the reaction and workup must be selected. Improving energy and resource efficiency in manufacturing metabolites, such as reducing energy use and avoiding protection–deprotection schemes in lengthy chemical synthesis from fossil-based raw materials or minimizing biological waste in low-yield extractive procedures from biobased raw materials, is essential in building more resilient and sustainable manufacturing routes to metabolites. A resilient manufacturing route to a metabolite means a route with a robust and stable production process with the ability to effectively cope with changing boundary conditions, respond, and maintain reliable manufacturing. The molecular transformations catalyzed by enzymes and metabolic pathways in nature, as well as the great advances of biocatalysis, provide a blueprint and rich sources of knowledge for designing and engineering biocatalytic systems for the synthesis of metabolites.
Biocatalytic systems using whole cells have been developed into numerous fermentation processes as well as biotransformation processes at large industrial scales for the production of metabolites [50,51,52,53]. Fermentation processes use the cultivation of microbial whole cells in a fermenter containing the medium with the nutrients required for cell growth and product biosynthesis, followed by subsequent product recovery and purification, while biotransformation processes make use of microbial whole cells in a resting state for transforming an advanced intermediate to the product under physiological conditions. The vast knowledge base and the rich diversity of biocatalytic whole-cell systems have led to biocatalytic metabolite production through suitable growing or resting whole cells (see Figure 2). Metabolic engineering and synthetic biology enable improvements in titer, rate, and space–time yield of metabolite synthesis from the starting materials (a) by increasing the performance of the biosynthetic pathways to the metabolite, and (b) by deleting any biocatalytic degradation reaction of the final metabolite and of any metabolic intermediates. Biocatalytic whole-cell systems are also connected with a high degree of complexity, which can be reduced by using cell-free biocatalytic systems (see Figure 2) in different forms of purification, from crude cell-free extracts to isolated and purified enzymes. Whatever biocatalytic system is considered, bioprocess design and engineering need to address and optimize various parameters such as biocatalytic pathway selection, form and status of the biocatalysts, reaction engineering, downstream processing, and purification of the metabolites [35,54]. Cell-free biocatalytic systems have the advantage of reducing the bioprocess complexity through the absence of interfering and degrading enzymes or the removal of mass-transfer limitations for substrates and products [54].
While the step economy of a biosynthetic pathway is already considered in the design phase, possible improvements such as complexity reduction, intermediate purification steps, and the number of separate reactors needed are also taken into account by the degree of process integration involved in biocatalytic metabolite synthesis. Depending on how many reaction steps are needed for transforming the starting materials to the metabolites, one-step to multistep enzymatic reactions are developed, if possible in one pot. If all the enzymatic reactions of a biosynthetic pathway can be performed well, enzymatic total synthesis can be achieved [55].

3. Synthesis of Naturally Occurring Metabolites

The milestone discoveries that small molecules of life could not only be isolated from nature, but could also be synthesized in the laboratory from inorganic chemical precursors, started the new era of synthetic organic chemistry [56]. The synthesis of urea by Friedrich Wöhler in 1828 [57] or acetic acid by Hermann Kolbe in 1845 [58] were clear demonstrations that organic compounds, which are found and formed naturally in living organisms, could be prepared starting from inorganic materials. This sparked tremendous interest in synthetic organic chemistry, and impressive advances have been achieved in the art and science of total synthesis of numerous more complex natural products (see Figure 3), such as cholesterol and penicillin [59,60,61,62,63]. Total synthesis has not only been key to the final proof of structures and to correcting mistakenly assigned structures [64], but it has also been a key driver for novel synthetic methods in organic chemistry, which remain at its heart up to the present time [65].
The tools and methodologies used for broadly applicable synthetic reactions, either reactions which are already well established or newly discovered and emerging reactions from modern organic chemistry, have enabled the total synthesis of an impressive number of naturally occurring metabolites [66,67,68,69] to be synthesized from very simple structures without any stereogenic centers into the most complex metabolites (see Figure 3 and Figure 4). Synthesizing naturally occurring polycyclic metabolites, such as steroids, from simple building blocks has attracted much interest [66]. The first total synthesis of cholesterol was described by R.B. Woodward, whereby the methyl-3-ketoetio-allo-cholanate synthesized previously was converted via cholestan-3-ol, cholestane-3-one, and 4-cholesten-3-one into cholesterol [67,68]. With large amounts of industrially manufactured cholesterol extracted from animal-derived raw materials, new short and straightforward synthetic routes starting from non-animal-derived biobased raw materials are of much interest, such as the conversion of plant-based diosgenin to cholesterol in four steps [69]. Overcoming the great challenges of the chemical synthesis of vitamin B12, a microbial metabolite of high molecular complexity (see Figure 4), has not only resulted in many significant discoveries and novel methods on the road to that goal, but has led to the epochal milestone of its first total synthesis by the research groups of Alfred Eschenmoser in Zurich and R.B. Woodward in Cambridge [70,71]. The identification of the microbial enzymes and the aerobic and anaerobic enzymatic pathways to vitamin B12 in nature [72] have shown the power of biocatalytic total synthesis and are also of fundamental interest in the context of the origin of vitamin B12 and related compounds [73].
The tremendous task of establishing the complete stereochemistry and total synthesis of the marine natural product palytoxin (see Figure 5), which has a molar mass of 2680 and contains 63 stereogenic centers as well as four trans- and three cis-carbon-carbon double bonds, has been achieved by Yoshito Kishi and coworkers [74].
In this pioneering work and landmark achievement, eight building blocks were coupled in seven reactions to the fully protected palytoxin carboxylic acid containing 43 protecting groups in total, from which the palytoxin carboxylic acid was obtained in a 35% overall yield after removal of the protecting groups [74]. The palytoxin carboxylic acid was then treated with acetic acid in order to obtain the corresponding δ-lactone, which was then converted into the final palytoxin [74]. The development of new chemical reactions, such as the powerful Fe(III)-mediated coupling of catharanthine and vindoline to anhydrovinblastine, which may perhaps be also involved in the natural plant biosynthesis of vinblastine, has enabled the synthesis of vinblastine (see Figure 3) and related compounds in 8 to 13 reaction steps [75]. Total synthesis without using protecting groups is of significant interest for reducing the complexity, cost, and number of steps, as shown in the synthesis of marine natural products [76].
The milestone discoveries of living microbial whole cells [77], their biosynthetic capabilities, and the elucidation of the organic chemistry of the underlying biocatalytic reactions exerted by cell-free extracts [78] have started the era of synthetic biochemistry (see Table 2 for an overview of the natural metabolites covered in this review, with their respective pathways involved).
Although the biological formation of urea is known, the elucidation of urea biosynthesis in animals from ammonia and carbon dioxide [79,80] required major scientific breakthroughs. Early preparative applications of biocatalysts focused on particular reactions, such as stereoselective sugar oxidation through the use of microbial whole cells, stereoselective carbonyl reductions using baker’s yeast or alcohol dehydrogenases, kinetic resolutions, or desymmetrization reactions using hydrolases such as pig liver esterase (PLE).
With the tremendous development of recombinant enzymes and enzyme engineering, a range of biocatalytic reaction platforms have become the first choice, such as asymmetric ketone reductions catalyzed by recombinant ketoreductases at a large scale. Numerous bioprocesses have been developed to an industrial large scale (see Figure 6) for producing naturally occurring metabolites, such as citric acid [81] and other organic acids, steroids [82], beta-lactams and other antibiotics [83], L-carnitin and other amino acids [84,85], vitamins [86], and microbial metabolites such as anticancer and immunosuppressant drugs [87]. Bioprocesses for functionalized and modified steroids, steroidal intermediates, and metabolites provide shortened routes compared with chemical synthesis and have been well established in the pharmaceutical industry for decades [88]. 17β-estradiol has been obtained with greater than 99% diastereomeric excess and a 64.8% yield through stereoselective reduction of estrone using Saccharomyces cerevisiae whole cells [89]. Industrial bioprocesses for the efficient production of 4-androstene-3,17-dione and related metabolites from phytosterols at high substrate concentrations [90] make these metabolites attractive precursors for the sustainable production of steroids from plant-based raw materials [82]. Therefore, the selective biocatalytic reduction of 4-androstene-3,17-dione to testosterone has attracted much interest among the various biocatalytic routes to testosterone. From an efficient and complete conversion of 4-androstene-3,17-dione at a substrate concentration of 28.8 g L−1, the product testosterone has been obtained with a purity of greater than 97% in 10 h [91].
Of particular interest for steroid bioprocesses are biocatalysts, which are able to catalyze highly selective reactions such as the regio- and stereoselective biocatalytic hydroxylation of unique C(sp3)-H positions of the steroid backbone [92], for example, the 17α-hydroxylase, the 21-hydroxylase, and the 11β-hydroxylase in the conversion of progesterone to hydrocortisone [93]. Bioprocesses for the biocatalytic synthesis of steroidal compounds from simple carbon sources are emerging [94], and the biocatalytic cyclization of linear precursors to the steroid backbone catalyzed by cyclases in one step is noteworthy [95]. Cyclases are also involved in the biocatalytic synthesis of all vitamin E components; α-, β-, γ-, and δ-tocopherol; and α-, β-, γ-, and δ-tocotrienol. Additionally, bioprocesses using in vitro plant cell cultures are of much interest [96] for the preparation of enantiomerically pure vitamin E components such as (R,R,R)-α-tocopherol. A new route to vitamin E has been established at a large scale by combining biocatalytic and chemical routes, such as the chemical synthesis of α-tocopherol via isophytol from β-farnesene, which is manufactured through fermentation [97]. Route shortening is also of much interest for the synthesis of lipid mediators, and a stereoselective biocatalytic Baeyer–Villiger oxidation and a diastereoselective ketoreductase-catalyzed enone reduction have been key steps in a chemoenzymatic synthesis strategy for prostaglandins [98].
Natural metabolites continue to be highly important as sources for small molecule pharmaceuticals for treating diseases such as infectious diseases, cancer, cardiovascular diseases, diabetes, glaucoma, or multiple sclerosis [99], and large-scale manufacturing is key for adequate supply. Therefore, efficient and reliable bioprocesses for metabolites have been important to replace: (a) extraction processes from endangered biological species which may become extinct and which may give yields depending on various environmental factors, or (b) non-sustainable production procedures combining extraction from biological species and synthetic modifications. A sustainable plant cell fermentation process for producing the natural diterpenoid paclitaxel (registered trade name Taxol(R)) at an industrial large scale (see Figure 7) preserves Taxus plants and is able to provide the required amounts of this Taxus species metabolite, which the WHO lists as an essential medicine and is used in cancer treatment [100,101]. The cell culture medium contains a simple monosaccharide as a carbon source, one or more amino acids as a nitrogen source, and a silver ion or complex, jasmonic acid methyl ester, auxin-related growth regulators, and phenylpropanoid pathway inhibitors like 3,4-methylenedioxy-6-nitrocinnamic acid for enhancing the production of taxol [100]. Human health and the quality of life of a large number of people around the world have greatly benefitted from a long history of dedicated work to discover and develop small molecules, such as artemisinin and avermectins, formed by biological organisms as gifts from nature for their use as anti-infectives in the 20th century [102,103]. It is, however, not the time for complacency today, as the appearance of new infectious agents, human negligence, and economic boundary conditions require actions towards the discovery, development, and reliable production of novel anti-infectives [104,105,106].
Beyond the manufacturing of metabolites through bioprocesses at industrial large scales such as anticancer drugs, anti-infectives, and other pharmaceuticals for human health [83,87], biomanufacturing of a variety of other metabolites such as vitamins [86,107], flavors, and fragrances [108] has become increasingly attractive for various industrial sectors [109]. Fermentation using Pseudomonas denitrificans or Pseudomonas freudenreichii strains downstream and purification processes has been developed for the manufacturing of vitamin B12 at an industrial large scale [109]. The efficient enzymatic cyclization of (E,E)-homofarnesol, which can be produced from the fermentation product (E)-β-farnesene, to the fragrance ingredient (-)-ambrox catalyzed by engineered squalene hopene cyclase at an industrial scale (see Figure 8) represents a significant improvement in carbon efficiency and sustainability [108].
Another important area is the synthesis of metabolites for analytical or diagnostic applications, such as the use of metabolites in analytical or diagnostic devices, as standards, or for measuring enzyme activities. The synthesis of metabolites which act as ionophores has been of much interest for the analysis and monitoring of biomedically and environmentally relevant ions via ion-selective electrodes and sensors [110], for example, the biomanufacturing of highly pure nonactin as a neutral ionophore in monitoring ammonium ions [111,112].
Metabolites also need to be synthesized for analytical investigations involving the measurement of enzyme activities, such as the analysis of enzyme activities relevant to clinical chemistry, food analysis, enzymology, enzyme production, environment, verification or discovery of novel enzyme functions, development of enzyme inhibitors, or the analysis of activating, signaling, or regulatory functions.
Energy metabolism and glycolytic pathways are central to biological organisms, and the metabolites of the monosaccharide catabolic pathways are essential. It is therefore desirable to synthesize, in pure and stable form, the metabolites (see Figure 9) of the most common pathways for the breakdown of D-glucose, the Emden–Meyerhof–Parnas, the Entner–Doudoroff, and the pentose phosphate pathway, because these are central to kingdoms of life.
The enantiomerically pure metabolite D-glyceraldehyde-3-phosphate, which occurs in all the three most common D-glucose catabolic pathways, as well as the enantiomerically pure L-glyceraldehyde-3-phosphate, which is toxic to cells, have been synthesized from their corresponding aldehydes through biocatalytic ATP-dependent phosphorylation using enantiocomplementary kinases and phosphoenol pyruvate/pyruvatekinase for regenerating ATP. Glycerol kinase has been used in the enantioselective phosphorylation of L-glyceraldehyde [113,114], which could be prepared through glycerol dehydrogenase-catalyzed resolution of racemic glyceraldehyde [115,116]. Dihydroxyacetone kinase has been found as the corresponding enantiocomplementary enzyme to enantioselectively catalyze the phosphorylation of D-glyceraldehyde [117]. The enolase substrate D-glycerate-2-phosphate has been prepared by phosphorylating D-glycerate using recombinant glycerate-2-kinase, ATP as cofactor, and phosphoenolpyruvate/pyruvatekinase for ATP regeneration [118]. In the pentose phosphate pathway, D-xylulose-5-phosphate has been prepared through two different routes; either through transketolase-catalyzed condensation of D-glyceraldehyde-3-phosphate [119,120] with hydroxypyruvate, which serves as irreversible C2-donor, or through xylulokinase-catalyzed ATP-dependent phosphorylation of D-xylulose [121,122]. This latter approach has also been extended to the synthesis of L-xylulose-5-phosphate by using an enantiocomplementary xylulokinase [121]. In the D-tagatose catabolic pathways, D-tagatose-6-phosphate 1-kinase-catalyzed phosphorylation of D-tagatose-6-phophate enabled the preparation of the central metabolite D-tagatose-1,6-diphosphate [122]. A characteristic metabolite for the Entner–Doudoroff pathway is 2-keto-3-deoxy-6-phosphogluconate, which can be synthesized in one step by eliminating water from 6-phosphogluconate using 6-phosphogluconate dehydratase [123]. In a similar way, metabolites of other monosaccharide non-phosphorylative catabolic pathways can be synthesized in a straightforward way from the corresponding sugar acid, such as 2-keto-3-deoxy-D-galactonate from D-galactonate or 2-keto-3-deoxy-D-xylonate from D-xylonate using D-xylonate dehydratase [124]. D-gluconate dehydratase-catalyzed water elimination from D-gluconate allows for the straightforward preparation of 2-keto-3-deoxy-D-gluconate [125]. In energy metabolism, the high energy of the phosphorus–nitrogen bond in phosphagens is a key energy source. Selective biocatalytic synthesis enables straightforward access, such as in the one step synthesis of Nω-phospho-L-arginine [126].
Biocatalytic methods are also very useful in synthesizing a number of key metabolites from other metabolic pathways. Shikimic acid-3-phosphate can be prepared through shikimate kinase-catalyzed ATP-dependent phosphorylation of shikimic acid and ATP regeneration using phosphoenolpyruvate and pyruvate kinase [127]. Pyridoxamine-5’-phosphate was synthesized from pyridoxal-5-phosphate through biocatalytic transamination using an ω-transaminase [128]. The biocatalytic L-arginine addition reaction to fumaric acid enabled efficient one-step access to the urea cycle metabolite L-argininosuccinate [129,130].
Biocatalytic methods have also been of much interest for the synthesis of vitamin D metabolites. Highly selective side-chain hydroxylation of vitamin D3 in the 25-position has been achieved at a laboratory scale using different biocatalytic approaches, such as cytochrome P450 monooxygenases, complex electron donors, and oxygen in whole-cell systems [131,132], through hydrogen peroxide-dependent peroxygenase [133], or through ferricyanide-dependent biocatalytic hydroxylation using a vitamin D3 hydroxylase as cell-free extract or as purified enzyme from Sterolibacterium denitrificans [134]. Efficient biocatalytic production of 573 mg of 25-hydroxyvitamin D3 per liter has been achieved using nisin-treated cells of Rhodococcus erythropolis containing an engineered vitamin D3 hydroxylase from Pseudonocardia autotrophica (573 mg of 25-hydroxyvitamin D3 per liter within 2 h) [131], and by using a Bacillus cereus strain (830 mg of 25-hydroxyvitamin D3 per liter within 60 h) [132]. A facile ferricyanide-dependent hydroxylation using a vitamin D3 hydroxylase, either as a cell-free extract or as purified enzyme from Sterolibacterium denitrificans, has enabled a simplified preparation of 25-hydroxyvitamin D3 in a yield greater than 99% at a 1 mM substrate concentration [134]. The biocatalytic synthesis of 1α,25-dihydroxyvitamin D3 (calcitriol) from vitamin D3 has been achieved through double hydroxylation of vitamin D3 using whole cells of Pseudonocardia sp., with more than 30% yield and a titer of approximately 62 mg L−1 [135]. Biocatalytic hydroxylation of 25-hydroxyvitamin D3 catalyzed by 25-hydroxyvitamin D3 24-hydroxylase and formation of 24R,25-dihydroxyvitamin D3 have been demonstrated on an analytical scale [136].
The 25-hydroxyvitamin D2 has been obtained as the sole product in a 90% yield through the regioselective hydroxylation of vitamin D2 catalyzed by a peroxygenase from Coprinopsis cinerea [133]. An engineered triple variant of CYP105A1 with increased 1α-hydroxylase activity has enabled the formation of 1α,25-dihydroxyvitamin D2, while an engineered double variant of CYP105A1 showed increased 26-hydroxylase activity and was best for the formation of 25,26-dihydroxyvitamin D2 [137]. Whole cells of Bacillus megaterium expressing the highly selective vitamin D2 hydroxylase CYP109E1 were used for the biocatalytic two-step hydroxylation of vitamin D2, whereby a titer of 12 mg L−1 of 24R,25-dihydroxyvitamin D2 was obtained in 48 h [138]. Simple biocatalytic routes are also of special interest for disease-specific metabolites in order to support and simplify the diagnostics, for example, of inborn errors of metabolism, cancers, and cardiovascular and metabolic diseases [139,140,141].

4. Synthesis of Isotope-Labelled Metabolites

The use of radioactive isotopes such as 3H (tritium), 14C, or 32P has been instrumental for the discovery of major metabolic pathways, such as the path of carbon in photosynthesis [142]. Biocatalytic methods, which have been developed for the synthesis of metabolites labelled with a radioactive isotope at a specific position, such as tritium- or 14C-labelled NAD+ or 14C-labelled nicotinamide riboside [143,144], can also be translated to methods for the synthesis of the corresponding metabolites labelled with a stable isotope at a specific position, such as 13C-labelled NAD+ or 13C-labelled nicotinamide riboside [144]. In contrast to radioactive labels, working with stable isotope labels such as the biogenic isotopes 2H (deuterium), 13C, 15N, or 18O does not involve any health hazards and is not subject to regulations regarding radiation safety. The technology of isotope separation has enabled a continuous increase in the production of stable isotopes of light elements [145]. Compounds in which an atom like 1H, 12C, 14N, or 16O is replaced by a corresponding isotope with a higher atomic mass are of significant interest to numerous applications, because the chemical structure and physical properties remain unchanged. Major types of applications of labelling with stable isotopes, such as 2H, 13C, or 15N in stable bonds, in which the label is non-exchangeable under physiological conditions, are related to biological cells in both health and disease. These applications include quantification methods for specific metabolites, methods for analyzing metabolic fluxes and pathways, and the localization of metabolites through imaging methods. For quantification methods, which are important in diagnostics, stable isotope labelled metabolites, drugs, or metabolite like molecules are preferable. Alternatively, other approaches may be followed, like chemically labelling the unlabeled analytes using derivation reagents containing stable isotopes or employing quantitative NMR of native metabolites [146]. For stable isotope tracer methods, the analysis of carbon metabolic fluxes and pathways [147] and stable isotope resolved metabolomics [146]. This is important for analyzing disease-specific metabolic pathway alterations, such as cancer cell metabolism [148]. For these methods, as well as for imaging methods of biological and pathological processes [149,150], the 13C-labelled precursors or nutrients such as universally 13C-labelled D-glucose can be used for in vivo labelling. Metabolites labelled with an equal number of stable isotope atoms but at different positions can be distinguished using NMR as isotopomers. Metabolites which differ by their isotope number and composition can be distinguished through MS as isotopologues [146].
The impressive advances of highly sensitive MS and NMR instrumentation, with their powerful methodologies and analyses with high information content, have shifted the interest to the use of the stable isotopes 2H, 13C, 15N, or 18O [151,152]. As isotope separation is demanding and requires highly specialized equipment and facilities for the production of stable isotope-labelled starting chemicals with high chemical and isotopic purity, the precious stable isotope-labelled starting materials should then be fully utilized for the synthesis of the desired metabolites. Therefore, highly selective synthetic methods are needed, which are able to efficiently incorporate to a high degree the stable isotope from the starting material into a defined position of the product and to completely convert the starting material to the target metabolite. The thermodynamics of biochemical reactions and the universe of biocatalysts provide a significant knowledge base from which suitable biocatalytic reactions for selective labelling with stable isotopes can be selected (see Figure 10).
Biocatalytic methods for site- and stereoselective deuteration are of much interest for short routes to deuterated metabolites, for example, in the synthesis of selectively deuterated phosphatidyl-sn-glycerol, amino acids deuterated in the α- and/or β-position, deuterated NAD+/ NADH cofactors, or deuterated aldehydes [153,154,155,156,157]. The biocatalytic synthesis of 3′,4′,5′,5′-tetradeuterated 5-phospho-D-ribosyl α-1-pyrophosphate (PRPP) was prepared from 3′,4′,5′,5′-tetradeuterated D-ribose through ribokinase-catalyzed phosphorylation and PRPP synthetase-catalyzed pyrophosphorylation [158]. The tetradeuterated PRPP was then converted through multistep enzymatic processes in high yields to the 3′,4′,5′,5′-tetradeuterated nucleotides ATP, CTP, GTP, and UTP [158].
The synthesis of 13C-labelled metabolites has been very useful for various metabolomics applications, such as for growing cells on media containing 13C-labelled carbon sources as nutrients [159], for detailed investigations of cellular metabolism and the functional properties of complex metabolic networks through 13C-based metabolic flux analysis [27,160], or for overcoming matrix effects in accurate and reliable metabolite analyses and quantitative metabolomics [161]. The key to these applications and the discovery of novel biosynthetic pathways like the deoxyxylulose phosphate pathway [162] has been the synthesis of 13C-labelled biochemicals, such as 13C-labelled acetate or isotope isomers (isotopomers) of D-glucose, where specific 12C-atoms are replaced by their 13C-isotopes. Enzymatic methods have facilitated access to 13C-labelled metabolites, such as monosaccharides [163], as well as 13C-labelled amino acids [164,165,166,167]. The use of very simple and inexpensive 13C-labelled precursors is thereby attractive, such as 13C-labelled pyruvate for the enzymatic synthesis of 13C-labelled aromatic amino acids [168]. Direct utilization of 13carbon dioxide is not only of significant interest for investigating the metabolism of photosynthetic organisms, but also for the biocatalytic synthesis of 13C-labelled biochemicals, such as 13C-labelled L-malate [169].
As many natural products contain nitrogen, the introduction of the stable nitrogen isotope 15N is very useful for discovering natural products and characterizing their biosynthetic pathways and metabolic intermediates [170]. Biocatalytic synthesis of 15N-labelled metabolites has been achieved in a straightforward way by introducing a stable isotope from 15NH4 salts, by using biocatalytic systems with isolated enzymes, or by making use of biosynthesis in whole cells growing in media containing 15NH4 salts. A very efficient NAD+-dependent amino acid dehydrogenase-catalyzed preparation has been demonstrated through the synthesis of 15N-labelled L-serine, L-methionine, and L-glutamic acid from the corresponding α-keto acids using alanine dehydrogenase, leucine dehydrogenase, and glutamate dehydrogenase, respectively, whereby NADH regeneration was performed using the glucose/glucose dehydrogenase system [171]. The four 15N-labelled cobalamin standards hydroxocobalamin, adenosylcobalamin, methylcobalamin, and cyanocobalamin have been prepared through biosynthesis by growing Propionibacterium freudenreichii whole cells in a chemically defined medium containing (15NH4)2SO4 instead of (14NH4)2SO4 [172].
In addition to using a single type of stable isotope for labelling, biocatalytic synthesis has also been attractive for the introduction of more than one type of stable isotope. Pentose phosphate and purine pathway enzymes, together with biocatalytic regeneration cycles for nucleoside triphosphate, folate, aspartate, glutamine, and NAD+, have been utilized for labelling purine nucleotides with 13C and 15N [29]. This flexible and robust one-pot biocatalytic system enabled the preparation of uniformly 13C- and/or 15N-labelled GTP, or 13C-labelled ATP in the C2- and C8-position from labelled serine, ammonium, glucose, and carbon dioxide [30].

5. Synthesis of Pharmaceutical Drug Metabolites

The development of new molecular entities for the effective treatment of human diseases with minimized side effects requires an understanding of its interactions with the biological cells of humans and their microbiome. The investigation of potential in vitro and in vivo pharmaceutical drug metabolism involving human and microbial enzymes, the biocatalytic reactions converting administered pharmaceutical drugs to derived metabolites, and the identity and biological activity of pharmaceutical drug metabolites, are of key importance to the treatment response, drug safety, and side effects. Knowledge about pharmaceutical drug metabolism reactions, data on drug metabolizing enzymes, and structures of drug metabolites has been growing significantly over the past years, as shown by their increasing numbers in the DrugBank database and its most recent version DrugBank 5.0 [173]. The complexity of pharmaceutical drug metabolism is increased further because pharmaceutical drugs not only interact with human metabolism but also with the human microbiome [174], as demonstrated by the biotransformation capabilities of the human gut microbiome towards numerous pharmaceutical drugs [175,176,177].
When applications of a pharmaceutical drug candidate show that it is enzymatically converted, from the site where it is administered to the desired drug action site, to a less active drug metabolite, this results in a poor treatment response and requires further drug development. For optimizing drug effectiveness while minimizing side effects, ensuring correct dosing, and avoiding drug overdoses of therapeutics with undesirable side effects, therapeutic drug monitoring has become an important tool for precision medicine, where isotope-labelled drugs are routinely used as reference standards for LC-MS/MS-analyses in clinical chemistry. Although pharmaceutical drug metabolism often leads to inactivation, there are also cases where therapeutic benefits may derive from biotransformation to a pharmacologically more active metabolite [178], for example, when a prodrug with better cell permeability is enzymatically converted to the active pharmaceutical in diseased cells. Pharmacologically inactive, or less active by three orders of magnitude, small molecular weight compounds, which are enzymatically converted in vivo to their active pharmaceuticals, have been developed through different paths. They have been discovered by chance, from rescuing a drug discovery project, or by designing a prodrug [179]. Chemically reactive or toxic drug metabolites have received increased attention, and the investigation of potential side effects, safety, and toxicity issues of pharmaceutical drug metabolites, which may potentially be formed through biotransformations in the human body, has also evolved with the “Metabolites In Safety Testing” (MIST) guidance and the framework for identifying, quantifying, and assessing human drug metabolite safety [180,181,182]. The investigations of the possible effects of such modified drugs require sufficient amounts of pure drug metabolites; therefore, straightforward methods for their synthetic access are highly desirable and can be highly significant for the timeline of projects. As selective chemical modification of complex drugs with stereocenters may be challenging [183], using biocatalysts which are involved in human drug metabolism provides an attractive selective approach [184].
When the orally active synthetic pharmaceutical drug dydrogesterone is used for treating progesterone deficiency and various gynecological conditions, human metabolism is responsible for the formation of the drug metabolite 20α-dihydrodydrogesterone. This drug metabolite has been prepared through the efficient stereo- and regioselective reduction of dydrogesterone (see Figure 11) catalyzed by recombinant human 20α-hydroxysteroid dehydrogenase AKR1C1 expressed in Schizosaccharomyces pombe [185]. With chemical reduction of the C20-keto group in dydrogesterone leading only to the 20β-dihydrodydrogesterone, biocatalytic reduction is key for obtaining 20α-dihydrodydrogesterone, which is also pharmacologically active [185]. The drug metabolite (S)-fingolimod-phosphate, which is a modulator of sphingosine 1-phosphate receptor 1, is formed in vivo through sphingosine kinase 2-catalyzed phosphorylation (see Figure 6) of the drug fingolimod, which has been approved as a pharmaceutical drug for the therapy of multiple sclerosis in more than 80 countries [186]. As different forms of mycophenolic acid have various therapeutic applications as pharmaceutical drugs, for example, as immunosuppressants, their pharmacologically active metabolite, mycophenolic acid acylglucuronide has attracted interest, as it also inhibits inosine monophosphate dehydrogenase II like mycophenolic acid [187]. For the biocatalytic synthesis of the acylglucuronide of mycophenolic acid, only horse liver homogenate was found to catalyze the glucuronidation of mycophenolic acid, using UDP-glucuronic acid as donor, but the acylglucuronide was formed in a 1:1 mixture with the 7-O-glucuronide [187]. Through the optimization of the reaction temperature, the concentrations of the liver homogenate, and the UDP-glucuronide, the degree of conversion was increased to 54% and an acylglucuronide to 7-O-glucuronide ratio of 4.9:1 could be obtained, leading to the drug metabolite mycophenolic acid acylglucuronide (see Figure 11) in a >95% purity and a 34% isolated yield [187].
Biocatalytic transformations of synthetic pharmaceutical drug derivatives through human metabolism in vivo are important aspects in the design of prodrugs in order to overcome barriers in delivering and releasing the parent drug, and to improve cell permeability and water solubility [188].
Although many drug metabolites are pharmacologically inactive or much less active than their parent drugs, as exemplified by the pharmacologically inactive 7-O-glucuronide of mycophenolic acid as the major mycophenolic acid metabolite in humans, ensuring synthetic access is important [189]. The 7-O-glucuronide of mycophenolic was obtained with 97% purity through biocatalytic glucuronidation of mycophenolic acid using horse liver homogenate as the biocatalyst and UDP-glucuronic acid as the donor [187].
Chemically reactive drug metabolites, which are formed by human metabolism, can react with relevant molecules of biological cells and lead to functional changes and adverse drug reactions [190]. In the drug metabolism of the nonsteroid anti-inflammatory drug diclofenac, which is used for treating rheumatoid disorders, the two metabolic pathways of the cytochrome P450-catalyzed oxidation to the corresponding chemically reactive quinone imine metabolites and the enzymatic glucuronidation are thought to be involved in several adverse drug reactions [191,192]. Cytochrome P450 enzymes have been shown to catalyze the drug conversion to chemically reactive metabolites, which can cause toxic effects, such as the oxidation of acetaminophen to the toxic metabolite N-acetyl-p-benzoquinone imine catalyzed by human cytochrome P450 2E1, 1A2, and 3A4 [192,193,194].

6. Synthesis of Metabolite-like Compounds

The concepts of natural product-likeness [195,196] and metabolite-likeness [197,198] are attractive because the transport of natural products, nutrients, and metabolites is omnipresent in biological organisms. Natural-product-like and metabolite-like structures have been found to be present in a significant number of approved pharmaceutical drugs [99,199].
Variations in metabolite and natural product structures have attracted increasing interest due to the functional group differences which have been found to exist between natural products and synthetic molecules [200], the changing structural characteristics and properties of approved pharmaceutical drugs over time [201,202], and the interactions between metabolites and natural products with target proteins, including biosynthetic and transport proteins [203]. With the increasing knowledge about specific interactions between metabolites of the human microbiome and human disease-relevant biomolecules such as ligand–receptor or enzyme inhibitor–enzyme interactions, the concept of utilizing microbial metabolites as rich molecular spaces is very attractive for discovering metabolite-like compounds as novel pharmaceutical drugs for highly precise therapies and the derisking of adverse drug reactions [204,205]. The development of metabolite-like compounds from bacterial tryptophan metabolism for discovering novel ligands which bind directly to the pregnane X receptor and are not cytotoxic is of much interest for further development and opens up a wide range of opportunities [206]. Excellent biocatalytic tools and methodologies have been developed for synthesizing non-natural chiral amino acids with high enantioselectivity, and the diversity of possible functional groups and chiral centers in non-natural amino acids is attractive for their use as fragments in small molecule pharmaceutical drugs (for some examples, see Figure 12) for the therapy of diseases [207,208,209,210].
Biocatalytic reactions for highly stereoselective formation of carbon–carbon bonds is of central importance for the synthesis of both natural metabolites as well as metabolite-like compounds (see Figure 12), but practical applications are not very common [208]. The thermodynamic advantage of decarboxylative carbon–carbon bond formation, as discussed previously in the context of transketolase-catalyzed reactions, has been demonstrated by reactions catalyzed by the enzyme UstD and its evolved variants [210]. This leads to the stereoselective formation of a series of γ-hydroxy-α-amino acids from various aldehydes and L-aspartic acid with the release of carbon dioxide (see example in the middle of Figure 12). The thermodynamic equilibrium of a stereoselective aldol reaction catalyzed by an engineered deoxyribose 5-phosphate aldolase has been overcome by coupling it to the subsequent reactions in the biocatalytic synthetic route to islatravir (see example at the bottom of Figure 12) and removing the inorganic phosphate side product [211]. Numerous advances have been achieved in recent years in stereoselective alkylations, acylations, oxidative carbon–carbon coupling reactions, cyclizations, and carbene transfer reactions [212]. The synthesis of various pyrroloindolines has been achieved through the regio- and stereoselective methylation of various indoles at their C3-position when catalyzed by the S-adenosyl-L-methionine (SAM)-dependent methyltransferase PsmD from Streptomyces griseofuscus, whereby SAM was regenerated by methyliodide and halide methyltransferase from Chloracidobacterium thermophilum [213]. Recombinant trans-α-hydroxybenzylidenepyruvate hydratase-aldolase (NahE) has been demonstrated to catalyze stereoselective Michael addition reactions of pyruvate to various β-nitrostyrenes, from which then the corresponding β-aryl-γ-nitrobutyric acids can be obtained [214].
The advances in the characterization and engineering of enzymes catalyzing routine reactions in organic chemistry laboratories, such as Diels–Alder reactions, Claisen, and Cope rearrangements are very promising for translating biocatalytic complexity-generating reactions into routine operations [215]. The development of highly active Diels-Alderases from nature, through engineering, and by design for catalyzing highly stereoselective intermolecular [4+2] cycloaddition reactions is of great interest for the synthesis of natural products [216,217] and for catalyzing abiological hetero-Diels–Alder reactions [218]. Their versatility has been increased by reversing the exo-selectivity of natural Diels-Alderases to the endo-selectivity of an engineered Diels-Alderase, catalyzing the intermolecular [4+2] cycloaddition of the same substrates with high enantioselectivity and broadening the scope of the dienes and dienophiles accepted as substrates [219].
Combining biocatalytic and chemical reaction steps in chemoenzymatic syntheses offers new opportunities for overcoming challenges and diversifying metabolite structures. Advances have been achieved in the synthesis of diversified compounds of the plant metabolite cis-(+)-12-oxophytodienoic acid [220], a variety of non-natural nucleosides and nucleoside building blocks (see Figure 12), as well as in the synthesis of nucleoside analogue drugs through enzymatic cascades [221,222].

7. Discussion

The growing number of metabolite structures identified in biological organisms and the renewed interest in the elucidation of the fundamental roles and useful properties of metabolites have also increased the necessity of developing analytical and synthetic methods and tools for a great diversity of metabolites. The identification of natural metabolic pathways has provided great starting points for the development of straightforward biocatalytic synthesis routes inspired by nature. Biocatalytic systems offer particular benefits not only for producing naturally occurring metabolites but also for stable isotope-labelled metabolites, drug metabolites, and non-natural metabolite-like compounds. The molecular economy and reduced complexity of highly selective and protecting group-free biocatalytic reactions have enabled resource-efficient and robust production procedures for these metabolite classes. When the extraction yields of naturally occurring metabolites from biological resources are rather low and quite variable, or if challenging and lengthy chemical routes using protection–deprotection schemes require significant purification steps, biocatalytic production procedures using recombinant biocatalysts are a preferred choice. The strategic advantages of biocatalytic systems enable the fast generation of molecular complexity through shorter synthetic routes in a straightforward way, as demonstrated by the development of the biocatalytic synthesis of the antiviral compound islatravir. The cascade of the three biocatalytic reaction steps leading to a single stereoisomer of islatravir from achiral 2-ethinylglycerol (see Figure 12) shortens the number of reaction steps to less than half of the reaction steps reported in previous routes [211].
Engineering enzymes towards desired activities and selectivities under certain reaction conditions and the combination of the necessary enzymes in a cascade reaction enable us to overcome unfavorable thermodynamic equilibria, enzyme inhibition, and the isolation of intermediates. Minimizing the time for the development, manufacturing, and supply of a compound in high demand and urgently needed is always very critical, but it has been even more crucial during the COVID-19 pandemic. The fast development of a short scalable biocatalytic route to the antiviral agent Molnupiravir using engineered ribosyl-1-kinase and uridine phosphorylase as key enzymes was impressive [223]. The advantages of utilizing a biocatalytic reaction cascade for preparing complex non-natural small molecules in one pot have also been demonstrated through the efficient biocatalytic synthesis of the cyclic dinucleotide MK-1454 using three engineered kinases and an engineered cyclic guanosine–adenosine synthase as key enzymes [224]. An isolated yield of 62%, based on the starting material of the nucleotide monothiophosphates, has been achieved for a single diastereomer of MK-1454 without the use of any protecting group [224]. When manufacturing at larger scale is needed due to increasing demand, raw materials, resource efficiency, and reliability issues have become increasingly important. Sustainability benefits can also be achieved when biological resources which are rare or in danger of becoming extinct are preserved and abundant bio-based resources are used for metabolite production using biocatalytic systems.
Excellent opportunities appear for exploring uncharted territory regarding novel metabolites and their biological functions [225], natural product drugs [226,227], and metabolite-like compounds. Discovering and characterizing novel enzyme functions and pathways from nature, as well as engineering and evolving enzymes which catalyze reactions that are new to nature [228,229], are important for extending the frontiers of biocatalytic reactions in syntheses. Unlocking the power of enzymes can transform the synthesis of metabolites, natural products, and non-natural small molecules derived thereof in various ways, from individual enzymatic reactions in chemoenzymatic synthesis to their full utilization in enzymatic total synthesis [230,231,232].

8. Future Directions

With the rising gap between the number of known metabolites and the number of metabolites whose syntheses have been reported, there are more than enough concrete problems to be solved, such as the synthesis of important metabolites with high biological activity and inherent instability, or biologically active metabolites which are synthesized by bifunctional enzymes carrying synthetic and degrading functions within the same protein. Rapid advances in enzyme engineering and computational methodologies have been outlined and appear promising on the road towards the design of robust enzymes with desired enzyme functional properties [233]. Many roads can lead to discovering novel enzymes from nature, such as guiding the correct assignment of enzyme functions to gene annotations and domains of unknown functions on genomes [234,235,236], unlocking natural product biosynthetic enzymes from metagenomes [237], or identifying missing enzymes in biosynthetic pathways to metabolites [238]. Mining microbial genomes for biosynthetic gene clusters [239] and deciphering precise genome–metabolome relationships of bacteria and fungi are very promising approaches for finding novel biosynthetic enzymes and pathways to novel metabolites [240,241,242,243]. Advanced metabolic engineering and synthetic biology tools and methodologies [244] and efficient gene expression to highly functional and fit-for-use enzymes are key for engineering novel biocatalytic pathways in viable and sustainable production processes. Therefore, the future of synthesizing naturally occurring metabolites [245,246], metabolite-like compounds, drug metabolites, and stable isotope-labelled metabolites using biocatalytic systems looks bright.
Table 2. Overview of Natural Metabolites covered in this review.
Table 2. Overview of Natural Metabolites covered in this review.
MetabolitesMetabolic PathwayReference
Acetic acidCarbon metabolism[58]
(-)-AmbroxHomofarnesol cyclization[108]
4-Androstene-3,17-dioneSteroid biosynthesis[90]
L-Argininosuccinate Urea cycle[129,130]
ArtemisininArtemisinin bioynthesis[103]
AvermectinPolyketide biosynthesis[102]
AzadirachtinTetranortriterpenoid biosynthesis[61]
CholesterolSteroid biosynthesis[67,69]
Citric acidCitrate cycle (TCA cycle)[81]
1α,25-Dihydroxyvitamin D2 Steroid biosynthesis[137]
1α,25-Dihydroxyvitamin D3 Steroid biosynthesis[135]
24R,25-Dihydroxyvitamin D2Steroid biosynthesis[138]
24R,25-Dihydroxyvitamin D3Steroid biosynthesis[136]
17β-EstradiolSteroid biosynthesis[89]
L-GlyceraldehydePentose and glucuronate interconversions[115,116]
D-Glyceraldehyde-3-phosphateEmbden-Meyerhof-Parnas pathway
Carbon fixation in photosynthesis
Pentose phosphate pathway
[117]
L-Glyceraldehyde-3-phosphateIsomerase bypass[113,114]
D-Glycerate-2-phosphate Embden-Meyerhof-Parnas pathway[118]
25-Hydroxyvitamin D2Steroid biosynthesis[133]
25-Hydroxyvitamin D3Steroid biosynthesis[131,132,133,134]
2-Keto-3-deoxy-D-galactonate Galactose metabolism[124]
2-Keto-3-deoxy-D-gluconate Non-phosphorylative Entner-Doudoroff pathway Pentose phosphate pathway[125]
2-Keto-3-deoxy-6-phosphogluconateEntner-Doudoroff pathway
Pentose phosphate pathway
[123]
2-Keto-3-deoxy-D-xylonatePentose and glucuronate interconversions[124]
PalytoxinPalytoxin biosynthesis[74]
Penicillin VPenicillin biosynthesis[59]
Nω-Phospho-L-arginine Phosphagen pathway[126]
Pyridoxamine-5’-phosphateVitamin B6 metabolism[128]
Shikimic acid-3-phosphate Shikimate pathway[127]
D-Tagatose-1,6-diphosphate Galactose metabolism
D-Tagatose pathway
[122]
TaxolTaxol biosynthesis[63,100,101]
TestosteroneSteroid biosynthesis[91]
UreaUrea cycle[57,79,80]
VinblastineIndole alkaloid biosynthesis[75,238]
Vitamin B12Cobalamine biosynthesis[70,71,72,73,107]
D-Xylulose-5-phosphate Pentose phosphate pathway[119,120,121]
L-Xylulose-5-phosphate Pentose and glucuronate interconversions[121]

Funding

This research received no external funding.

Acknowledgments

The author thanks all the reviewers for their work and comments.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Schreiber, S.L. Small molecules: The missing link in the central dogma. Nat. Chem. Biol. 2005, 1, 64–66. [Google Scholar] [CrossRef]
  2. McKnight, S.L. Back to the future: Molecular biology meets metabolism. Cold Spring Harb. Symp. Quant. Biol. 2011, 76, 403–411. [Google Scholar] [CrossRef] [PubMed]
  3. Aharoni, A.; Goodacre, R.; Fernie, A.R. Plant and microbial sciences as key drivers in the development of metabolomics research. Proc. Natl. Acad. Sci. USA 2023, 120, e2217383120. [Google Scholar] [CrossRef] [PubMed]
  4. Kell, D.B.; Oliver, S. The metabolome 18 years on: A concept comes of age. Metabolomics 2016, 12, 148. [Google Scholar] [CrossRef]
  5. Fiehn, O. Metabolomics by Gas Chromatography-Mass Spectrometry: The combination of targeted and untargeted profiling. Curr. Protoc. Mol. Biol. 2017, 114, 30.4.1–30.4.32. [Google Scholar] [CrossRef] [PubMed]
  6. Plumb, R.S.; Gethings, L.A.; Rainville, P.D.; Isaac, G.; Trengove, R.; King, A.M.; Wilson, I.D. Advances in high throughput LC/MS based metabolomics: A review. Trends Anal. Chem. 2023, 160, 116954. [Google Scholar] [CrossRef]
  7. Emwas, A.H.; Roy, R.; McKay, R.T.; Tenori, L.; Saccenti, E.; Gowda, G.N.; Raftery, D.; Alahmari, F.; Jaremko, L.; Jaremko, M.; et al. NMR spectroscopy for metabolomics research. Metabolites 2019, 9, 123. [Google Scholar] [CrossRef]
  8. Guijas, C.; Montenegro-Burke, J.R.; Domingo-Almenara, X.; Palermo, A.; Warth, B.; Hermann, G.; Koellensperger, G.; Huan, T.; Uritboonthai, W.; Aisporna, A.E.; et al. METLIN: A Technology Platform for Identifying Knowns and Unknowns. Anal. Chem. 2018, 90, 3156–3164. [Google Scholar] [CrossRef]
  9. Wishart, D.S.; Cheng, L.L.; Copié, V.; Edison, A.S.; Eghbalnia, H.R.; Hoch, J.C.; Gouveia, G.J.; Pathmasiri, W.; Powers, R.; Schock, T.B.; et al. NMR and metabolomics—A roadmap for the future. Metabolites 2022, 12, 678. [Google Scholar] [CrossRef]
  10. Hoch, J.C.; Baskaran, K.; Burr, H.; Chin, J.; Eghbalnia, H.R.; Fujiwara, T.; Gryk, M.R.; Iwata, T.; Kojima, C.; Kurisu, G.; et al. Biological Magnetic Resonance Data Bank. Nucleic Acids Res. 2023, 51, D368–D376. [Google Scholar] [CrossRef]
  11. Kanehisa, M.; Goto, S.; Sato, Y.; Kawashima, M.; Furumichi, M.; Tanabe, M. Data, information, knowledge and principle: Back to metabolism in KEGG. Nucleic Acids Res. 2014, 42, D199–D205. [Google Scholar] [CrossRef]
  12. Hastings, J.; Owen, G.; Dekker, A.; Ennis, M.; Kale, N.; Muthukrishnan, V.; Turner, S.; Swainston, N.; Mendes, P.; Steinbeck, C. ChEBI in 2016: Improved services and an expanding collection of metabolites. Nucleic Acids Res. 2016, 44, D1214–D1219. [Google Scholar] [CrossRef] [PubMed]
  13. van Santen, J.A.; Jacob, G.; Singh, A.L.; Aniebok, V.; Balunas, M.J.; Bunsko, D.; Neto, F.C.; Castaño-Espriu, L.; Chang, C.; Clark, T.N.; et al. The Natural Products Atlas: An Open Access Knowledge Base for Microbial Natural Products Discovery. ACS Cent. Sci. 2019, 5, 1824–1833. [Google Scholar] [CrossRef] [PubMed]
  14. Wishart, D.S.; Guo, A.; Oler, E.; Wang, F.; Anjum, A.; Peters, H.; Dizon, R.; Sayeeda, Z.; Tian, S.; Lee, B.L.; et al. HMDB 5.0: The human metabolome database for 2022. Nucleic Acids Res. 2022, 50, D622–D631. [Google Scholar] [CrossRef] [PubMed]
  15. Wishart, D.S.; Oler, E.; Peters, H.; Guo, A.; Girod, S.; Han, S.; Saha, S.; Lui, V.W.; LeVatte, M.; Gautam, V.; et al. MiMeDB: The Human Microbial Metabolome Database. Nucleic Acids Res. 2023, 51, D611–D620. [Google Scholar] [CrossRef]
  16. Sajed, T.; Marcu, A.; Ramirez, M.; Pon, A.; Guo, A.; Knox, C.; Wilson, M.; Grant, J.; Djoumbou, Y.; Wishart, D. ECMDB 2.0: A richer resource for understanding the biochemistry of E. coli. Nucleic Acids Res. 2016, 44, D495–D501. [Google Scholar] [CrossRef]
  17. Huang, W.; Luke, K.; Brewer, L.K.; Jace, W.; Jones, J.W.; Angela, T.; Nguyen, A.T.; Ana Marcu, A.; David, S.; Wishart, D.S.; et al. PAMDB: A comprehensive Pseudomonas aeruginosa metabolome database. Nucleic Acids Res. 2018, 46, D575–D580. [Google Scholar] [CrossRef] [PubMed]
  18. Moumbock, A.F.A.; Gao, M.; Qaseem, A.; Li, J.; Kirchner, P.A.; Ndingkokhar, B.; Bekono, B.D.; Simoben, C.V.; Babiaka, S.M.; Malange, Y.I.; et al. StreptomeDB 3.0: An updated compendium of streptomycetes natural products. Nucleic Acids Res. 2021, 49, D600–D604. [Google Scholar] [CrossRef]
  19. Jones, M.R.; Pinto, E.; Torres, M.A.; Dörr, F.; Mazur-Marzec, H.; Szubert, K.; Tartaglione, L.; Dell’Aversano, C.; Miles, C.O.; Beach, D.G.; et al. CyanoMetDB, a comprehensive public database of secondary metabolites from cyanobacteria. Water Res. 2021, 196, 117017. [Google Scholar] [CrossRef]
  20. Wang, D.G.; Wang, C.Y.; Hu, J.Q.; Wang, J.J.; Liu, W.C.; Zhang, W.J.; Du, X.R.; Wang, H.; Zhu, L.L.; Sui, H.Y.; et al. Constructing a Myxobacterial Natural Product Database to Facilitate NMR-Based Metabolomics Bioprospecting of Myxobacteria. Anal. Chem. 2023, 95, 5256–5266. [Google Scholar] [CrossRef]
  21. Ramirez-Gaona, M.; Marcu, A.; Pon, A.; Guo, A.C.; Sajed, T.; Wishart, N.A.; Karu, N.; Djoumbou Feunang, Y.; Arndt, D.; Wishart, D.S. YMDB 2.0: A significantly expanded version of the yeast metabolome database. Nucleic Acids Res. 2017, 45, D440–D445. [Google Scholar] [CrossRef] [PubMed]
  22. Foroutan, A.; Fitzsimmons, C.; Mandal, R.; Piri-Moghadam, H.; Zheng, J.; Guo, A.; Li, C.; Guan, L.L.; Wishart, D.S. The Bovine Metabolome. Metabolites 2020, 10, 233. [Google Scholar] [CrossRef] [PubMed]
  23. Takeshi Ara, T.; Sakurai, N.; Takahashi, S.; Waki, N.; Suganuma, H.; Aizawa, K.; Matsumura, Y.; Kawada, T.; Shibata, D. TOMATOMET: A metabolome database consists of 7118 accurate mass values detected in mature fruits of 25 tomato cultivars. Plant Direct. 2021, 5, e00318. [Google Scholar] [CrossRef]
  24. Link, H.; Kochanowski, K.; Sauer, U. Systematic identification of allosteric protein-metabolite interactions that control enzyme activity in vivo. Nat. Biotechnol. 2013, 31, 357–361. [Google Scholar] [CrossRef]
  25. Piazza, I.; Kochanowski, K.; Cappelletti, V.; Fuhrer, T.; Noor, E.; Sauer, U.; Picotti, P. A map of protein-metabolite interactions reveals principles of chemical communication. Cell 2018, 172, 358–372. [Google Scholar] [CrossRef]
  26. Rinschen, M.R.; Ivanisevic, J.; Giera, M.; Siuzdak, G. Identification of bioactive metabolites using activity metabolomics. Nat. Rev. Mol. Cell Biol. 2019, 20, 353–367. [Google Scholar] [CrossRef]
  27. Buescher, J.M.; Antoniewicz, M.R.; Boros, L.G.; Burgess, S.C.; Brunengraber, H.; Clish, C.B.; DeBerardinis, R.J.; Feron, O.; Frezza, C.; Ghesquiere, B.; et al. A roadmap for interpreting 13C metabolite labeling patterns from cells. Curr. Opin. Biotechnol. 2015, 34, 189–201. [Google Scholar] [CrossRef] [PubMed]
  28. Yu, D.; Zhou, L.; Liu, X.; Xu, G. Stable isotope-resolved metabolomics based on mass spectrometry: Methods and their applications. Trends Anal. Chem. 2023, 160, 116985. [Google Scholar] [CrossRef]
  29. Han, S.; Van Treuren, W.; Fischer, C.R.; Merrill, B.D.; DeFelice, B.C.; Sanchez, J.M.; Higginbottom, S.K.; Guthrie, L.; Fall, L.A.; Dodd, D.; et al. A metabolomics pipeline for the mechanistic interrogation of the gut microbiome. Nature 2021, 595, 415–420. [Google Scholar] [CrossRef]
  30. Schultheisz, H.L.; Szymczyna, B.R.; Scott, L.G.; Williamson, J.R. Pathway Engineered Enzymatic de novo Purine Nucleotide Synthesis. ACS Chem Biol. 2008, 3, 499–511. [Google Scholar] [CrossRef]
  31. Rowbotham, J.S.; Ramirez, M.A.; Lenz, O.; Reeve, H.A.; Vincent, K.A. Bringing biocatalytic deuteration into the toolbox of asymmetric isotopic labelling techniques. Nat. Commun. 2020, 11, 1454. [Google Scholar] [CrossRef] [PubMed]
  32. Nicolaou, K.C.; Snyder, S.A. Chasing molecules that were never there: Misassigned natural products and the role of chemical synthesis in modern structure elucidation. Angew. Chem. Int. Ed. 2005, 44, 1012–1044. [Google Scholar] [CrossRef] [PubMed]
  33. Sunazuka, T.; Hirose, T.; Omura, S. Efficient Total Synthesis of Novel Bioactive Microbial Metabolites. Acc. Chem. Res. 2008, 41, 302–314. [Google Scholar] [CrossRef] [PubMed]
  34. Nicolaou, K.C. Organic synthesis: The art and science of replicating the molecules of living nature and creating others like them in the laboratory. Proc. R. Soc. A 2014, 470, 20130690. [Google Scholar] [CrossRef] [PubMed]
  35. Wohlgemuth, R. Route Selection and Reaction Engineering for Sustainable Metabolite Synthesis. React. Chem. Eng. 2023, 8, 2109–2118. [Google Scholar] [CrossRef]
  36. Arnold, F.H. Directed evolution: Bringing new chemistry to life. Angew. Chem. Int. Ed. 2018, 57, 4143–4148. [Google Scholar] [CrossRef]
  37. List, B. Introduction: Organocatalysis. Chem. Rev. 2007, 107, 5413–5415. [Google Scholar] [CrossRef]
  38. MacMillan, D. The advent and development of organocatalysis. Nature 2008, 455, 304–308. [Google Scholar] [CrossRef]
  39. Murray, J.; Hodgson, D.R.W.; O’Donoghue, A.C. Going Full Circle with Organocatalysis and Biocatalysis: The Latent Potential of Cofactor Mimics in Asymmetric Synthesis. J. Org. Chem. 2023, 88, 7619–7629. [Google Scholar] [CrossRef]
  40. Demain, A.L. From natural products discovery to commercialization: A success story. J. Ind. Microbiol. Biotechnol. 2006, 33, 486–495. [Google Scholar] [CrossRef]
  41. Hoff, B.; Plassmeier, J.; Blankschien, M.; Letzel, A.C.; Kourtz, L.; Schröder, H.; Koch, W.; Zelder, O. Unlocking Nature’s Biosynthetic Power—Metabolic Engineering for the Fermentative Production of Chemicals. Angew. Chem. Int. Ed. 2021, 60, 2258–2278. [Google Scholar] [CrossRef] [PubMed]
  42. Yang, D.; Eun, H.; Prabowo, C.P.S.; Cho, S.; Lee, S.Y. Metabolic and cellular engineering for the production of natural products. Curr. Opin. Biotechnol. 2022, 77, 102760. [Google Scholar] [CrossRef] [PubMed]
  43. Wohlgemuth, R. Selective biocatalytic defunctionalization of raw materials. ChemSusChem 2022, 15, e202200402. [Google Scholar] [CrossRef] [PubMed]
  44. Kaiser, R. Scent of the Vanishing Flora; Wiley-VHCA AG: Zürich, Switzerland, 2010; ISBN 978-3-906390-64-2. [Google Scholar]
  45. Walsh, C.T.; Tang, Y. Natural Product Biosynthesis—Chemical Logic and Enzymatic Machinery; Royal Society of Chemistry: London, UK, 2017. [Google Scholar]
  46. Alcántara, A.R.; Dominguez de Maria, P.; Littlechild, J.A.; Schürmann, M.; Sheldon, R.A.; Wohlgemuth, R. Biocatalysis as key to sustainable industrial chemistry. ChemSusChem 2022, 15, e202102709. [Google Scholar] [CrossRef]
  47. Wohlgemuth, R. Tools and ingredients for the biocatalytic synthesis of metabolites. Biotechnol. J. 2009, 4, 1253–1265. [Google Scholar] [CrossRef]
  48. Wohlgemuth, R. Horizons of systems biocatalysis and renaissance of metabolite synthesis. Biotechnol. J. 2018, 13, 1700620. [Google Scholar] [CrossRef]
  49. Oberg, N.; Zallot, R.; Gerlt, J.A. EFI-EST, EFI-GNT, and EFI-CGFP: Enzyme Function Initiative (EFI) Web Resource for Genomic Enzymology Tools. J. Mol. Biol. 2023, 435, 168018. [Google Scholar] [CrossRef]
  50. Wittmann, C.; Liao, J.C. (Eds.) Industrial Biotechnology: Microorganism, 1st ed.; Wiley-VCH: Weinheim, Germany, 2017; ISBN 978-3-527-34179-5. [Google Scholar]
  51. Wittmann, C.; Liao, J.C. (Eds.) Industrial Biotechnology: Products and Processes, 1st ed.; Wiley-VCH: Weinheim, Germany, 2017; ISBN 978-3-527-34181-8. [Google Scholar]
  52. Lee, S.Y.; Nielsen, J.; Stephanopoulos, G. (Eds.) Metabolic Engineering—Concepts and Applications, 1st ed.; Wiley-VCH: Weinheim, Germany, 2021; ISBN 978-3-527-34662-2. [Google Scholar]
  53. Flickinger, M.C. (Ed.) Encyclopedia of Industrial Biotechnology: Bioprocess, Bioseparation, and Cell Technology; John Wliey & Sons: Hoboken, NJ, USA, 2010; Volumes 1–7, ISBN 978-0-471-79930-6. [Google Scholar]
  54. Wohlgemuth, R.; Littlechild, J. Complexity reduction and opportunities in the design, integration and intensification of biocata-lytic processes for metabolite synthesis. Front. Bioeng. Biotechnol. 2022, 10, 958606. [Google Scholar] [CrossRef]
  55. Wohlgemuth, R. Biocatalysis–Key enabling tools from biocatalytic one-step and multi-step reactions to biocatalytic total synthesis. New Biotechnol. 2021, 60, 113–123. [Google Scholar] [CrossRef]
  56. Wentrup, C. Origins of Organic Chemistry and Organic Synthesis. Eur. J. Org. Chem. 2022, 2022, e202101492. [Google Scholar] [CrossRef]
  57. Wöhler, F. Ueber künstliche Bildung des Harnstoffs. Ann. Phys. 1828, 88, 253–256. [Google Scholar] [CrossRef]
  58. Kolbe, H. Beiträge zur Kenntnis der gepaarten Verbindungen. Ann. Chem. Pharm. 1845, 54, 145–188. [Google Scholar] [CrossRef]
  59. Sheehan, J.C.; Henery-Logan, K.R. The total synthesis of penicillin V. J. Am. Chem. Soc. 1957, 79, 1262–1263. [Google Scholar] [CrossRef]
  60. Nicolaou, K.C.; Vourloumis, D.; Winssinger, N.; Baran, P.S. The art and science of total synthesis at the dawn of the twenty-first century. Angew. Chem. Int. Ed. 2000, 39, 44–122. [Google Scholar] [CrossRef]
  61. Veitch, G.E.; Boyer, A.; Ley, S.V. The Azadirachtin Story. Angew. Chem. Int. Ed. 2008, 47, 9402–9429. [Google Scholar] [CrossRef]
  62. Nicolaou, K.C.; Rigol, S. Perspectives from nearly five decades of total synthesis of natural products and their analogues for biology and medicine. Nat. Prod. Rep. 2020, 37, 1404–1435. [Google Scholar] [CrossRef] [PubMed]
  63. Min, L.; Han, J.C.; Zhang, W.; Gu, C.C.; Zou, Y.P.; Li, C.C. Strategies and Lessons Learned from Total Synthesis of Taxol. Chem. Rev. 2023, 123, 4934–4971. [Google Scholar] [CrossRef]
  64. Nicolaou, K.C.; Rigol, S.; Yu, R. Total synthesis endeavors and their contributions to science and society: A personal account. CCS Chemistry 2019, 1, 3–37. [Google Scholar] [CrossRef]
  65. Peters, D.S.; Pitts, C.R.; McClymont, K.S.; Stratton, T.P.; Bi, C.; Baran, P.S. Ideality in Context: Motivations for Total Synthesis. Acc. Chem. Res. 2021, 54, 605–617. [Google Scholar] [CrossRef]
  66. Cardwell, H.M.E.; Cornforth, J.W.; Duff, S.R.; Holtermann, H.; Robinson, R. Total synthesis of androgenic hormones. Chem. Ind. 1951, 20, 389–390. [Google Scholar]
  67. Woodward, R.B.; Sondheimer, F.; Taub, D. The total synthesis of cholesterol. J. Am. Chem. Soc. 1951, 73, 3548. [Google Scholar] [CrossRef]
  68. Woodward, R.B.; Sondheimer, F.; Taub, D.; Heusler, K.; McLamore, W.M. The Total Synthesis of Steroids. J. Am. Chem. Soc. 1952, 74, 4223–4251. [Google Scholar] [CrossRef]
  69. Munt, M.; Spieß, O.; Indolese, A.; Roux, L.; Giraud, M.; Schinzer, D. Short and Scalable Synthesis of Plant-Based Cholesterol in GMP Grade. Adv. Synth. Catal. 2023, 365, 2406–2409. [Google Scholar] [CrossRef]
  70. Eschenmoser, A.; Wintner, C.E. Natural Product Synthesis and Vitamin B12: Total synthesis of vitamin B12 provided a framework for exploration in several areas of organic chemistry. Science 1977, 196, 1410–1420. [Google Scholar] [CrossRef]
  71. Woodward, R.B. The total synthesis of vitamin B12. Pure Appl. Chem. 1973, 33, 145–178. [Google Scholar] [CrossRef] [PubMed]
  72. Scott, A.I. Discovering Nature’s Diverse Pathways to Vitamin B12: A 35-Year Odyssey. J. Org. Chem. 2003, 68, 2529–2539. [Google Scholar] [CrossRef]
  73. Eschenmoser, A. Vitamin B12: Experiments concerning the origin of its molecular structure. Angew. Chem. Int. Ed. 1988, 27, 5–39. [Google Scholar] [CrossRef]
  74. Kishi, Y. Palytoxin: An inexhaustible source of inspiration—Personal perspective. Tetrahedron 2002, 58, 6239–6258. [Google Scholar] [CrossRef]
  75. Sears, J.E.; Boger, D.L. Total Synthesis of Vinblastine, Related Natural Products, and Key Analogues and Development of Inspired Methodology Suitable for the Systematic Study of Their Structure−Function Properties. ACC Chem. Res. 2015, 48, 653–662. [Google Scholar] [CrossRef]
  76. Baran, P.S.; Maimone, T.J.; Richter, J.M. Total synthesis of marine natural products without using protecting groups. Nature 2007, 446, 404–408. [Google Scholar] [CrossRef]
  77. Pasteur, L. Mémoire sur la fermentation appelée lactique. Comptes Rendus Chim. 1857, 45, 913–916. [Google Scholar]
  78. Buchner, E. Alkoholische Gährung ohne Hefezellen. Berichte Der Dtsch. Chem. Ges. 1897, 30, 117–124. [Google Scholar] [CrossRef]
  79. Krebs, H.A.; Henseleit, K. Untersuchungen über die Harnstoffbildung im Tierkörper. Klin. Wochenschr. 1932, 11, 757–759. [Google Scholar] [CrossRef]
  80. Krebs, H.A.; Henseleit, K. Untersuchungen über die Harnstoffbildung im Tierkörper. II. Klin. Wochenschr. 1932, 11, 1137–1139. [Google Scholar] [CrossRef]
  81. Amato, A.; Becci, A.; Beolchini, F. Citric acid bioproduction: The technological innovation change. Crit. Rev. Biotechnol. 2020, 40, 199–212. [Google Scholar] [CrossRef]
  82. Feng, J.; Wu, Q.; Zhu, D.; Ma, Y. Biotransformation Enables Innovations Toward Green Synthesis of Steroidal Pharmaceuticals. ChemSusChem 2022, 15, e2021023. [Google Scholar] [CrossRef]
  83. Elander, R.P. Industrial production of β-lactam antibiotics. Appl. Microbiol. Biotechnol. 2003, 61, 385–392. [Google Scholar] [CrossRef]
  84. Meyer, H.P.; Robins, K.T. Large scale bioprocess for the production of optically pure L-carnitine. Monatsh. Chem. 2005, 136, 1269–1277. [Google Scholar] [CrossRef]
  85. Sanchez, S.; Rodríguez-Sanoja, R.; Ramos, A.; Demain, A.L. Our microbes not only produce antibiotics, they also overproduce amino acids. J. Antibiot. 2018, 71, 26–36. [Google Scholar] [CrossRef]
  86. Vandamme, E.J.; Revuelta, J.L. Industrial Biotechnology of Vitamins, Biopigments, and Antioxidants; Wiley-VCH: Weinheim, Germany, 2016. [Google Scholar] [CrossRef]
  87. Ramírez-Rendon, D.; Passari, A.K.; Ruiz-Villafán, B.; Rodríguez-Sanoja, R.; Sánchez, S.; Demain, A.L. Impact of novel microbial secondary metabolites on the pharma industry. Appl. Microbiol. Biotechnol. 2022, 106, 1855–1878. [Google Scholar] [CrossRef] [PubMed]
  88. Fernández-Cabezón, L.; Galán, B.; García, J.L. New Insights on Steroid Biotechnology. Front. Microbiol. 2018, 9, 958. [Google Scholar] [CrossRef]
  89. Cheng, C.; Tsai, H.R. Yeast-mediated stereo-selective reduction of estrone by continuous cell culture with dual stirred tanks for product yield improvement. J. Chem. Technol. Biotechnol. 2011, 86, 601–607. [Google Scholar] [CrossRef]
  90. Chang, H.; Zhang, H.; Zhu, L.; Zhang, W.; You, S.; Qi, W.; Qian, J.; Su, R.; He, Z. A combined strategy of metabolic pathway regulation and two-step bioprocess for improved 4-androstene-3, 17-dione production with an engineered Mycobacterium neoaurum. Biochem. Eng. J. 2020, 164, 107789. [Google Scholar] [CrossRef]
  91. Su, B.M.; Zhao, H.R.; Xu, L.; Xu, X.Q.; Wang, L.C.; Lin, J.; Lin, W. Construction of an Efficient Non-natural Enzyme System for Preparation of Testosterone in High Space-Time Yield. ACS Sustain. Chem. Eng. 2022, 10, 3373–3382. [Google Scholar] [CrossRef]
  92. Kille, S.; Zilly, F.E.; Acevedo, J.P.; Reetz, M.T. Regio- and stereoselectivity of P450-catalysed hydroxylation of steroids controlled by laboratory evolution. Nat. Chem. 2011, 3, 738–743. [Google Scholar] [CrossRef]
  93. Pan, H.; Chang, S.; Qu, Y.; Liu, M.; Tian, W.; Chang, Z. Hydrocortisone production using whole-cell biocatalysts in recombinant Escherichia coli. Biochem. Eng. J. 2023, 198, 109023. [Google Scholar] [CrossRef]
  94. Gu, Y.; Jiao, X.; Ye, L.; Yu, H. Metabolic engineering strategies for de novo biosynthesis of sterols and steroids in yeast. Bioresour. Bioprocess. 2021, 8, 110. [Google Scholar] [CrossRef]
  95. Thoma, R.; Schulz-Gasch, T.; D’Arcy, B.; Benz, J.; Aebi, J.; Dehmlow, H.; Hennig, M.; Stihle, M.; Ruf, A. Insight into steroid scaffold formation from the structure of human oxidosqualene cyclase. Nature 2004, 432, 118–122. [Google Scholar] [CrossRef] [PubMed]
  96. Muthulakshmi, M.V.; Srinivasan, A.; Srivastava, S. Antioxidant Green Factories: Toward Sustainable Production of Vitamin E in Plant In Vitro Cultures. ACS Omega 2023, 8, 3586–3605. [Google Scholar] [CrossRef]
  97. Ye, Z.; Shi, B.; Huang, Y.; Ma, T.; Xiang, Z.; Hu, B.; Kuang, Z.; Huang, M.; Lin, X.; Tian, Z.; et al. Revolution of vitamin E production by starting from microbial fermented farnesene to isophytol. Innovation 2022, 3, 100228. [Google Scholar] [CrossRef]
  98. Zhu, K.; Jiang, M.; Ye, B.; Zhang, G.T.; Li, W.; Tang, P.; Huang, Z.; Chen, F. A unified strategy to prostaglandins: Chemoenzymatic total synthesis of cloprostenol, bimatoprost, PGF2a, fluprostenol, and travoprost guided by biocatalytic retrosynthesis. Chem. Sci. 2021, 12, 10362–10370. [Google Scholar] [CrossRef]
  99. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs over the Nearly Four Decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef]
  100. Bringi, V.; Kadkade, P.G.; Prince, C.L.; Roach, B.L. Enhanced Production of Taxol and Taxanes by Cell Cultures of Taxus Species. U.S. Patent 2013/0017582 A1, 17 January 2013. [Google Scholar]
  101. Sharma, A.; Bhatia, S.K.; Banyal, A.; Chanana, I.; Kumar, A.; Chand, D.; Kulshrestha, S.; Kumar, P. An Overview on Taxol Production Technology and Its Applications as Anticancer Agent. Biotechnol. Bioprocess Engin. 2022, 27, 706–728. [Google Scholar] [CrossRef]
  102. Omura, S. A Splendid Gift from the Earth: The Origins and Impact of the Avermectins (Nobel Lecture). Angew. Chem. Int. Ed. 2016, 55, 10190–10209. [Google Scholar] [CrossRef] [PubMed]
  103. Tu, Y. Artemisinin—A Gift from Traditional Chinese Medicine to the World (Nobel Lecture). Angew. Chem. Int. Ed. 2016, 55, 10210–10226. [Google Scholar] [CrossRef] [PubMed]
  104. Fischbach, M.A.; Walsh, C.T. Antibiotics for Emerging Pathogens. Science 2009, 325, 1089–1093. [Google Scholar] [CrossRef]
  105. Lewis, K. The Science of Antibiotic Discovery. Cell 2020, 181, 29–45. [Google Scholar] [CrossRef] [PubMed]
  106. Miethke, M.; Pieroni, M.; Weber, T.; Brönstrup, M.; Hammann, P.; Halby, L.; Arimondo, P.B.; Glaser, P.; Aigle, B.; Bode, H.B.; et al. Towards the sustainable discovery and development of new antibiotics. Nat. Rev. Chem. 2021, 5, 726–749. [Google Scholar] [CrossRef]
  107. Calvillo, A.; Pellicer, T.; Carnicer, M.; Planas, A. Bioprocess Strategies for Vitamin B12 Production by Microbial Fermentation and Market Applications. Bioengineering 2022, 9, 365. [Google Scholar] [CrossRef]
  108. Eichhorn, E.; Locher, E.; Guillemer, S.; Wahler, D.; Fourage, L.; Schilling, B. Biocatalytic process for (−)-Ambrox production using squalene hopene cyclase. Adv. Synth. Catal. 2018, 360, 2339–2351. [Google Scholar] [CrossRef]
  109. Meyer, H.P.; Eichhorn, E.; Hanlon, S.; Lütz, S.; Schürmann, M.; Wohlgemuth, R.; Coppolecchia, R. The use of enzymes in organic synthesis and the life sciences: Perspectives from the Swiss Industrial Biocatalysis Consortium (SIBC). Catal. Sci. Technol. 2013, 3, 29–40. [Google Scholar] [CrossRef]
  110. Bühlmann, P.; Pretsch, E.; Bakker, E. Carrier-Based Ion-Selective Electrodes and Bulk Optodes. 2. Ionophores for Potentiometric and Optical Sensors. Chem. Rev. 1998, 98, 1593–1687. [Google Scholar] [CrossRef] [PubMed]
  111. Jani, P.; Emmert, J.; Wohlgemuth, R. Process analysis of macrotetrolide biosynthesis during fermentation by means of direct infusion LC-MS. Biotechnol. J. 2008, 3, 202–208. [Google Scholar] [CrossRef] [PubMed]
  112. Cuartero, M.; Colozza, N.; Fernández-Pérez, B.M.; Crespo, G.A. Why ammonium detection is particularly challenging but insightful with ionophore-based potentiometric sensors—An overview of the progress in the last 20 years. Analyst 2020, 145, 3188–3210. [Google Scholar] [CrossRef] [PubMed]
  113. Gauss, D.; Schoenenberger, B.; Wohlgemuth, R. Chemical and enzymatic methodologies for the synthesis of enantiomerically pure glyceraldehyde 3-phosphates. Carbohydr. Res. 2014, 389, 18–24. [Google Scholar] [CrossRef] [PubMed]
  114. Molla, G.S.; Kinfu, B.M.; Chow, J.; Streit, W.; Wohlgemuth, R.; Liese, A. Bioreaction engineering leading to efficient synthesis of L-glyceraldehyd-3-phosphate. Biotechnol. J. 2017, 12, 1600625. [Google Scholar] [CrossRef] [PubMed]
  115. Richter, N.; Neumann, M.; Liese, A.; Wohlgemuth, R.; Eggert THummel, W. Characterisation of a Recombinant NADP-Dependent Glycerol Dehydrogenase from Gluconobacter oxydans and its Application in the Production of L-Glyceraldehyde. ChemBioChem 2009, 10, 1888–1896. [Google Scholar] [CrossRef]
  116. Richter, N.; Neumann, M.; Liese, A.; Wohlgemuth, R.; Weckbecker, A.; Eggert, T.; Hummel, W. Characterization of a whole-cell catalyst co-expressing glycerol dehydrogenase and glucose dehydrogenase and its application in the synthesis of L-glyceraldehyde. Biotechnol. Bioeng. 2010, 106, 541–552. [Google Scholar] [CrossRef]
  117. Gauss, D.; Sánchez-Moreno, I.; Oroz-Guinea, I.; García-Junceda, E.; Wohlgemuth, R. Phosphorylation catalyzed by dihydroxyacetone kinase. Eur. J. Org. Chem. 2018, 23, 2892–2895. [Google Scholar] [CrossRef]
  118. Hardt, N.; Kinfu, B.M.; Chow, J.; Schoenenberger, B.; Streit, W.R.; Obkircher, M.; Wohlgemuth, R. Biocatalytic Asymmetric Phosphorylation Catalyzed by Recombinant Glycerate-2-Kinase. ChemBioChem 2017, 18, 1518–1522. [Google Scholar] [CrossRef]
  119. Shaeri, J.; Wohlgemuth, R.; Woodley, J.M. Semiquantitative process screening for the biocatalytic synthesis of D-xylulose-5-phosphate. Org. Proc. Res. Dev. 2006, 10, 605–610. [Google Scholar] [CrossRef]
  120. Shaeri, J.; Wright, I.; Rathbone, E.B.; Wohlgemuth, R.; Woodley, J.M. Characterization of enzymatic D-xylulose 5-phosphate synthesis. Biotechnol. Bioeng. 2008, 101, 761–767. [Google Scholar] [CrossRef] [PubMed]
  121. Hardt, N.; Kind, S.; Schoenenberger, B.; Eggert, T.; Obkircher, M.; Wohlgemuth, R. Facile synthesis of D-xylulose-5-phosphate and L-xylulose-5-phosphate by xylulokinase-catalyzed phosphorylation. Biocatal. Biotransform. 2020, 38, 35–45. [Google Scholar] [CrossRef]
  122. Schoenenberger, B.; Kind, S.; Meier, R.; Eggert, T.; Obkircher, M.; Wohlgemuth, R. Efficient biocatalytic synthesis of D-tagatose 1, 6-diphosphate by LacC-catalysed phosphorylation of D-tagatose 6-phosphate. Biocatal. Biotransform. 2020, 38, 53–63. [Google Scholar] [CrossRef]
  123. Krevet, S.; Shen, L.; Bohnen, T.; Schoenenberger, B.; Meier, R.; Obkircher, M.; Bangert, K.; Koehling, R.; Allenspach, E.; Wohlgemuth, R.; et al. Enzymatic synthesis of 2-keto-3-deoxy-6-phosphogluconate by the 6-phosphogluconate-dehydratase from Caulobacter crescentus. Front. Bioeng. Biotechnol. 2020, 8, 185. [Google Scholar] [CrossRef] [PubMed]
  124. Shen, L.; Kohlhaas, M.; Enoki, J.; Meier, R.; Schönenberger, B.; Wohlgemuth, R.; Kourist, R.; Niemeyer, F.; van Niekerk, D.; Bräsen, C.; et al. A combined experimental and modelling approach for the Weimberg pathway optimisation. Nat. Commun. 2020, 11, 1098. [Google Scholar] [CrossRef]
  125. Matsubara, K.; Köhling, R.; Schönenberger, B.; Kouril, T.; Esser, D.; Bräsen, C.; Siebers, B.; Wohlgemuth, R. One-step synthesis of 2-keto-3-deoxy-D-gluconate by biocatalytic dehydration of D-gluconate. J. Biotechnol. 2014, 191, 69–77. [Google Scholar] [CrossRef]
  126. Schoenenberger, B.; Wszolek, A.; Milesi, T.; Brundiek, H.; Obkircher, M.; Wohlgemuth, R. Synthesis of Nω-Phospho-L-arginine by Biocatalytic Phosphorylation of L-Arginine. ChemCatChem 2017, 9, 121–126. [Google Scholar] [CrossRef]
  127. Schoenenberger, B.; Wszolek, A.; Meier, R.; Brundiek, H.; Obkircher, M.; Wohlgemuth, R. Recombinant AroL-Catalyzed Phosphorylation for the Efficient Synthesis of Shikimic Acid 3-Phosphate. Biotechnol. J. 2018, 13, 1700529. [Google Scholar] [CrossRef]
  128. Schell, U.; Wohlgemuth, R.; Ward, J.M. Synthesis of pyridoxamine 5′-phosphate using an MBA: Pyruvate transaminase as biocatalyst. J. Mol. Catal. B Enzym. 2009, 59, 279–285. [Google Scholar] [CrossRef]
  129. Schoenenberger, B.; Wszolek, A.; Meier, R.; Brundiek, H.; Obkircher, M.; Wohlgemuth, R. Biocatalytic asymmetric Michael addition reaction of l-arginine to fumarate for the green synthesis of N-(([(4S)-4-amino-4-carboxy-butyl] amino) iminomethyl)-L-aspartic acid lithium salt (L-argininosuccinic acid lithium salt). RSC Adv. 2017, 7, 48952–48957. [Google Scholar] [CrossRef]
  130. Schoenenberger, B.; Wszolek, A.; Meier, R.; Brundiek, H.; Obkircher, M.; Wohlgemuth, R. Biocatalytic Asymmetric Aza-Michael Addition Reactions and Synthesis of L-Argininosuccinate by Argininosuccinate Lyase ARG4-Catalysed Aza-Michael Addition of L-Arginine to Fumarate. In Applied Biocatalysis; Whittall, J., Sutton, P.W., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2021; pp. 204–210. [Google Scholar]
  131. Yasutake, Y.; Nishioka, T.; Imoto, N.; Tamura, T. A Single Mutation at the Ferredoxin Binding Site of P450 Vdh Enables Efficient Biocatalytic Production of 25-Hydroxyvitamin D3. ChemBioChem 2013, 14, 2284–2291. [Google Scholar] [CrossRef] [PubMed]
  132. Tang, D.; Liu, W.; Huang, L.; Cheng, L.; Xu, Z. Efficient biotransformation of vitamin D3 to 25-hydroxyvitamin D3 by a newly isolated Bacillus cereus strain. Appl. Microbiol. Biotechnol. 2020, 104, 765–774. [Google Scholar] [CrossRef] [PubMed]
  133. Babot, E.D.; del Río, J.C.; Kalum, L.; Martínez, A.T.; Gutiérrez, A. Regioselective Hydroxylation in the Production of 25-Hydroxyvitamin D by Coprinopsis cinerea Peroxygenase. ChemCatChem 2015, 7, 283–290. [Google Scholar] [CrossRef]
  134. Warnke, M.; Jung, T.; Dermer, J.; Hipp, K.; Jehmlich, N.; von Bergen, M.; Ferlaino, S.; Fries, A.; Müller, M.; Boll, M. 25-Hydroxyvitamin D3 Synthesis by Enzymatic Steroid Side-Chain Hydroxylation with Water. Angew. Chem. Int. Ed. 2016, 55, 1881–1884. [Google Scholar] [CrossRef]
  135. Kang, D.J.; Im, J.H.; Kang, J.H.; Kim, K.H. Whole cell bioconversion of vitamin D3 to calcitriol using Pseudonocardia sp. KCTC 1029BP. Bioprocess Biosyst. Eng. 2015, 38, 1281–1290. [Google Scholar] [CrossRef]
  136. Taniguchi, T.; Eto, T.A.; Shiotsuki, H.; Sueta, H.; Higashi, S.; Iwamura, T.; Okuda, K.I.; Setoguchi, T. Newly established assay method for 25-hydroxyvitamin D3 24-hydroxylase revealed much lower Km for 25-hydroxyvitamin D3 than for 1alpha,25-dihydroxyvitamin D3. J. Bone Miner. Res. 2001, 16, 57–62. [Google Scholar] [CrossRef]
  137. Yasuda, K.; Yogo, Y.; Sugimoto, H.; Mano, H.; Takita, T.; Ohta, M.; Kamakura, M.; Ikushiro, S.; Yasukawa, K.; Shiro, Y.; et al. Production of an active form of vitamin D2 by genetically engineered CYP105A1. Biochem. Biophys. Res. Commun. 2017, 486, 336–341. [Google Scholar] [CrossRef]
  138. Putkaradze, N.; König, L.; Kattner, L.; Hutter, M.C.; Bernhardt, R. Highly regio- and stereoselective hydroxylation of vitamin D2 by CYP109E1. Biochem. Biophys. Res. Commun. 2020, 524, 295–300. [Google Scholar] [CrossRef]
  139. Coene, K.L.; Kluijtmans, L.A.; van der Heeft, E.; Engelke, U.F.; de Boer, S.; Hoegen, B.; Kwast, H.J.; van de Vorst, M.; Huigen, M.C.; Keularts, I.M.; et al. Next-generation metabolic screening: Targeted and untargeted metabolomics for the diagnosis of inborn errors of metabolism in individual patients. J. Inherit. Metab. Dis. 2018, 41, 337–353. [Google Scholar] [CrossRef]
  140. Pavlova, N.N.; Thompson, C.B. The emerging hallmarks of cancer metabolism. Cell Metab. 2016, 23, 27–47. [Google Scholar] [CrossRef] [PubMed]
  141. Nemet, I.; Saha, P.P.; Gupta, N.; Zhu, W.; Romano, K.A.; Skye, S.M.; Cajka, T.; Mohan, M.L.; Li, L.; Wu, Y.; et al. A cardiovascular disease-linked gut microbial metabolite acts via adrenergic receptors. Cell 2020, 180, 862–877. [Google Scholar] [CrossRef] [PubMed]
  142. Calvin, M. The Path of Carbon in Photosynthesis: The carbon cycle is a tool for exploring chemical biodynamics and the mechanism of quantum conversion. Science 1962, 135, 879–889. [Google Scholar] [CrossRef] [PubMed]
  143. Rising, K.A.; Schramm, V.L. Enzymatic Synthesis of NAD+ with the Specific Incorporation of Atomic Labels. J. Am. Chem. Soc. 1994, 116, 6531–6536. [Google Scholar] [CrossRef]
  144. Tran, A.; Yokose, R.; Cen, Y. Chemo-enzymatic synthesis of isotopically labelled nicotinamide riboside. Org. Biomol. Chem. 2018, 16, 3662–3671. [Google Scholar] [CrossRef]
  145. Khoroshilov, A.V. Production of stable isotopes of light elements: Past, present and future. J. Phys. Conf. Ser. 2018, 1099, 012002. [Google Scholar] [CrossRef]
  146. Letertre, M.P.M.; Dervilly, G.; Giraudeau, P. Combined Nuclear Magnetic Resonance Spectroscopy and Mass Spectrometry Approaches for Metabolomics. Anal. Chem. 2021, 93, 500–518. [Google Scholar] [CrossRef]
  147. Sauer, U.; Lasko, D.R.; Fiaux, J.; Hochuli, M.; Glaser, R.; Szyperski, T.; Wüthrich, K.; Bailey, J.E. Metabolic flux ratio analysis of genetic and environmental modulations of Escherichia coli central carbon metabolism. J. Bacteriol. 1999, 181, 6679–6688. [Google Scholar] [CrossRef]
  148. Faubert, B.; Tasdogan, A.; Morrison, S.J.; Mathews, T.P.; DeBerardinis, R.J. Stable isotope tracing to assess tumor metabolism in vivo. Nat. Protoc. 2021, 16, 5123–5145. [Google Scholar] [CrossRef]
  149. Steinhauser, M.L.; Bailey, A.P.; Senyo, S.E.; Guillermier, C.; Perlstein, T.S.; Gould, A.P.; Lee, R.T.; Lechene, C.P. Multi-isotope imaging mass spectrometry quantifies stem cell division and metabolism. Nature 2012, 481, 516–519. [Google Scholar] [CrossRef]
  150. Grey, A.C.; Tang, M.; Zahraei, A.; Guo, G.; Demarais, N.J. Applications of stable isotopes in MALDI imaging: Current approaches and an eye on the future. Anal. Bioanal. Chem. 2021, 413, 2637–2653. [Google Scholar] [CrossRef]
  151. Fan, T.W.M.; Lane, A.N. Applications of NMR spectroscopy to systems biochemistry. Prog. Nucl. Magn. Reson. Spectrosc. 2016, 92–93, 18–53. [Google Scholar] [CrossRef]
  152. Giraudeau, P. Quantitative NMR spectroscopy of complex mixtures. Chem. Commun. 2023, 59, 6627–6642. [Google Scholar] [CrossRef]
  153. Wohlgemuth, R.; Waespe-Sarcevic, N.; Seelig, J. Bilayers of Phosphatidylglycerol. A Deuterium and Phosphorus NuclearMagnetic Resonance Study of the Head-Group Region. Biochemistry 1980, 19, 3315–3321. [Google Scholar] [CrossRef] [PubMed]
  154. Chun, S.W.; Narayan, A.R.H. Biocatalytic, Stereoselective Deuteration of α-Amino Acids and Methyl Esters. ACS Catal. 2020, 10, 7413–7418. [Google Scholar] [CrossRef] [PubMed]
  155. Doyon, T.J.; Buller, A.R. Site-Selective Deuteration of Amino Acids through Dual-Protein Catalysis. J. Am. Chem. Soc. 2022, 144, 7327–7336. [Google Scholar] [CrossRef] [PubMed]
  156. Rowbotham, J.S.; Hardy, A.P.; Reeve, H.A.; Vincent, K.A. Synthesis of [4S-2H] NADH, [4R-2H] NADH, [4-2H2] NADH and [4-2H] NAD+ cofactors through heterogeneous biocatalysis in heavy water. J. Label. Compd. Radiopharm. 2021, 64, 181–186. [Google Scholar] [CrossRef] [PubMed]
  157. Xu, J.; Lou, Y.; Wang, L.; Wang, Z.; Xu, W.; Ma, W.; Chen, Z.; Chen, X.; Wu, Q. Rational Design of Biocatalytic Deuteration Platform of Aldehydes. ACS Catal. 2021, 11, 13348–13354. [Google Scholar] [CrossRef]
  158. Tolbert, T.J.; Williamson, J.R. Preparation of Specifically Deuterated RNA for NMR Studies Using a Combination of Chemical and Enzymatic Synthesis. J. Am. Chem. Soc. 1996, 118, 7929–7940. [Google Scholar] [CrossRef]
  159. Bennett, B.D.; Yuan, J.; Kimball, E.H.; Rabinowitz, J.D. Absolute quantitation of intracellular metabolite concentrations by an isotope ratio-based approach. Nat. Protoc. 2008, 3, 1299–1311. [Google Scholar] [CrossRef]
  160. Sauer, U. Metabolic networks in motion: 13C-based flux analysis. Mol. Syst. Biol. 2006, 2, 62. [Google Scholar] [CrossRef] [PubMed]
  161. Arrivault, S.; Guenther, M.; Fry, S.C.; Fuenfgeld, M.F.F.F.; Veyel, D.; Mettler-Altmann, T.; Stitt, M.; Lunn, J.E. Synthesis and Use of Stable-Isotope-Labelled Internal Standards for Quantification of Phosphorylated Metabolites by LC–MS/MS. Anal. Chem. 2015, 87, 6896–6904. [Google Scholar] [CrossRef]
  162. Eisenreich, W.; Schwarz, M.; Cartayrade, A.; Arigoni, D.; Zenk, M.H.; Bacher, A. The deoxyxylulose phosphate pathway of terpenoid biosynthesis in plants and microorganisms. Chem. Biol. 1998, 5, R221–R233. Available online: http://biomednet.com/elecref/10745521005R0221 (accessed on 8 August 2023). [CrossRef] [PubMed]
  163. Zhang, W.; Zhao, S.; Serianni, A.S. Labeling monosaccharides with stable isotopes. Methods Enzymol. 2015, 565, 423–458. [Google Scholar] [CrossRef] [PubMed]
  164. Goux, W.J.; Rench, L.; Weber, D.S. Stereoselective synthesis of stable isotope labeled L-α-amino acids: The enzymatic preparation of 13C-labeled L-glutamic acids. J. Label. Compd. Radiopharm. 1993, 33, 181–193. [Google Scholar] [CrossRef]
  165. Maeda, H.; Takata, K.; Toyoda, A.; Niitsu, T.; Iwakura, M.; Shibata, K. Production of L-[3-13C] serine from [13C] formaldehyde and glycine using an enzyme system combined with tetrahydrofolate regeneration. J. Ferment. Bioeng. 1997, 83, 113–115. [Google Scholar] [CrossRef]
  166. Jemielity, J.; Kańska, M.; Kański, R. Enzymatic Synthesis of [1-13C]-and [1-14C]-L-Phenyl-Alanine. Isot. Environ. Health Stud. 1998, 34, 335–339. [Google Scholar] [CrossRef]
  167. Akita, H.; Suzuki, H.; Doi, K.; Ohshima, T. Efficient synthesis of D-branched-chain amino acids and their labeled compounds with stable isotopes using D-amino acid dehydrogenase. Appl. Microbiol. Biotechnol. 2014, 98, 1135–1143. [Google Scholar] [CrossRef]
  168. Van Raad, D.; Huber, T.; Otting, G. Improved spectral resolution of [13C,1H]-HSQC spectra of aromatic amino acid residues in proteins produced by cell-free synthesis from inexpensive 13C-labelled precursors. J. Biomol. NMR 2023, 77, 183–190. [Google Scholar] [CrossRef]
  169. Guo, X.; Liu, Y.; Wang, Q.; Wang, X.; Li, Q.; Liu, W.; Zhao, Z.K. Non-natural Cofactor and Formate-Driven Reductive Carboxylation of Pyruvate. Angew. Chem. Int. Ed. 2020, 59, 3143–3146. [Google Scholar] [CrossRef]
  170. Morgan, K.D. The use of nitrogen-15 in microbial natural product discovery and biosynthetic characterization. Front. Microbiol. 2023, 14, 1174591. [Google Scholar] [CrossRef] [PubMed]
  171. Chiriaca, M.; Lupan, I.; Popa, F.; Palibroda, N.; Popescu, O. Enzymatic synthesis of some 15N-labelled L-amino acids. Isot. Environ. Health Stud. 2010, 46, 249–254. [Google Scholar] [CrossRef] [PubMed]
  172. Wang, M.; Asam, S.; Chen, J.; Ehrmann, M.; Rychlik, M. Production of Four 15N-Labelled Cobalamins via Biosynthesis Using Propionibacterium freudenreichii. Front. Microbiol. 2021, 12, 713321. [Google Scholar] [CrossRef] [PubMed]
  173. Wishart, D.S.; Feunang, Y.D.; Guo, A.C.; Lo, E.J.; Marcu, A.; Grant, J.R.; Sajed, T.; Johnson, D.; Li, C.; Sayeeda, Z.; et al. DrugBank 5.0: A major update to the DrugBank database for 2018. Nucleic Acids Res. 2018, 46, D1074–D1082. [Google Scholar] [CrossRef]
  174. Clayton, T.A.; Baker, D.; Lindon, J.C.; Everett, J.R.; Nicholson, J.K. Pharmacometabonomic identification of a significant host-microbiome metabolic interaction affecting human drug metabolism. Proc. Natl. Acad. Sci. USA 2009, 106, 14728–14733. [Google Scholar] [CrossRef]
  175. Zimmermann, M.; Zimmermann-Kogadeeva, M.; Wegmann, R.; Goodman, A.L. Mapping human microbiome drug metabolism by gut bacteria and their genes. Nature 2019, 570, 462–467. [Google Scholar] [CrossRef]
  176. Javdan, B.; Lopez, J.G.; Chankhamjon, P.; Lee, Y.C.J.; Hull, R.; Wu, Q.; Wang, X.; Chatterjee, S.; Donia, M.S. Personalized mapping of drug metabolism by the human gut microbiome. Cell 2020, 181, 1661–1679.e22. [Google Scholar] [CrossRef]
  177. Heinken, A.; Hertel, J.; Acharya, G.; Ravcheev, D.A.; Nyga, M.; Okpala, O.E.; Hogan, M.; Magnúsdóttir, S.; Martinelli, F.; Nap, B.; et al. Genome-scale metabolic reconstruction of 7,302 human microorganisms for personalized medicine. Nat. Biotechnol. 2023, 41, 1320–1331. [Google Scholar] [CrossRef]
  178. Fura, A.; Shu, Y.Z.; Zhu, M.; Hanson, R.L.; Roongta, V.; Humphreys, W.G. Discovering Drugs through Biological Transformation: Role of Pharmacologically Active Metabolites in Drug Discovery. J. Med. Chem. 2004, 47, 4339–4351. [Google Scholar] [CrossRef]
  179. Rautio, J.; Meanwell, N.A.; Di, L.; Hageman, M.J. The expanding role of prodrugs in contemporary drug design and development. Nat. Rev. Drug Discov. 2018, 17, 559–587. [Google Scholar] [CrossRef]
  180. Schadt, S.; Bister, B.; Chowdhury, S.K.; Funk, C.; Hop, C.E.C.A.; Humphreys, W.G.; Igarashi, F.; James, A.D.; Kagan, M.; Khojasteh, S.C.; et al. A Decade in the MIST: Learnings from Investigations of Drug Metabolites in Drug Development under the “Metabolites in Safety Testing” Regulatory Guidance. Drug Metab. Dispos. 2018, 46, 865–878. [Google Scholar] [CrossRef] [PubMed]
  181. US Food and Drug Administration (FDA). Safety Testing of Drug Metabolites. 2020. Available online: https://www.fda.gov/media/72279/download (accessed on 8 August 2023).
  182. Luffer-Atlas, D.; Obach, R.S.; Smith, D.A. A MIST conception: What has been learned from twenty years of human metabolite safety assessment? Med. Chem. Res. 2023, 32, 1933–1949. [Google Scholar] [CrossRef]
  183. Chhatrapati Bisen, A.; Nashik Sanap, S.; Agrawal, S.; Biswas, A.; Sankar Bhatta, R. Chemical metabolite synthesis and profiling: Mimicking in vivo biotransformation reactions. Bioorg. Chem. 2023, 139, 106722. [Google Scholar] [CrossRef]
  184. Winkler, M.; Geier, M.; Hanlon, S.P.; Nidetzky, B.; Glieder, A. Human enzymes for organic synthesis. Angew. Chem. Int. Ed. 2018, 57, 13406–13423. [Google Scholar] [CrossRef] [PubMed]
  185. Naumann, J.M.; Zöllner, A.; Drăgan, C.A.; Messinger, J.; Adam, J.; Bureik, M. Biotechnological Production of 20-alpha-Dihydrodydrogesterone at Pilot Scale. Appl. Biochem. Biotechnol. 2011, 165, 190–203. [Google Scholar] [CrossRef] [PubMed]
  186. Brinkmann, V.; Billich, A.; Baumruker, T.; Heining, P.; Schmouder, R.; Francis, G.; Aradhye, S.; Burtin, P. Fingolimod (FTY720): Discovery and development of an oral drug to treat multiple sclerosis. Nat. Rev. Drug Discov. 2010, 9, 883–897. [Google Scholar] [CrossRef]
  187. Kittelmann, M.; Rheinegger, U.; Espigat, A.; Oberer, L.; Aichholz, R.; Francotte, E.; Ghisalba, O. Preparative Enzymatic Synthesis of the Acylglucuronide of Mycophenolic Acid. Adv. Synth. Catal. 2003, 345, 825–829. [Google Scholar] [CrossRef]
  188. Rautio, J.; Kumpulainen, H.; Heimbach, T.; Oliyai, R.; Oh, D.; Järvinen, T.; Savolainen, J. Prodrugs: Design and clinical applications. Nat. Rev. Drug Discov. 2008, 7, 255–270. [Google Scholar] [CrossRef]
  189. Wolff, N.A.; Burckhardt, B.C.; Burckhardt, G.; Oellerich, M.; Armstrong, V.W. Mycophenolic acid (MPA) and its glucuronide metabolites interact with transport systems responsible for excretion of organic anions in the basolateral membrane of the human kidney. Nephrol. Dial. Transplant. 2007, 22, 2497–2503. [Google Scholar] [CrossRef]
  190. Park, B.K.; Boobis, A.; Clarke, S.; Goldring, C.E.P.; Jones, D.; Kenna, J.G.; Lambert, C.; Laverty, H.G.; Naisbitt, D.J.; Nelson, S.; et al. Managing the challenge of chemically reactive metabolites in drug development. Nat. Rev. Drug Discov. 2011, 10, 292–306. [Google Scholar] [CrossRef]
  191. Tateishi, Y.; Ohe, T.; Ogawa, M.; Takahashi, K.; Nakamura, S.; Mashino, T. Development of Novel Diclofenac Analogs Designed to Avoid Metabolic Activation and Hepatocyte Toxicity. ACS Omega 2020, 5, 32608–32616. [Google Scholar] [CrossRef] [PubMed]
  192. Guengerich, F.P. A history of the roles of cytochrome P450 enzymes in the toxicity of drugs. Toxicol. Res. 2021, 37, 1–23. [Google Scholar] [CrossRef]
  193. Dahlin, D.C.; Miwa, G.T.; Lu, A.Y.; Nelson, S.D. N-acetyl-p-benzoquinone imine: A cytochrome P-450-mediated oxidation pro-duct of acetaminophen. Proc. Natl. Acad. Sci. USA 1984, 81, 1327–1331. [Google Scholar] [CrossRef] [PubMed]
  194. Bender, R.P.; Lindsey, R.H., Jr.; Burden, D.A.; Osheroff, N. N-Acetyl-p-benzoquinone Imine, the Toxic Metabolite of Acetaminophen, Is a Topoisomerase II Poison. Biochemistry 2004, 43, 3731–3739. [Google Scholar] [CrossRef] [PubMed]
  195. Ertl, P.; Roggo, S.; Schuffenhauer, A. Natural Product-likeness Score and Its Application for Prioritization of Compound Libraries. J. Chem. Inf. Model. 2008, 48, 68–74. [Google Scholar] [CrossRef] [PubMed]
  196. Tay, D.W.P.; Yeo, N.Z.X.; Adaikkappan, K.; Lim, Y.H.; Ang, S.J. 67 million natural product-like compound database generated via molecular language processing. Sci. Data 2023, 10, 296. [Google Scholar] [CrossRef] [PubMed]
  197. Dobson, P.D.; Patel, Y.; Kell, D.B. Metabolite-likeness’ as a criterion in the design and selection of pharmaceutical drug libraries. Drug Discov. Today 2009, 14, 31–40. [Google Scholar] [CrossRef]
  198. O’Hagan, S.; Kell, D.B. Understanding the foundations of the structural similarities between marketed drugs and endogenous human metabolites. Front. Pharmacol. 2015, 6, 105. [Google Scholar] [CrossRef]
  199. O′Hagan, S.; Swainston, N.; Handl, J.; Kell, D.B. A ‘rule of 0.5’ for the metabolite-likeness of approved pharmaceutical drugs. Metabolomics 2015, 11, 323–339. [Google Scholar] [CrossRef]
  200. Ertl, P. Substituents of life: The most common substituent patterns present in natural products. Bioorg. Med. Chem. 2022, 54, 116562. [Google Scholar] [CrossRef]
  201. Shultz, M.D. Two Decades under the Influence of the Rule of Five and the Changing Properties of Approved Oral Drugs. J. Med. Chem. 2019, 62, 1701–1714. [Google Scholar] [CrossRef] [PubMed]
  202. Brown, D.G.; Wobst, J.J. A Decade of FDA-Approved Drugs (2010−2019): Trends and Future Directions. J. Med. Chem. 2021, 64, 2312–2338. [Google Scholar] [CrossRef] [PubMed]
  203. Young, R.J.; Flitsch, S.L.; Grigalunas, M.; Leeson, P.D.; Quinn, R.J.; Turner, N.J.; Waldmann, H. The Time and Place for Nature in Drug Discovery. JACS Au 2022, 2, 2400–2416. [Google Scholar] [CrossRef] [PubMed]
  204. Saha, S.; Rajpal, D.K.; Brown, J.R. Human microbial metabolites as a source of new drugs. Drug Discov. Today 2016, 21, 692–698. [Google Scholar] [CrossRef] [PubMed]
  205. Li, H.; Ranhotra, H.S.; Mani, S.; Dvořák, Z.; Sokol, H.; Müller, R. Human microbial metabolite mimicry as a strategy to expand the chemical space of potential drugs. Drug Discov. Today 2020, 25, 1575–1579. [Google Scholar] [CrossRef]
  206. Dvorák, Z.; Kopp, F.; Costello, C.M.; Kemp, J.S.; Li, H.; Vrzalová, A.; Stepánková, M.; Bartonková, I.; Jiskrová, E.; Poulíková, K.; et al. Targeting the pregnane X receptor using microbial metabolite mimicry. EMBO Mol. Med. 2020, 12, e11621. [Google Scholar] [CrossRef]
  207. Xue, Y.P.; Cao, C.H.; Zheng, Y.G. Enzymatic asymmetric synthesis of chiral amino acids. Chem. Soc. Rev. 2018, 47, 1516–1561. [Google Scholar] [CrossRef]
  208. Herger, M.; van Roye, P.; Romney, D.K.; Brinkmann-Chen, S.; Buller, A.R.; Arnold, F.H. Synthesis of β-branched tryptophan analogues using an engineered subunit of tryptophan synthase. J. Am. Chem. Soc. 2016, 138, 8388–8391. [Google Scholar] [CrossRef]
  209. Alfonzo, E.; Das, A.; Arnold, F.H. New Additions to the Arsenal of Biocatalysts for Noncanonical Amino Acid Synthesis. Curr. Opin. Green Sustain. Chem. 2022, 38, 100701. [Google Scholar] [CrossRef]
  210. Ellis, J.M.; Campbell, M.E.; Kumar, P.; Geunes, E.P.; Bingman, C.A.; Buller, A.R. Biocatalytic synthesis of non-standard amino acids by a decarboxylative aldol reaction. Nat. Catal. 2022, 5, 136–143. [Google Scholar] [CrossRef]
  211. Huffman, M.A.; Fryszkowska, A.; Alvizo, O.; Borra-Garske, M.; Campos, K.R.; Canada, K.A.; Devine, P.N.; Duan, D.; Forstater, J.H.; Grosser, S.T.; et al. Design of an in vitro biocatalytic cascade for the manufacture of islatravir. Science 2019, 366, 1255–1259. [Google Scholar] [CrossRef]
  212. Zetzsche, L.E.; Narayan, A.R.H. Broadening the scope of biocatalytic C–C bond formation. Nat. Rev. Chem. 2020, 4, 334–346. [Google Scholar] [CrossRef] [PubMed]
  213. Schneider, P.; Henßen, B.; Paschold, B.; Chapple, B.P.; Schatton, M.; Seebeck, F.P.; Classen, T.; Pietruszka, J. Biocatalytic C3-Indole Methylation—A Useful Tool for the Natural-Product-Inspired Stereoselective Synthesis of Pyrroloindoles. Angew. Chem. Int. Ed. 2021, 60, 23412–23418. [Google Scholar] [CrossRef]
  214. Fansher, D.J.; Palmer, D.R. A Type 1 Aldolase, NahE, Catalyzes a Stereoselective Nitro-Michael Reaction: Synthesis of β-Aryl-γ-nitrobutyric Acids. Angew. Chem. Int. Ed. 2023, 62, e202214539. [Google Scholar] [CrossRef]
  215. Walsh, C.T.; Tang, Y. Recent Advances in Enzymatic Complexity Generation: Cyclization Reactions. Biochemistry 2018, 57, 3087–3104. [Google Scholar] [CrossRef] [PubMed]
  216. Gao, L.; Su, C.; Du, X.; Wang, R.; Chen, S.; Zhou, Y.; Liu, C.; Liu, X.; Tian, R.; Zhang, L.; et al. FAD-dependent enzyme-catalysed intermolecular [4+2] cycloaddition in natural product biosynthesis. Nat. Chem. 2020, 12, 620–628. [Google Scholar] [CrossRef] [PubMed]
  217. Liu, X.; Yang, J.; Gao, L.; Zhang, L.; Lei, X. Chemoenzymatic Total Syntheses of Artonin I with an Intermolecular Diels-Alderase. Biotechnol. J. 2020, 15, 2000119. [Google Scholar] [CrossRef]
  218. Basler, S.; Studer, S.; Zou, Y.; Mori, T.; Ota, Y.; Camus, A.; Bunzel, H.A.; Helgeson, R.C.; Houk, K.N.; Jiménez-Osés, G.; et al. Efficient Lewis acid catalysis of an abiological reaction in a de novo protein scaffold. Nat. Chem. 2021, 13, 231–235. [Google Scholar] [CrossRef]
  219. Gao, L.; Zou, Y.; Liu, X.; Yang, J.; Du, X.; Wang, J.; Yu, X.; Fan, J.; Jiang, M.; Li, Y.; et al. Enzymatic control of endo- and exo-stereoselective Diels–Alder reactions with broad substrate scope. Nat. Catal. 2021, 4, 1059–1069. [Google Scholar] [CrossRef]
  220. Löwe, J.; Dietz, K.J.; Gröger, H. From a Biosynthetic Pathway toward a Biocatalytic Process and Chemocatalytic Modifications: Three-Step Enzymatic Cascade to the Plant Metabolite cis-(+)-12-OPDA and Metathesis-Derived Products. Adv. Sci. 2020, 7, 1902973. [Google Scholar] [CrossRef]
  221. Westarp, S.; Kaspar, F.; Neubauer, P.; Kurreck, A. Industrial potential of the enzymatic synthesis of nucleoside analogs: Existing challenges and perspectives. Curr. Opin. Biotechnol. 2022, 78, 102829. [Google Scholar] [CrossRef]
  222. Cosgrove, S.C.; Miller, G.J. Advances in biocatalytic and chemoenzymatic synthesis of nucleoside analogues. Expert Opin. Drug Discov. 2022, 17, 355–364. [Google Scholar] [CrossRef]
  223. McIntosh, J.A.; Benkovics, T.; Silverman, S.M.; Huffman, M.A.; Kong, J.; Maligres, P.E.; Itoh, T.; Yang, H.; Verma, D.; Pan, W.; et al. Engineered ribosyl-1-kinase enables concise synthesis of molnupiravir, an antiviral for COVID-19. ACS Cent. Sci. 2021, 7, 1980–1985. [Google Scholar] [CrossRef] [PubMed]
  224. McIntosh, J.A.; Liu, Z.; Andresen, B.M.; Marzijarani, N.S.; Moore, J.C.; Marshall, N.M.; Borra-Garske, M.; Obligacion, J.V.; Fier, P.S.; Peng, F.; et al. A kinase-cGAS cascade to synthesize a therapeutic STING activator. Nature 2022, 603, 439–444. [Google Scholar] [CrossRef] [PubMed]
  225. Trung, M.N.; Kieninger, S.; Fandi, Z.; Qiu, D.; Liu, G.; Mehendale, N.K.; Saiardi, A.; Jessen, H.; Keller, B.; Fiedler, D. Stable Isotopomers of myo-Inositol Uncover a Complex MINPP1-Dependent Inositol Phosphate Network. ACS Cent. Sci. 2022, 8, 1683–1694. [Google Scholar] [CrossRef] [PubMed]
  226. Shen, B. A New Golden Age of Natural Products Drug Discovery. Cell 2015, 163, 1297–1300. [Google Scholar] [CrossRef]
  227. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; Supuran, C.T. Natural products in drug discovery: Advances and opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [CrossRef]
  228. Reetz, M.T.; Sun, Z.; Qu, G. Enzyme Engineering: Selective Catalysts for Applications in Biotechnology, Organic Chemistry, and Life Science; Wiley-VCH: Weinhem, Germany, 2023; ISBN 978-3-527-35033-9. [Google Scholar]
  229. Chen, K.; Arnold, F.H. Engineering new catalytic activities in enzymes. Nat. Catal. 2020, 3, 203–213. [Google Scholar] [CrossRef]
  230. Zetzsche, L.E.; Chakrabarty, S.; Narayan, A.R. The transformative power of biocatalysis in convergent synthesis. J. Am. Chem. Soc. 2022, 144, 5214–5225. [Google Scholar] [CrossRef]
  231. Stout, C.N.; Wasfy, N.M.; Chen, F.; Renata, H. Charting the Evolution of Chemoenzymatic Strategies in the Syntheses of Complex Natural Products. J. Am. Chem. Soc. 2023, 145, 18161–18181. [Google Scholar] [CrossRef]
  232. Li, F.; Deng, H.; Renata, H. Chemoenzymatic approaches for exploring structure–activity relationship studies of bioactive natural products. Nat. Synth. 2023, 2, 708–718. [Google Scholar] [CrossRef]
  233. Lovelock, S.L.; Crawshaw, R.; Basler, S.; Levy, C.; Baker, D.; Hilvert, D.; Green, A.P. The road to fully programmable protein catalysis. Nature 2022, 606, 49–58. [Google Scholar] [CrossRef] [PubMed]
  234. Hermann, J.C.; Marti-Arbona, R.; Fedorov, A.A.; Fedorov, E.; Almo, S.C.; Shoichet, B.K.; Raushel, F.M. Structure-based activity prediction for an enzyme of unknown function. Nature 2007, 448, 775–779. [Google Scholar] [CrossRef] [PubMed]
  235. Zallot, R.; Oberg, N.; Gerlt, J.A. The EFI web resource for genomic enzymology tools: Leveraging protein, genome, and metagenome databases to discover novel enzymes and metabolic pathways. Biochemistry 2019, 58, 4169–4182. [Google Scholar] [CrossRef]
  236. Price, M.N.; Wetmore, K.M.; Waters, R.J.; Callaghan, M.; Ray, J.; Liu, H.; Kuehl, J.V.; Melnyk, R.A.; Lamson, J.S.; Suh, Y.; et al. Mutant phenotypes for thousands of bacterial genes of unknown function. Nature 2018, 557, 503–509. [Google Scholar] [CrossRef]
  237. Robinson, S.L.; Piel, J.; Sunagawa, S. A roadmap for metagenomic enzyme discovery. Nat. Prod. Rep. 2021, 38, 1994–2023. [Google Scholar] [CrossRef]
  238. Caputi, L.; Franke, J.; Farrow, S.C.; Chung, K.; Payne, R.M.; Nguyen, T.D.; Dang, T.T.T.; Soares Teto Carqueijeiro, I.; Koudounas, K.; Dugé de Bernonville, T.; et al. Missing enzymes in the biosynthesis of the anticancer drug vinblastine in Madagascar periwinkle. Science 2018, 360, 1235–1239. [Google Scholar] [CrossRef]
  239. Blin, K.; Shaw, S.; Augustijn, H.E.; Reitz, Z.L.; Biermann, F.; Alanjary, M.; Fetter, A.; Terlouw, B.R.; Metcalf, W.W.; Helfrich, E.J.N.; et al. antiSMASH 7.0: New and improved predictions for detection, regulation, chemical structures and visualisation. Nucleic Acids Res. 2023, 51, W46–W50. [Google Scholar] [CrossRef]
  240. Caesar, L.K.; Montaser, R.; Keller, N.P.; Kelleher, N.L. Metabolomics and genomics in natural products research: Complementary tools for targeting new chemical entities. Nat. Prod. Rep. 2021, 38, 2041–2065. [Google Scholar] [CrossRef]
  241. Scherlach, K.; Hertweck, C. Mining and unearthing hidden biosynthetic potential. Nat. Commun. 2021, 12, 3864. [Google Scholar] [CrossRef]
  242. Klapper, M.; Hübner, A.; Ibrahim, A.; Wasmuth, I.; Borry, M.; Haensch, V.G.; Zhang, S.; Al-Jammal, W.K.; Suma, H.; Fellows Yates, J.A.; et al. Natural products from reconstructed bacterial genomes of the Middle and Upper Paleolithic. Science 2023, 380, 619–624. [Google Scholar] [CrossRef] [PubMed]
  243. Caesar, L.K.; Butun, F.A.; Robey, M.T.; Ayon, N.J.; Gupta, R.; Dainko, D.; Bok, J.W.; Nickles, G.; Stankey, R.J.; Johnson, D.; et al. Correlative metabologenomics of 110 fungi reveals metabolite–gene cluster pairs. Nat. Chem. Biol. 2023, 19, 846–854. [Google Scholar] [CrossRef] [PubMed]
  244. Smanski, M.J.; Zhou, H.; Claesen, J.; Shen, B.; Fischbach, M.; Voigt, C.A. Synthetic biology to access and expand nature’s chemical diversity. Nat. Rev. Microbiol. 2016, 14, 135–149. [Google Scholar] [CrossRef] [PubMed]
  245. Erb, T.J.; Jones, P.R.; Bar-Even, A. Synthetic metabolism: Metabolic engineering meets enzyme design. Curr. Opin. Chem. Biol. 2017, 37, 56–62. [Google Scholar] [CrossRef]
  246. Yi, J.; Li, Z. Artificial multi-enzyme cascades for natural product synthesis. Curr. Opin. Biotechnol. 2022, 78, 102831. [Google Scholar] [CrossRef]
Figure 1. Selected reasons for the synthesis of metabolites and metabolite-like compounds. (A) For academic research applications in basic and applied sciences, where small quantities are usually needed, for example in biochemistry and other life sciences for the analysis of metabolic pathways, the discovery of unknown biological pathways, in preclinical research and development, and medical and diagnostic applications. (B) For applications in industry, where usually large quantities are needed of active ingredients and intermediates of pharmaceuticals, vitamins, flavors, fragrances, dyes, or agrochemicals. In clinical chemistry, standards are needed for blood and urine tests of biomarkers for diagnostic purposes and therapeutic drug monitoring or drug abuse. In medicinal chemistry, a diversity of test compounds and standards are needed for studying biological activity and efficacy, metabolism, and pharmacokinetics, and for determining the parameters of absorption, distribution, metabolism, excretion, and toxicity of novel molecular entities.
Figure 1. Selected reasons for the synthesis of metabolites and metabolite-like compounds. (A) For academic research applications in basic and applied sciences, where small quantities are usually needed, for example in biochemistry and other life sciences for the analysis of metabolic pathways, the discovery of unknown biological pathways, in preclinical research and development, and medical and diagnostic applications. (B) For applications in industry, where usually large quantities are needed of active ingredients and intermediates of pharmaceuticals, vitamins, flavors, fragrances, dyes, or agrochemicals. In clinical chemistry, standards are needed for blood and urine tests of biomarkers for diagnostic purposes and therapeutic drug monitoring or drug abuse. In medicinal chemistry, a diversity of test compounds and standards are needed for studying biological activity and efficacy, metabolism, and pharmacokinetics, and for determining the parameters of absorption, distribution, metabolism, excretion, and toxicity of novel molecular entities.
Metabolites 13 01097 g001
Figure 2. Basic types of biocatalytic systems for the synthesis of metabolites and metabolite-like compounds.
Figure 2. Basic types of biocatalytic systems for the synthesis of metabolites and metabolite-like compounds.
Metabolites 13 01097 g002
Figure 3. Selected naturally occurring metabolites with demonstrated total synthesis.
Figure 3. Selected naturally occurring metabolites with demonstrated total synthesis.
Metabolites 13 01097 g003
Figure 4. Structures of vitamin B12 and coenzyme B12.
Figure 4. Structures of vitamin B12 and coenzyme B12.
Metabolites 13 01097 g004
Figure 5. Complete stereochemistry and planar structure of palytoxin.
Figure 5. Complete stereochemistry and planar structure of palytoxin.
Metabolites 13 01097 g005
Figure 6. Selected naturally occurring metabolites produced through bioprocesses at an industrial large scale.
Figure 6. Selected naturally occurring metabolites produced through bioprocesses at an industrial large scale.
Metabolites 13 01097 g006
Figure 7. Plant cell fermentation for the industrial production of paclitaxel.
Figure 7. Plant cell fermentation for the industrial production of paclitaxel.
Metabolites 13 01097 g007
Figure 8. Enzymatic process for the industrial production of the fragrance ingredient (-)-Ambrox.
Figure 8. Enzymatic process for the industrial production of the fragrance ingredient (-)-Ambrox.
Metabolites 13 01097 g008
Figure 9. Biocatalytic synthesis of selected naturally occurring metabolites of the Emden–Meyerhof–Parnas, the Entner–Doudoroff, and the pentose phosphate pathways.
Figure 9. Biocatalytic synthesis of selected naturally occurring metabolites of the Emden–Meyerhof–Parnas, the Entner–Doudoroff, and the pentose phosphate pathways.
Metabolites 13 01097 g009
Figure 10. Selected biocatalytic systems for the synthesis of metabolites labelled with 2H-, 13C-, or 15N-stable isotopes.
Figure 10. Selected biocatalytic systems for the synthesis of metabolites labelled with 2H-, 13C-, or 15N-stable isotopes.
Metabolites 13 01097 g010
Figure 11. Selected biocatalytic systems for pharmaceutical drug metabolite synthesis.
Figure 11. Selected biocatalytic systems for pharmaceutical drug metabolite synthesis.
Metabolites 13 01097 g011
Figure 12. Biocatalytic synthesis of selected metabolite-like compounds.
Figure 12. Biocatalytic synthesis of selected metabolite-like compounds.
Metabolites 13 01097 g012
Table 1. Selection of Species-specific Metabolite Databases.
Table 1. Selection of Species-specific Metabolite Databases.
Biological
Species
Name of Species-
Specific
Metabolite Database
Abbreviation
of Metabolite
Database Name
Website of
Metabolite Database
ReferenceAccessed Date
for the URL
HumanHuman Metabolome
Database
HMDBhttps://hmdb.ca/[14]accessed on 16 July 2023
Human
Microbiome
Human Microbial
Metabolome Database i
MiMeDBhttps://mimedb.org/[15]accessed on 16 July 2023
Escherichia coliEscherichia coli
Metabolome Database
ECMDB http://www.ecmdb.ca/[16]accessed on 16 July 2023
Pseudomonas
aeruginosa
Pseudomonas aeruginosa
Metabolome Database
PAMDBhttp://pseudomonas.umaryland.edu/[17]accessed on 16 July 2023
Streptomyces sp. Streptomyces Natural
Products Database
StreptomeDBhttp://www.pharmbioinf.uni-freiburg.de/streptomedb/[18]accessed on 30 July 2023
CyanobacteriaComprehensive database
of secondary metabolites
from cyanobacteria
CyanoMetDBhttps://zenodo.org/record/7922070/[19]accessed on 30 July 2023
MyxobacteriaMyxobacterial Natural
Product Database
MyxoDBhttps://www.myxonpdb.sdu.edu.cn/[20]accessed on 30 July 2023
YeastYeast Metabolome
Database
YMDBhttp://www.ymdb.ca/[21]accessed on 16 July 2023
BovineBovine Metabolome
Database
BMDBhttps://bovinedb.ca/[22]accessed on 16 July 2023
TomatoTomato Metabolome
Database
TOMATOMEThttp://metabolites.in/tomato-fruits/[23]accessed on 24 July 2023
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wohlgemuth, R. Synthesis of Metabolites and Metabolite-like Compounds Using Biocatalytic Systems. Metabolites 2023, 13, 1097. https://doi.org/10.3390/metabo13101097

AMA Style

Wohlgemuth R. Synthesis of Metabolites and Metabolite-like Compounds Using Biocatalytic Systems. Metabolites. 2023; 13(10):1097. https://doi.org/10.3390/metabo13101097

Chicago/Turabian Style

Wohlgemuth, Roland. 2023. "Synthesis of Metabolites and Metabolite-like Compounds Using Biocatalytic Systems" Metabolites 13, no. 10: 1097. https://doi.org/10.3390/metabo13101097

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop