ACS Publications. Most Trusted. Most Cited. Most Read
Enabling Broader Adoption of Biocatalysis in Organic Chemistry
My Activity
  • Open Access
Perspective

Enabling Broader Adoption of Biocatalysis in Organic Chemistry
Click to copy article linkArticle link copied!

  • Evan O. Romero
    Evan O. Romero
    Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
  • Anthony T. Saucedo
    Anthony T. Saucedo
    Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
  • José R. Hernández-Meléndez
    José R. Hernández-Meléndez
    Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
  • Di Yang
    Di Yang
    Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
    More by Di Yang
  • Suman Chakrabarty
    Suman Chakrabarty
    Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
  • Alison R. H. Narayan*
    Alison R. H. Narayan
    Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
    *Email: [email protected]. Phone: +1 (734) 615-5505.
Open PDF

JACS Au

Cite this: JACS Au 2023, 3, 8, 2073–2085
Click to copy citationCitation copied!
https://doi.org/10.1021/jacsau.3c00263
Published July 19, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Biocatalysis is becoming an increasingly impactful method in contemporary synthetic chemistry for target molecule synthesis. The selectivity imparted by enzymes has been leveraged to complete previously intractable chemical transformations and improve synthetic routes toward complex molecules. However, the implementation of biocatalysis in mainstream organic chemistry has been gradual to this point. This is partly due to a set of historical and technological barriers that have prevented chemists from using biocatalysis as a synthetic tool with utility that parallels alternative modes of catalysis. In this Perspective, we discuss these barriers and how they have hindered the adoption of enzyme catalysts into synthetic strategies. We also summarize tools and resources that already enable organic chemists to use biocatalysts. Furthermore, we discuss ways to further lower the barriers for the adoption of biocatalysis by the broader synthetic organic chemistry community through the dissemination of resources, demystifying biocatalytic reactions, and increasing collaboration across the field.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2023 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Throughout history, humans have employed complex molecules for therapeutic purposes. In modern times, the ability to synthesize these molecules through chemical methods has expanded access to both naturally occurring and synthetic medicinal compounds. (1,2) As a result, synthetic strategies toward complex molecules have evolved tremendously over the past two centuries to arrive at a level of sophistication that directly impacts medicine, agriculture, and other industries. (3,4) People have used enzymes through fermentation since ancient history, but in the 21st century, especially, the use of enzymes for selective synthesis of complex molecules has expanded. (5−7) In particular, the use of enzymes in chemical synthesis has seen a resurgence in the past decade with increasing demand for cleaner and greener reactions. (8) The potential of biocatalysis in complex molecule synthesis is aptly described by Professor Brian Stoltz, a leader and practitioner of organic synthesis: “As a synthetic organic chemist primarily interested in accessing complex structures, I believe that we need to be open to using any and all tools available to build the bonds we desire to make. Enzymes and biocatalysts offer a unique and often orthogonal set of instruments to accomplish this molecular surgery. Recent advances in enzyme engineering bolster our ability to employ these catalysts in our day-to-day laboratory experience. I can only imagine that our toolset of enzymes and biocatalytic transformations will greatly increase in the future.”
The precision and efficiency possible with enzyme catalysts has led to the incorporation of biocatalytic strategies into synthetic campaigns in both academic and industrial spheres. (9,10) Enzymes can operate under ambient conditions and possess the potential for high chemo- and site-selectivity which can override the need to protect nonparticipating functional groups in complex substrates. (11,12) Additionally, the ambient and low-temperature reaction conditions typical for enzymatic reactions enhance their suitability for academic and industrial settings by improving the safety profile, reducing the energy required, and maximizing the overall procedural simplicity. (11−14) Consequently, biocatalytic methods are increasingly being embraced by mainstream organic chemists. (6,7,15−17) The advantages of leveraging enzymes in synthesis can be seen in academic settings, with chemoenzymatic synthesis becoming another tool for elegant total syntheses of complex molecules. (7,14,18−25) For example, biocatalysis has been used to form challenging C–C bonds (Figure 1A), (7,17,19) to introduce C–O bonds in a selective fashion (Figure 1B–D), and to carry out other difficult transformations. (14,21,22,24−26) Furthermore, biocatalytic methods are also becoming commonplace on process scale in the industrial production of pharmaceutical compounds, which is exemplified by Merck’s multienzyme cascade synthesis of islatravir (18) among a growing set of contemporary examples (Figure 1E). (13,14,27,28)

Figure 1

Figure 1. Select examples of chemical structures accessed by using biocatalysis. (A) Compounds formed through C–C bond forming reactions. (B) Compounds accessed using C–H hydroxylation reactions. (C) Hydroxylative dearomatization in the total synthesis of azaphilone natural products. (D) Amino acid C–H hydroxylation in the synthesis of manzacidin C. (E) Multienzyme cascade toward the process-scale total synthesis of islatravir.

Despite these elegant examples, the organic chemistry community’s overall adoption of biocatalytic methods has been slow; however, with emerging technological advances, biocatalysis is becoming a tool that anyone can integrate into their syntheses. In this Perspective, we provide an analysis of how the use of enzymes in organic chemistry has evolved and become accessible to chemists in the lab. We begin our discussion by highlighting the historical use of enzymes and the technological barriers that have slowed the implementation of enzymatic catalysis in organic chemistry. We then provide insight into potential solutions to these barriers by highlighting powerful resources that currently exist and new steps that can be implemented to broaden the incorporation of biocatalytic and hybrid chemoenzymatic tools in mainstream chemical synthesis.

2. Historical Barriers for Biocatalysis

Click to copy section linkSection link copied!

2.1. Biocatalysis throughout History

Enzymatic methods have played a pivotal role in the food and beverage industries throughout history, primarily through fermentation processes relying on microorganisms. (29) However, it was not until the 1800s that scientists began to uncover the chemical processes occurring within microorganisms. Crucially, in 1897, Eduard Buchner demonstrated that yeast cell-free extracts could still produce alcohol from glucose, indicating that living cells were not necessary for the process. (30) This discovery was essential in identifying the biomolecules responsible for these reactions and paved the way for further studies of enzyme-mediated processes. Further discoveries in molecular biology and enzymology propelled the field forward as individual enzymes were identified and their mechanisms were uncovered. (31)
Even with these advances, a distinct limitation of the application of biocatalysis in organic chemistry was the accessibility of enzymes capable of performing the desired reactions (Figure 2A). Traditionally, enzymes were accessed using the cells or cell-free extract of the organism where the enzyme is found natively. (32) This meant that researchers would have to obtain the specific strain of an organism. For example, in 1981 Oliveri and co-workers found that a strain of Agrobacterium radiobacter could perform two sequential stereospecific hydrolyses of racemic hydantoins to produce d-amino acids (Figure 2B), molecules which were difficult to access with chemical methods. (33) These products were highly useful for synthesizing antibiotics; however, this enzymatic reaction required a specific microorganism, limiting the widespread applicability of this work. (34)

Figure 2

Figure 2. (A) Historical access to enzymes and enzyme products was a time-consuming process. The understanding of biological systems and the lack of enabling technologies make it difficult to efficiently develop new biocatalysts. (B) Example of an early application of biocatalysts in the synthesis of d-amino acids. This process required the use of a specific strain of bacteria to complete the transformation.

Although enzyme access could be a barrier for research, the mid- to late-1900s saw an increase in research in this field. Seminal examples of enzyme catalysis include dynamic resolutions, hydrolyses, oxidations, reductions, and ligations. (31) Around this time, more companies also began to sell enzymes extracted from organisms, making biocatalytic transformations slightly more accessible. (31) Scientists also found that individual enzymes could be immobilized to produce desired products more efficiently, such as in the industrial manufacture of penicillanic acid. (35)
Despite this early adoption of enzymes in the production of valuable molecules, the broader implementation of biocatalysis in synthesis was slow due in part to the time it took to discover and understand new enzymes. This is exemplified in the discovery and subsequent understanding of the biosynthesis of morphine, a vital pain medication that has been used for centuries. (36) The structure of morphine was first determined in 1925, (37) but the enzymes involved in its biosynthesis were not elucidated until the 1960s. (36) Clearly, Nature had assembled a group of powerful enzymes that could stitch together a highly complex structure, but researchers did not have the resources or understanding to determine precisely how this was accomplished.
Overall, enzymes have been employed extensively throughout history. Foundational advances in enzymology allowed for specific enzymes to be used for synthetic transformations and laid the groundwork for biocatalysis in organic synthesis.

2.2. Effects of Technological Barriers

In the past, limited knowledge of enzyme synthesis, structure, and function has been a significant barrier to the advancement of biocatalysis. Because enzymes could only be accessed from specific organisms, the breadth of enzymes that could be characterized was meager compared to the millions of proteins that are present in Nature. (38) Furthermore, many of the reports of enzymes throughout history have focused on determining the function of various enzymes, with little emphasis on the breadth of individual enzyme’s capabilities. (31) Consequently, there has been a distinct lack of data covering the overall reactivity of enzymes. This information gap has traditionally hindered the applications of enzymes in organic synthesis as it limits the known starting points for integrating enzymes into synthetic campaigns.
Paralleling this lack of enzyme reactivity and the limited examples of enzymes in organic synthesis, biocatalysis has not traditionally been incorporated into organic chemistry education. Even though enzymes can offer distinct advantages in organic synthesis, they rarely find their way into an organic chemistry classroom or introductory laboratory course. (39) Consequently, students have limited exposure to this valuable strategy of building molecules. This may cause developing scientists to view biocatalysis as a niche tool rather than one they could use in the laboratory.

3. Biocatalysis Today

Click to copy section linkSection link copied!

Although the aforementioned barriers have historically limited the access to enzymes and their application in organic synthesis laboratories, within the past 50 years there has been significant progress in how we conduct research with enzymes. Many of these advances are centered around determining DNA sequence, access to DNA for genes of interest, and methods for producing enzymes efficiently. (31) The technological innovations developed for these purposes have drastically expanded our knowledge of enzymes and lowered the barriers for obtaining an enzyme of interest.
Knowing the exact sequence of a molecule of DNA was once a massive hurdle; however, the development of robust sequencing technologies means that an exponentially growing number of genomes are publicly available, and that DNA sequencing can be done within individual laboratories, core facilities, and a growing number of companies. (40) As a result, researchers can send a DNA sample to one of the many companies that conduct the sequencing and have sequence data in 1–2 days.
These advances in DNA sequencing technology have led to an explosion of publicly available genetic information. Partial or entire organism genome sequences are consistently being added to online databases, resulting in billions of individual gene sequence entries. (41,42) These DNA sequences of individual annotated genes can be automatically translated to afford the sequence of individual proteins. In most cases, these protein sequences can then be automatically assigned to a protein family based on their similarity to previously characterized proteins. In addition, protein sequences are housed in online databases, such as UniProt, where users can access the sequence and any other information about the proteins. (43) As a result, there is now a wealth of data that researchers can access when choosing enzymes for synthesis.
We recognize that the experimental techniques and equipment used to access biocatalysts and run reactions may be an unknown territory for many synthetic chemists. Below, we outline the basic steps to acquire a biocatalyst and to run biocatalytic reactions (Figure 3). This is not a comprehensive tutorial, but it is intended to provide helpful context to synthetic chemists that may not have previously interacted with this field. Additionally, as there are an increasing number of technologies and companies that are lowering barriers within this process, different entry points will be highlighted to emphasize the growing accessibility of obtaining enzymes.

Figure 3

Figure 3. Accessing biocatalysts with today’s methods. (A) General workflow for producing enzymes from the gene encoding for an enzyme of interest. The various entry points where a scientist could step into the process are highlighted. (B) Enzymes can be used in biocatalytic reactions at various levels of purity.

3.1. Access to DNA for Genes of Interest

The connection between genetic information and enzymes is crucial for understanding and manipulating enzymes. It was not until the mid-20th century that the link between DNA and proteins was established. This discovery was a major breakthrough in the field of molecular biology as it enabled scientists to understand how genetic information is encoded in DNA and translated into proteins. Even after this discovery, the efficient synthesis of DNA remained a significant challenge, as the technologies to enable this had not yet been developed.
That all changed with the advancement in DNA synthesis technologies such as microarray-based gene synthesis, (44) Gibson assembly, (45) as well as engineering of Terminal deoxynucleotidyl Transferase (TdT) (46) DNA that encode biocatalysts of interest which can be synthesized with high efficiency and high throughput. The DNA fragments can then be cloned into an expression vector of choice for downstream recombinant protein production. Several companies provide DNA synthesis services at low cost and high efficiency. DNA fragments from 300 base pairs up to 10 kilobase pairs can be synthesized, cloned into expression vectors, and delivered to users within a short time frame. At the time of publication of this Perspective, a 1000 base pair gene cloned into a standard expression vector would cost approximately $200.

3.2. Transformation and Heterologous Expression

Heterologous protein expression (also known as recombinant protein expression) can be performed to produce an enzyme of interest. First, the gene that encodes the biocatalyst of interest must be cloned into an expression vector to construct a plasmid, a self-replicating circular DNA with specific selection markers that can be used for gene expression. The gene can then be delivered into the host of choice through transformation, transfection, or transduction. (47) Common heterologous protein expression hosts include Escherichia coli (E. coli), (48) yeast (Pichia pastoris), (49) and mammalian cells. (50) Cells carrying the gene of interest can be selected using the selection marker and protein production can be induced by the addition of a specific chemical or by using self-inducing media. (51) All of the cell strains that are required for heterologous protein expression and the chemicals needed for selection and induction are commercially available, making the process streamlined and straightforward to execute.
This method for producing enzymes represents a significant improvement compared with the traditional way of accessing enzymes, which required isolation from the native organism. Heterologous protein expression allows for a standardized way of producing large quantities of protein. An emerging method for protein production is cell-free expression, where cell lysate containing the molecular machinery for transcription and translation is used in vitro with DNA as the template. This strategy may offer advantages such as more efficient protein production and enhanced control overexpression conditions. (52) The development of protein production methods has been highly enabling in biocatalysis research as it gives researchers a way to rapidly access a protein of interest.

3.3. Preparing Biocatalysts for Reactions

Once cultures have been grown and used to produce the desired protein, it is time to decide how to use the catalyst (Figure 3B). A biocatalyst can be added to reactions in the form of whole cells, cell lysate, or purified protein with each of these methods successively increasing the time and equipment required. The simplest method is to leave the enzyme in the cells and use the whole cell to deliver the catalyst. This method does not require any additional equipment and uses minimal resources. (53) Some situations require the enzyme to be present in solution, which can be achieved by lysing cells mechanically or chemically. This produces what is known as a crude cell lysate, which is a common form of enzyme used in biocatalytic reactions. The lysate can also be lyophilized to produce a powder that can be easier to handle and store. (54) Finally, the most rigorous way to prepare the catalyst is to purify the enzyme from the lysate solution. This is commonly achieved by using affinity chromatography. An example of this is using a Nickel resin that binds a polyhistidine tag which is appended to the enzyme of interest, followed by washing and eluting steps that result in purified protein. (55) Having your target enzyme in the purified form allows for direct quantification of the amount of enzyme and eliminates potentially problematic impurities. That said, the required time and equipment needed to obtain purified protein create a barrier to using protein in this form. Fortunately, running reactions in a whole cell or crude cell lysate format can often be equally as effective as a pure enzyme. (53,54)

3.4. Entry Points for Accessing Biocatalysts

There are numerous entry points available for researchers to participate in the process of obtaining an enzyme. If an enzyme is not commercially available, it can be produced from the DNA that encodes the enzyme of interest. This process starts by obtaining the gene (“Entry Point 1”, Figure 3A), which can be custom ordered from several different vendors. Next, the gene can be inserted into a plasmid using the cloning process described above to give the recombinant plasmid.
Recombinant plasmids can also be readily obtained from commercial sources (“Entry Point 2”, Figure 3A), through either a plasmid repository or custom-made services. Plasmid repositories, such as Addgene, have an inventory of plasmids containing many enzymes of interest. These plasmids are often stocked by research laboratories that want to make their plasmids available to the broader research community. For example, the Addgene repository has over 100,000 plasmids available for purchase. Although these repositories have many plasmids, an enzyme of interest may not be available in a presynthesized plasmid. Fortunately, with the recent advances in gene synthesis, (56) there are now several companies, such as Twist Bioscience and Gene Script, that can routinely make and provide synthetic plasmid within a few weeks. This growing set of vendors removes several technical biochemical steps and is facilitating a wider use of biocatalysis.
With the recombinant plasmid in hand, the process can continue with transformation of the host organism. These transformed recombinant cells can be used directly or stored as glycerol stocks that can be later used to produce enzymes. If an enzyme is discussed in the literature, the research group may have a glycerol stock containing a plasmid for the enzyme of interest that they would be willing to send to other groups (“Entry Point 3”, Figure 3A). In this case, a simple request may allow a researcher to quickly access a strain capable of producing the enzyme of interest.
Some enzymes can be obtained directly from a chemical vendor, similar to ordering a commercially available transition metal catalyst (“Entry Point 4”, Figure 3A). Several companies offer a range of enzymes such as reductases, dehydrogenases, transaminases, and lipases. These do not require researchers to produce the biocatalysts themselves, and instead receive the enzyme as a lyophilized powder that can be directly used in synthesis. These collections of enzymes cover the most commonly used enzyme classes and can be sold in kits that allow nonexperts to screen for biocatalysts that could be applied in valuable functional group interconversions. Furthermore, some companies, such as Prozomix, allow academic researchers to license their enzymes to be sold. This allows commercial access to enzymes that are not typically available through chemical vendors.

3.5. Carrying out and Optimizing Enzymatic Reactions

With protein in hand, setting up a biocatalytic reaction is relatively straightforward and is analogous to a standard “dump and stir” small-molecule chemical reaction. In general, the process is to add buffered solvent, substrate, cosolvent (if necessary), cofactors, and the biocatalyst to a flask. (57) As biocatalytic reactions are often done under ambient conditions in aqueous solvent, rigorous glassware drying or solvent preparation is often unnecessary. Reactions can easily be tracked by common analysis techniques such as thin-layer chromatography, high-performance liquid chromatography, and NMR. (58−60)
Optimizing an enzymatic reaction can be as simple as screening reaction component concentrations, time, and temperature─just as one would for a small-molecule organic reaction. Enzymes can be sensitive and easily denatured outside of their optimal reaction environments (i.e., temperature, pH, solvent, etc.), (61,62) and these factors can often be addressed with simple variations of the reaction conditions. Additionally, new tools and methods are being developed to improve the way biocatalytic reactions are set up. For example, enzyme immobilization is a tool that has enabled the use of biocatalysis in a variety of commonplace organic solvents. (63) Furthermore, this technique has allowed for increased thermal stability, use in one-pot multistep processes, easier catalyst reuse, and improved purification of reaction products. (64) The use of these immobilization techniques has even extended to continuous flow methods, which have extensively demonstrated the robustness of these catalysts in both academic and industrial applications. (65)
In some cases, a given substrate may not be compatible with biocatalytic reaction conditions or with the enzyme itself. For example, organic substrates may have limited solubility in water, but this can often be solved by using a cosolvent or by modulating the substrate to be more soluble in water. In some instances, even poorly water-soluble substrates can still deliver high yields on process scales. (66) In addition, if the enzyme is inefficient at catalyzing a desired reaction, the substrate can sometimes be engineered to more closely match the native substrate of the enzyme. (67) This is straightforward for a synthetic laboratory to achieve and requires only limited biology knowledge and resources.
When developing small-molecule methods, substrate generality is often a point of emphasis. Conversely, in Nature, chemical transformations are often highly selective and enzymes may struggle to transform substrates outside of those that structurally resemble their native substrate. (62) For many synthetic organic chemists, this poses a great barrier to developing methods and syntheses using biocatalysis. However, many tools have recently been developed that may help researchers choose and optimize an enzyme’s reactivity. Some of these tools are summarized below in section 4, and many can be used by a researcher with limited biochemical knowledge.
Another established tool for improving a biocatalytic reaction is enzyme engineering and directed evolution. (68) This powerful strategy has been used to improve enzyme properties, increase selectivity, increase catalytic efficiency, and optimize non-native functions. (69) Unfortunately, enzyme engineering requires specialized technical skills in molecular biology, (70) which makes this approach less accessible to synthetic chemists seeking an improved biocatalyst. There are several companies that can assist with engineering campaigns, and these organizations will surely become more commonplace in the near future. Additionally, many academic groups are conducting research in biocatalyst engineering and are open to collaborations.

3.6. Reaction Miniaturization

Because biocatalysis spans chemistry and biochemistry, a host of tools are available from both fields that can be used to accelerate research. In particular, biochemistry laboratories commonly use workflows that enable researchers to study systems on a very small scale. This allows for high-throughput experiments and rapid analysis of a large set of conditions. Analogously, biocatalytic reactions can easily be miniaturized and subsequently screened to identify ideal conditions, enzymes, or substrates. (71) A common way to accomplish this is by using 96-well plates, (72) but systems utilizing 384-well plates, (73) and even microfluidics (74) have also been demonstrated. These high-throughput platforms allow for other enabling technologies, such as liquid-handling robotics, (75) that have greatly accelerated the pace at which competent biocatalysts can be developed.

3.7. Equipment Requirements

An essential element of performing science in a laboratory is having the necessary equipment. Very little synthetic chemistry can be done without a functioning rotary evaporator, fume hood, or even something as simple as a stir plate. Biocatalysis is not an exception when it comes to the need for specific equipment; however, we acknowledge that this extra expense may be a hindrance for groups interested in using biocatalysts. In some cases, standard laboratory equipment can serve as a suitable substitution for biocatalytic reactions (such as a flask/hot plate instead of an incubator shaker). In situations requiring more technical biochemical equipment, it is likely that other laboratories in most chemistry and biology departments have the basic equipment described in this section. In this case, only a knock on the door or a quick email to a colleague could simplify the way synthetic chemists can access the wide range of reactivity enzymes offer.

4. Current Resources Aiding Biocatalysis

Click to copy section linkSection link copied!

4.1. Resources for Choosing Biocatalysts

Bioinformatics and computational resources have greatly assisted in the rapid development and data storage of enzyme-mediated reactions. Using tools such as online databases, synthetic organic chemists have rapid and thorough access to increasingly prevalent information about biocatalytic reactions, as they would with traditional synthetic method reports.
To aid synthetic planning and adoption, information about biocatalytic methods should be accessible through existing databases currently utilized by organic chemists. In recent years, two prominent databases, Reaxys and SciFinder, have formatted their search engine features to allow specific searches for biocatalytic transformations. For example, when filtering through reaction results in Reaxys, users can specify “catalyst classes” and limit search results to “organism/enzymes.” Within this specification, users can select the reaction type (i.e., oxidation/reduction, hydrolysis, isomerization, etc.) they would like to investigate. Although SciFinder’s enzymatic reaction search is not as detailed as Reaxys’, this platform does allow its users to utilize a “Biosequence Search” tool, which enables one to search through vast amounts of sequence data (over 500 million biosequences to date). Although the integration of information relating to biocatalytic reactions into these databases is helpful to organic chemists, the depth of information relating to the enzymes necessary for these reactions is lackluster compared to nonbiological reagents. In order to make these biological tools more accessible to chemists, details related to areas like preparation, purification, physical data, source, etc. should be readily available through Reaxys’ and SciFinder’s platforms.
Outside of these platforms familiar to synthetic chemists, other impactful database tools seek to better incorporate biocatalysis into synthetic organic chemistry (Table 1). One example is RetroBioCat, a tool largely curated by William Finnigan and co-workers from Nicholas Turner’s group at the University of Manchester. (75) This online tool provides users with two efficient methods for designing biocatalytic reactions: retrosynthetic planning and known enzyme reactivity tools. Through the retrosynthetic planning tool, users submit a target compound’s SMILES string and generate a network of possible precursors that can be modified via known enzymatic transformations. Additionally, the pathway explorer tool allows users to submit a SMILES string, from which the program will generate pathways to the compound of interest over several steps set by the user. From this, the program ranks each generated pathway based on parameters that the user can also set.
Table 1. Useful Online Resources for Biocatalysis
databasedescription
RetroBioCat (retrobiocat.com)aids in design of biocatalytic reactions/cascades through retrosynthetic approach
EFI-EST/Cytoscape (efi.igb.illinois.edu/efi-est/; cytoscape.org)enables users to generate SSNs and view these generated networks for identification of enzyme homologues and orthologues
PrenDB (prendb.pharmazie.uni-marburg.de/prendb/home)enzyme database with information related to prenyltransferase enzymes
BioCatNet Databases (www.biocatnet.de)database with sequence, structure, and biocatalytic data pertaining to a variety of protein families
BioCyc Collection of Pathway/Genome Databases (www.biocyc.org)database containing extensive sequence data and a variety of bioinformatic tools
Expasy (www.expasy.org)a wide collection of bioinformatic prediction and analysis tools
MACiE (www.ebi.ac.uk/thornton-srv/m-csa)database with extensive details pertaining to enzymatic reaction mechanisms
GTD (randr.nist.gov/enzyme/Default.aspx)database providing details related to thermodynamic parameters of enzymatic reactions
UniProt (www.uniprot.org)database with information pertaining to structure and function of proteins
Protein Data Bank (www.rcsb.org)database that provides reported 3D structures of proteins
EAWAG-BBD (eawag-bbd.ethz.ch/index.html)database with information regarding microbial biocatalytic reactions and biodegradation pathways for chemical compounds
ExplorEnz (www.enzyme-database.org)enzyme database that emphasizes enzyme nomenclature and classification
ESTHER (bioweb.supagro.inra.fr/ESTHER/general?what=index)enzyme database with information related to superfamily of alpha/beta-hydrolases
MEROPS (www.ebi.ac.uk/merops/index.shtml)enzyme database with information related to peptidase enzymes
Lipase database (www.led.uni-stuttgart.de)enzyme database with information related to lipase enzymes
CAZy (www.cazy.org)enzyme database with information related to carbohydrate-active enzymes
RedoxiBase (peroxibase.toulouse.inra.fr)enzyme database with information related to oxidoreductase enzymes
In addition to its retrosynthetic planning tools, RetroBioCat offers an enzyme database that enables users to explore known enzymatic reactions and to explore the application of these enzymes on non-native substrates. The site also offers users the opportunity to contribute additional information to its enzyme database, which allows for increased diversity and greater accessibility to the enzymatic reactions available to synthetic organic chemists.
In cases where a researcher finds an enzyme that is known to perform their reaction of interest but is not capable of converting their substrate, there are bioinformatic tools that can be used to identify similar enzymes that might act on their substrate. One such tool is the National Center for Biotechnology Information’s (NCBI) Basic Local Alignment Search Tool (BLAST). (76) Through this resource, users are able to identify proteins with high sequence similarity to their enzymes of interest via an advanced algorithm that pulls information from enzyme databases hosted by NCBI. The output of this tool is a list of proteins and their corresponding sequence alignment scores relative to the enzyme of interest. These related enzymes can then be produced and assayed to look for the modulated activity.
Phylogenetic trees are another useful bioinformatic tool that scientists can use to look for similar enzymes and define the relationship between enzymes in a family. (77) The generated diagrams offer information related to the evolutionary history of protein families and enable chemists to explore highly related enzymes that may offer improved or alternate reactivity from a the initial query enzyme. (78−80) Phylogenetic trees may be constructed and visualized readily using free programs such as Molecular Evolutionary Genetics Analysis (MEGA) and Ensembl. (81−83)
Sequence similarity networks (SSNs) are also highly valuable in the development of biocatalytic reactions, as they allow users to compare an extensive collection of protein sequences available in online databases. (84,85) Fundamentally, SSNs are a visualization tool that assists scientists in organizing large volumes of protein sequence data, which enables them to better determine sequence-function relationships and identify enzymes for a target transformation. (84) This process is valuable to synthetic organic chemists who wish to explore libraries of enzymes for developing reactions with wide substrate scopes in which enzyme promiscuity or differential reactivity/selectivity is crucial. (26,86−91) SSNs have become widely accessible with the development of the Enzyme Function Initiative - Enzyme Similarity Tool (EFI-EST) provided by the University of Illinois at Urbana–Champaign. (92) This tool allows chemists to easily input sequence data from proteins of interest and generate SSNs within minutes. Furthermore, these generated SSNs can be viewed and using Cytoscape, an open-source platform for processing complex network data. (93) To further assist chemists, Cytoscape also offers many tutorials and demonstrations for new users in processing SSNs. In addition to these tools, key examples of other publicly available enzyme database tools are summarized in Table 1.

4.2. Literature Resources

A growing number of literary resources are available to organic chemists interested in exploring biocatalysis. (72) In his recent perspective, Roger Sheldon, a pioneer in green chemistry and biocatalysis, offers a guide on how traditional organic chemists can access biocatalysts, what enzymes are and how they operate, and discusses how chemists can employ these biological tools for a variety of chemical transformations. (10) In a recent review from the laboratories of the renowned synthetic chemist, Erick Carreira, and innovator in biocatalysis, Nicholas Turner, the authors provide a framework for the implementation of biocatalysis in synthetic organic chemistry through the scope of retrosynthetic analysis. The concept of retrosynthetic analysis has shaped the logic of synthetic planning for organic chemists and has enabled access to countless synthetic targets, and including biocatalysis further expands its usefulness. (94) Furthermore, the Supporting Information documents that are included with biocatalysis or chemoenzymatic reports in the literature may give detailed descriptions of how enzymes were accessed and used in reactions. These are great starting points for becoming familiar with some of the technical aspects of biocatalysis.
In addition to the variety of published articles detailing the opportunity to incorporate biocatalytic methods in synthetic organic chemistry, an assortment of books provide more in-depth information and ideas to assist organic chemists in implementing biocatalysis in areas such as complex molecule synthesis. (95,96) In his recently published book, “Biocatalysis in Organic Synthesis: The Retrosynthesis Approach,” Turner further provides organic chemists a framework for implementing biocatalytically driven transformations based on the familiar technique of retrosynthesis. (97)

4.3. Industry Investment in Biocatalysis

Industrial science often requires optimization for large scale processes which puts them in a position to drive innovation and fine-tune methods, making the tools even more useful for the synthetic community. Biocatalysis is increasingly being adopted by companies, with new examples of large scale enzymatic/chemoenzymatic methods being published every year. (27) These syntheses are often on process scale and can significantly improve previous syntheses. For example, the COVID-19 antiviral drug Molnupiravir was synthesized by Merck with two out of the three total steps of the synthesis using enzymes with a 69% overall yield (Figure 4). (28) Similarly, Pfizer used commercially available enzymes to perform a multikilogram chemoenzymatic synthesis of a selective γ-secretase inhibitor with potential antitumor activity. (98) The continued investment of companies in biocatalysis will surely push forward the development of new technologies and processes that will broaden the applicability of biocatalysis in synthesis.

Figure 4

Figure 4. Chemoenzymatic synthesis of Molnupiravir demonstrated by Merck (right) compared to the previous small-molecule route (left).

5. How Can We Make Biocatalysis More Accessible?

Click to copy section linkSection link copied!

Although these resources and tools have started to bring biocatalysis into the realm of synthetic organic chemistry, significant work remains to make enzyme catalysis more accessible to the average chemist. Broadly, there needs to be an expansion of the educational curriculum, technical biocatalysis resources, data repository development, access to a wider selection of enzymes, and increased cross-disciplinary collaboration. As these changes occur, biocatalysis will become a more accessible tool for synthetic organic chemistry laboratories.

5.1. Undergraduate Education and Training

Perhaps one of the most important ways to impact the chemistry community is the early exposure of the next generation of chemists to topics through undergraduate education and laboratory training. In many cases, classroom and laboratory courses are students’ first exposure to the various subsets of chemistry and can be foundational for their engagement with the topics as they progress through their education and careers. There have been significant strides in synthetic organic chemistry in the past 40 years, yet many modern techniques are slow to be incorporated into classroom or laboratory courses. (99,100) This remains true for enzymatic catalysis as there is a distinct lack of biocatalysis in chemistry curricula.
There are many ways that biocatalysis could be introduced to students. One simple way to do this is by presenting biocatalytic methods in parallel with small molecule methods when discussing a given chemical reaction. For example, most introductory organic chemistry courses include a section on carbonyl reduction where students learn traditional racemic and enantioselective methods for accomplishing this transformation. (101) Dehydrogenase enzymes could be presented in parallel with the textbook small molecule reducing agents as a complementary method for carbonyl reduction. (102) This particular example also has the opportunity to bring in meaningful historical context for the use of natural enzymes. (5,103)
Organic chemistry laboratory courses are also a highly impactful way for students to engage in synthetic methods; however, few courses include biocatalysis experiments. Biocatalysis is relatively easy to include in laboratory courses. This hands-on exposure to enzymes would begin dismantling the “black box” in which biocatalysis is often placed. It is possible to design an experiment that requires only standard laboratory equipment, a commercially available enzyme, and the requisite substrate(s).
Laboratory units like this have been designed and implemented previously and successfully teach many different aspects of organic synthesis. (104,105) For example, Johnston and co-workers published a laboratory unit on a multistep chemoenzymatic synthesis off endogenous cannabinergic ligand 2-arachidonoylglycerol (Figure 5). (104) This synthesis introduced students to enzymatic catalysis and also reinforced other essential synthetic techniques.

Figure 5

Figure 5. Example of a chemoenzymatic synthesis used in an undergraduate chemistry laboratory course.

Every chemistry department should consider evolving their introductory courses to include biocatalysis as the curriculum is updated. The examples referenced herein would be an excellent starting point for instructors who wish to incorporate biocatalysis into their laboratory curriculum.

5.2. Tutorials

Clear, comprehensive, and accessible explanations of what it takes to perform biocatalysis have, to our knowledge, not yet been consolidated as a resource for the community. The books, articles, and Supporting Information documents listed earlier in this Perspective contain excellent basic descriptions of how to do biocatalytic transformations; however, they lack a complete picture of the whole process. This points to the need for a readily available and centralized compilation of technical biocatalysis resources analogous to Not Voodoo, a web site developed by Alison Frontier at the University of Rochester that is an excellent resource for conducting traditional organic chemistry reactions. The tagline for the web site is “Demystifying synthetic organic chemistry since 2004”─a goal it has undoubtedly accomplished with its many visitors over the years. (106) The web site is a compendium of information geared toward enhancing the experimental skills of synthetic organic chemists and includes sections like Tips and Tricks, Chromatography, Workup, Troubleshooting, and several more. This web site offers valuable information for the expert and novice synthetic alike.
A similar resource for biocatalysis would begin to demystify the technical aspects of enzymatic reactions that can be a barrier to its use in the laboratory. A biocatalysis-focused web site could compile practical tips, describe how to find enzymes, address common issues, and any other information useful for enabling the incorporation of enzymes in synthetic chemistry. Furthermore, this web site could include video demonstrations of the various steps of the biocatalysis process to make it even more accessible. An additional component that could be useful is a forum where community members can ask questions and have them answered by biocatalysis experts. This would further the accessibility of enzyme catalysis, while also encouraging discourse that would progress the field.

5.3. Enzyme Activity Data Repositories

As previously noted in this perspective, there are select examples of databases that catalog enzyme activity data. However, many of these are still in their infancy. For example, the data currently represented in RetroBioCat (https://retrobiocat.com/leaderboard) comes from 420 papers with a total of 1773 enzyme sequences. Considering the many institutions and groups publishing on enzymatic reactions, these are representatively small totals.
For these databases to become more broadly useful, they need a more robust foundation of data. This will be accomplished, in part, over time as data continue to be collected and submitted to the database; however, a more proactive approach should be adopted to rapidly make these tools useful to chemists. If all groups participating in biocatalysis or fundamental enzyme research upload experimental enzyme activity data to these repositories, it would improve the utility of these resources for the chemistry community. The cooperative efforts of the chemistry and biochemistry communities will enable these data collections to grow and improve.
In addition to bolstering the effectiveness of enzyme-specific databases, popular chemistry repositories (i.e., Scifinder and Reaxys) must be modified to best facilitate the indexing and presentation of biocatalytic reactions. Although the functionality to search by organisms/enzymes exists, many publications that are focused on biocatalysts are not indexed in the same manner as more conventional organic chemistry. For example, a search in Reaxys for a publication from our lab that describes biocatalytic α-deuteration of amino acids and methyl esters reveals that none of the many “Substances” or “Reactions” have been cataloged. (107) This is in stark contrast to a publication from the same journal in the same year on photocatalytic construction of α-functionalized amine products, (108) which has 66 “Substances” and 35 “Reactions” indexed. These papers contain a similar number of chemical structures and schemes in both the article and Supporting Information, and there are no clear differences that would lead to this difference in indexing.
These types of large databases are often the first step in a chemist’s journey to implement new chemistries in their syntheses. If there is a disparity in the classes of reactions represented, the underrepresented methods will naturally be incorporated much slower. Furthermore, robust data availability is critically important for developing predictive models that can aid in the enzyme catalyst choice. Therefore, we enthusiastically encourage chemical search and retrieval systems like Reaxys and SciFinder to evaluate their methods of indexing new publications and ensure that all types of chemistry are represented in their databases. Scientists conducting research in enzymatic catalysis should also make certain their publications are included in these databases by contacting them if the data are not accurately reflected.

5.4. Continued Interdisciplinary Collaboration and Acceptance

Biocatalysis is a highly interdisciplinary field that combines aspects of biology, engineering, and organic chemistry. Because of this, it often leads to collaboration among experts in separate fields. In particular, to progress the use of biocatalysis in chemical synthesis, continued collaboration between biocatalysis and synthetic organic chemistry groups will be critical. This will progress the field of chemoenzymatic synthesis while also destigmatizing the use of enzymes in synthesis.
Indeed, these types of collaborations have become more regular over the past few decades, and prominent members of the organic chemistry and biocatalysis communities have worked together to demonstrate high-impacting science. For example, in 2020 Todd Hyster, David MacMillan, and co-workers presented a chemoenzymatic method for dynamic kinetic resolution of β-substituted ketones, combining the utility of modern photoredox catalysis and biocatalysis into a useful synthetic platform (Figure 6A). (109) Francis Arnold, Brian Stoltz, and co-workers also joined forces in their enantioselective chemoenzymatic total synthesis of norditerpenoid alkaloid nigelladine A. (15) The synthesis relied on late-stage C–H oxidation, which was not possible with current chemical methods. This was solved by implementing an engineered P450 enzyme which completed the desired oxidation in a regio- and stereoselective manner (Figure 6B).

Figure 6

Figure 6. Examples of chemoenzymatic and enzymatic methods that result from collaborations between organic and biocatalysis research groups.

In addition to these strong collaborations, some well-established members of the organic chemistry community have taken large steps into biocatalysis and have subsequently made major advances in the field. One such chemist is John Hartwig, who started with biocatalysis collaborations and whose group conducts research on biocatalysis in their own laboratory, making major contributions to the active research area of artificial metalloenzymes. (110,111) An example of this work is displayed in a recent publication where they use a non-natural iridium porphyrin with cytochromes P450 to site-selectively functionalize (sp3) C–H bonds (Figure 6C). (112)
When we asked him about the place of enzymes in organic chemistry, Hartwig commented, “The catalysts produced by Nature possess many advantages, but they also possess many disadvantages and limitations, relative to the catalysts produced by chemists in the laboratory. High selectivity for the creation of complex chemical architectures and genetic encoding that allows evolution are two important attributes of Nature’s catalysts, but the limited range of chemical transformations is one liability. For years, I, and presumably others, have dreamed of interleaving chemical and enzymatic catalysts to create complex structures biosynthetically in an organism and of an ability to evolve organometallic catalysts in the lab or in a living organism. Researchers in the field of artificial metalloenzymes are progressing toward this mutual vision, and one can imagine a future in which synthetic chemists are equally comfortable with small-molecule catalysts and large-molecule enzymes. Remember, it wasn’t so long ago that synthetic organic chemists considered a palladium catalyst useful as a last resort, an iridium catalyst too expensive, and gold best for a children’s bed-time story.”
These collaborations and new ventures from seasoned organic chemists represent only a small fraction of the potential for biocatalytic and chemoenzymatic synthetic strategies. Many early career scientists have already developed robust research programs focused on biocatalysis and are paving the way for others to join the community.

6. Conclusion

Click to copy section linkSection link copied!

Biocatalysis is a highly beneficial tool for organic synthesis. Despite this, historical barriers have slowed the adoption of enzyme catalysts in the broader organic chemistry toolbox. In this Perspective, we shared how such obstacles can be addressed and how new developments are empowering every organic chemist to use enzymes.
The best way to encourage growth for both organic chemistry and biocatalysis is by collaborating with one another. We encourage every organic chemist to consider contacting colleagues who research biocatalysis to help solve challenging chemical problems. Likewise, we urge those studying enzymes to join synthetic chemists in envisioning new ways to use biocatalysts in complex molecule synthesis and chemoenzymatic methods. Through strategic collaboration, existing barriers will be dismantled and biocatalysts will quickly become a tool not relegated to niche communities, but rather, the door will be open to all organic chemists.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Evan O. Romero - Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United StatesOrcidhttps://orcid.org/0000-0001-8553-1238
    • Anthony T. Saucedo - Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
    • José R. Hernández-Meléndez - Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
    • Di Yang - Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
    • Suman Chakrabarty - Life Sciences Institute & Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United StatesOrcidhttps://orcid.org/0000-0002-6611-3839
  • Author Contributions

    CRediT: Evan O. Romero writing-original draft, writing-review & editing; Anthony T. Saucedo writing-original draft, writing-review & editing; José R. Hernández-Meléndez writing-original draft, writing-review & editing; Di Yang writing-original draft, writing-review & editing; Suman Chakrabarty writing-original draft, writing-review & editing; Alison R. H. Narayan writing-original draft, writing-review & editing.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors acknowledge support from the University of Michigan Life Sciences Institute, University of Michigan Department of Chemistry, the National Institutes of Health Grant R35 GM124880, and NSF 2221346. E.O.R. and A.T.S acknowledge support from the NSF graduate research fellowship (DGE 1841052). J.R.H.M. acknowledges support from the NIH Chemistry-Biology Interface training program (5T32GM132046-03). The authors graciously thank Professor Brian Stoltz and Professor John Hartwig for providing statements included in this piece.

References

Click to copy section linkSection link copied!

This article references 112 other publications.

  1. 1
    Atanasov, A. G.; Zotchev, S. B.; Dirsch, V. M.; Orhan, I. E.; Banach, M.; Rollinger, J. M.; Barreca, D.; Weckwerth, W.; Bauer, R.; Bayer, E. A.; Majeed, M.; Bishayee, A.; Bochkov, V.; Bonn, G. K.; Braidy, N.; Bucar, F.; Cifuentes, A.; D’Onofrio, G.; Bodkin, M.; Diederich, M.; Dinkova-Kostova, A. T.; Efferth, T.; El Bairi, K.; Arkells, N.; Fan, T.-P.; Fiebich, B. L.; Freissmuth, M.; Georgiev, M. I.; Gibbons, S.; Godfrey, K. M.; Gruber, C. W.; Heer, J.; Huber, L. A.; Ibanez, E.; Kijjoa, A.; Kiss, A. K.; Lu, A.; Macias, F. A.; Miller, M. J. S.; Mocan, A.; Müller, R.; Nicoletti, F.; Perry, G.; Pittalà, V.; Rastrelli, L.; Ristow, M.; Russo, G. L.; Silva, A. S.; Schuster, D.; Sheridan, H.; Skalicka-Woźniak, K.; Skaltsounis, L.; Sobarzo-Sánchez, E.; Bredt, D. S.; Stuppner, H.; Sureda, A.; Tzvetkov, N. T.; Vacca, R. A.; Aggarwal, B. B.; Battino, M.; Giampieri, F.; Wink, M.; Wolfender, J.-L.; Xiao, J.; Yeung, A. W. K.; Lizard, G.; Popp, M. A.; Heinrich, M.; Berindan-Neagoe, I.; Stadler, M.; Daglia, M.; Verpoorte, R.; Supuran, C. T. Natural products in drug discovery: advances and opportunities. Nat. Rev. Drug Discovery 2021, 20, 200216,  DOI: 10.1038/s41573-020-00114-z
  2. 2
    Dandapani, S.; Marcaurelle, L. A. Grand Challenge Commentary: Accessing new chemical space for ’undruggable’ targets. Nat. Chem. Biol. 2010, 6, 861863,  DOI: 10.1038/nchembio.479
  3. 3
    Rotella, D. P. The Critical Role of Organic Chemistry in Drug Discovery. ACS Chem. Neurosci. 2016, 7, 13151316,  DOI: 10.1021/acschemneuro.6b00280
  4. 4
    Grygorenko, O. O.; Volochnyuk, D. M.; Ryabukhin, S. V.; Judd, D. B. The Symbiotic Relationship Between Drug Discovery and Organic Chemistry. Chem. Eur. J. 2020, 26, 11961237,  DOI: 10.1002/chem.201903232
  5. 5
    Pyser, J. B.; Chakrabarty, S.; Romero, E. O.; Narayan, A. R. H. State-of-the-Art Biocatalysis. ACS Cent. Sci. 2021, 7, 11051116,  DOI: 10.1021/acscentsci.1c00273
  6. 6
    Chakrabarty, S.; Romero, E. O.; Pyser, J. B.; Yazarians, J. A.; Narayan, A. R. H. Chemoenzymatic Total Synthesis of Natural Products. Acc. Chem. Res. 2021, 54, 13741384,  DOI: 10.1021/acs.accounts.0c00810
  7. 7
    Zetzsche, L. E.; Yazarians, J. A.; Chakrabarty, S.; Hinze, M. E.; Murray, L. A. M.; Lukowski, A. L.; Joyce, L. A.; Narayan, A. R. H. Biocatalytic oxidative cross-coupling reactions for biaryl bond formation. Nature 2022, 603, 7985,  DOI: 10.1038/s41586-021-04365-7
  8. 8
    Chakrabarty, S.; Wang, Y.; Perkins, J. C.; Narayan, A. R. H. Scalable biocatalytic C–H oxyfunctionalization reactions. Chem. Soc. Rev. 2020, 49, 81378155,  DOI: 10.1039/D0CS00440E
  9. 9
    Clouthier, C. M.; Pelletier, J. N. Expanding the organic toolbox: a guide to integrating biocatalysis in synthesis. Chem. Soc. Rev. 2012, 41, 15851605,  DOI: 10.1039/c2cs15286j
  10. 10
    Sheldon, R. A.; Brady, D.; Bode, M. L. The Hitchhiker’s guide to biocatalysis: recent advances in the use of enzymes in organic synthesis. Chem. Sci. 2020, 11, 25872605,  DOI: 10.1039/C9SC05746C
  11. 11
    Abdelraheem, E. M. M.; Busch, H.; Hanefeld, U.; Tonin, F. Biocatalysis explained: from pharmaceutical to bulk chemical production. React. Chem. Eng. 2019, 4, 18781894,  DOI: 10.1039/C9RE00301K
  12. 12
    Sheldon, R. A.; Brady, D. Broadening the Scope of Biocatalysis in Sustainable Organic Synthesis. ChemSusChem 2019, 12, 28592881,  DOI: 10.1002/cssc.201900351
  13. 13
    Hughes, G.; Lewis, J. C. Introduction: Biocatalysis in Industry. Chem. Rev. 2018, 118, 13,  DOI: 10.1021/acs.chemrev.7b00741
  14. 14
    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.; Halsey, H. M.; Hughes, G. J.; Jo, J.; Joyce, L. A.; Kolev, J. N.; Liang, J.; Maloney, K. M.; Mann, B. F.; Marshall, N. M.; McLaughlin, M.; Moore, J. C.; Murphy, G. S.; Nawrat, C. C.; Nazor, J.; Novick, S.; Patel, N. R.; Rodriguez-Granillo, A.; Robaire, S. A.; Sherer, E. C.; Truppo, M. D.; Whittaker, A. M.; Verma, D.; Xiao, L.; Xu, Y.; Yang, H. Design of an in vitro biocatalytic cascade for the manufacture of islatravir. Science 2019, 366, 12551259,  DOI: 10.1126/science.aay8484
  15. 15
    Loskot, S. A.; Romney, D. K.; Arnold, F. H.; Stoltz, B. M. Enantioselective Total Synthesis of Nigelladine A via Late-Stage C–H Oxidation Enabled by an Engineered P450 Enzyme. J. Am. Chem. Soc. 2017, 139, 1019610199,  DOI: 10.1021/jacs.7b05196
  16. 16
    Chen, K.; Huang, X.; Kan, S. B. J.; Zhang, R. K.; Arnold, F. H. Enzymatic construction of highly strained carbocycles. Science 2018, 360, 7175,  DOI: 10.1126/science.aar4239
  17. 17
    Zhang, X.; King-Smith, E.; Dong, L.-B.; Yang, L.-C.; Rudolf, J. D.; Shen, B.; Renata, H. Divergent synthesis of complex diterpenes through a hybrid oxidative approach. Science 2020, 369, 799806,  DOI: 10.1126/science.abb8271
  18. 18
    Nakamura, H.; Schultz, E. E.; Balskus, E. P. A new strategy for aromatic ring alkylation in cylindrocyclophane biosynthesis. Nat. Chem. Biol. 2017, 13, 916921,  DOI: 10.1038/nchembio.2421
  19. 19
    Schultz, E. E.; Braffman, N. R.; Luescher, M. U.; Hager, H. H.; Balskus, E. P. Biocatalytic Friedel–Crafts Alkylation Using a Promiscuous Biosynthetic Enzyme. Angew. Chem., Int. Ed. 2019, 58, 31513155,  DOI: 10.1002/anie.201814016
  20. 20
    Lau, W.; Sattely, E. S. Six enzymes from mayapple that complete the biosynthetic pathway to the etoposide aglycone. Science 2015, 349, 12241228,  DOI: 10.1126/science.aac7202
  21. 21
    Lowell, A. N.; DeMars, M. D.; Slocum, S. T.; Yu, F.; Anand, K.; Chemler, J. A.; Korakavi, N.; Priessnitz, J. K.; Park, S. R.; Koch, A. A.; Schultz, P. J.; Sherman, D. H. Chemoenzymatic Total Synthesis and Structural Diversification of Tylactone-Based Macrolide Antibiotics through Late-Stage Polyketide Assembly, Tailoring, and C─H Functionalization. J. Am. Chem. Soc. 2017, 139, 79137920,  DOI: 10.1021/jacs.7b02875
  22. 22
    Lukowski, A. L.; Denomme, N.; Hinze, M. E.; Hall, S.; Isom, L. L.; Narayan, A. R. H. Biocatalytic Detoxification of Paralytic Shellfish Toxins. ACS Chem. Biol. 2019, 14, 941948,  DOI: 10.1021/acschembio.9b00123
  23. 23
    Wang, J.; Zhang, Y.; Liu, H.; Shang, Y.; Zhou, L.; Wei, P.; Yin, W.-B.; Deng, Z.; Qu, X.; Zhou, Q. A biocatalytic hydroxylation-enabled unified approach to C19-hydroxylated steroids. Nat. Commun. 2019, 10, 3378,  DOI: 10.1038/s41467-019-11344-0
  24. 24
    Pyser, J. B.; Baker Dockrey, S. A.; Benítez, A. R.; Joyce, L. A.; Wiscons, R. A.; Smith, J. L.; Narayan, A. R. H. Stereodivergent, Chemoenzymatic Synthesis of Azaphilone Natural Products. J. Am. Chem. Soc. 2019, 141, 1855118559,  DOI: 10.1021/jacs.9b09385
  25. 25
    Zwick, C. R.; Renata, H. Remote C–H Hydroxylation by an α-Ketoglutarate-Dependent Dioxygenase Enables Efficient Chemoenzymatic Synthesis of Manzacidin C and Proline Analogs. J. Am. Chem. Soc. 2018, 140, 11651169,  DOI: 10.1021/jacs.7b12918
  26. 26
    Lukowski, A. L.; Liu, J.; Bridwell-Rabb, J.; Narayan, A. R. H. Structural basis for divergent C–H hydroxylation selectivity in two Rieske oxygenases. Nat. Commun. 2020, 11, 2991,  DOI: 10.1038/s41467-020-16729-0
  27. 27
    Wu, S.; Snajdrova, R.; Moore, J. C.; Baldenius, K.; Bornscheuer, U. T. Biocatalysis: Enzymatic Synthesis for Industrial Applications. Angew. Chem., Int. Ed. 2021, 60, 88119,  DOI: 10.1002/anie.202006648
  28. 28
    McIntosh, J. A.; Benkovics, T.; Silverman, S. M.; Huffman, M. A.; Kong, J.; Maligres, P. E.; Itoh, T.; Yang, H.; Verma, D.; Pan, W.; Ho, H.-I.; Vroom, J.; Knight, A. M.; Hurtak, J. A.; Klapars, A.; Fryszkowska, A.; Morris, W. J.; Strotman, N. A.; Murphy, G. S.; Maloney, K. M.; Fier, P. S. Engineered Ribosyl-1-Kinase Enables Concise Synthesis of Molnupiravir, an Antiviral for COVID-19. ACS Cent. Sci. 2021, 7, 19801985,  DOI: 10.1021/acscentsci.1c00608
  29. 29
    Bornscheuer, U. T.; Buchholz, K. Highlights in Biocatalysis – Historical Landmarks and Current Trends. Eng. Life Sci. 2005, 5, 309323,  DOI: 10.1002/elsc.200520089
  30. 30
    Buchner, E. Alkoholische Gährung ohne Hefezellen. Berichte der deutschen chemischen Gesellschaft 1897, 30, 117124,  DOI: 10.1002/cber.18970300121
  31. 31
    Heckmann, C. M.; Paradisi, F. Looking Back: A Short History of the Discovery of Enzymes and How They Became Powerful Chemical Tools. ChemCatChem. 2020, 12, 60826102,  DOI: 10.1002/cctc.202001107
  32. 32
    Whitesides, G. M. Applications of Cell-Free Enzymes in Organic Synthesis. In Ciba Foundation Symposium 111 - Enzymes in Organic Synthesis; Pitman: London, 1985; pp 7696.
  33. 33
    Olivieri, R.; Fascetti, E.; Angelini, L.; Degen, L. Microbial transformation of racemic hydantoins to d-amino acids. Biotechnol. Bioeng. 1981, 23, 21732183,  DOI: 10.1002/bit.260231002
  34. 34
    Liu, Y.; Zhu, L.; Qi, W.; Yu, B. Biocatalytic production of D-p-hydroxyphenylglycine by optimizing protein expression and cell wall engineering in Escherichia coli. Appl. Microbiol. Biotechnol. 2019, 103, 88398851,  DOI: 10.1007/s00253-019-10155-z
  35. 35
    Buchholz, K. A breakthrough in enzyme technology to fight penicillin resistance─industrial application of penicillin amidase. Appl. Microbiol. Biotechnol. 2016, 100, 38253839,  DOI: 10.1007/s00253-016-7399-6
  36. 36
    Wicks, C.; Hudlicky, T.; Rinner, U. Morphine alkaloids: History, biology, and synthesis. In The Alkaloids: Chemistry and Biology; Knölker, H.-J., Ed.; Academic Press: 2021; Vol. 86, Ch. 2, pp 145342.
  37. 37
    Gulland, J. M.; Robinson, R. Constitution of codeine and thebaine. Mem. Proc. Manchester Lit. Philos. Soc. 1925, 69, 7986
  38. 38
    Armstrong, E. F. Enzymes: A Discovery and its Consequences. Nature 1933, 131, 535537,  DOI: 10.1038/131535a0
  39. 39
    Mohan, R. S.; Mejia, M. P. Environmentally Friendly Organic Chemistry Laboratory Experiments for the Undergraduate Curriculum: A Literature Survey and Assessment. J. Chem. Educ. 2020, 97, 943959,  DOI: 10.1021/acs.jchemed.9b00753
  40. 40
    Heather, J. M.; Chain, B. The sequence of sequencers: The history of sequencing DNA. Genomics 2016, 107, 18,  DOI: 10.1016/j.ygeno.2015.11.003
  41. 41
    Baxevanis, A. D. Using Genomic Databases for Sequence-Based Biological Discovery. Mol. Med. 2003, 9, 185192,  DOI: 10.1007/BF03402130
  42. 42
    GenBank and WGS Statistics. https://www.ncbi.nlm.nih.gov/genbank/statistics/ (accessed 2023-02–01).
  43. 43
    The UniProt Consortium UniProt: the universal protein knowledgebase. Nucleic Acids Res. 2017, 45, D158D169,  DOI: 10.1093/nar/gkw1099
  44. 44
    LeProust, E. M.; Peck, B. J.; Spirin, K.; McCuen, H. B.; Moore, B.; Namsaraev, E.; Caruthers, M. H. Synthesis of high-quality libraries of long (150mer) oligonucleotides by a novel depurination controlled process. Nucleic Acids Res. 2010, 38, 25222540,  DOI: 10.1093/nar/gkq163
  45. 45
    Gibson, D. G.; Young, L.; Chuang, R.-Y.; Venter, J. C.; Hutchison, C. A.; Smith, H. O. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 2009, 6, 343345,  DOI: 10.1038/nmeth.1318
  46. 46
    Loftie-Eaton, W.; Heinisch, T.; Soskine, M.; Champion, E.; Godron, X.; Ybert, T. Novel Variants of Endonuclease V and Uses Thereof. WO2022/090057, 2022.
  47. 47
    Moustafa, K.; Makhzoum, A.; Trémouillaux-Guiller, J. Molecular farming on rescue of pharma industry for next generations. Crit. Rev. Biotechnol. 2016, 36, 840850,  DOI: 10.3109/07388551.2015.1049934
  48. 48
    Swartz, J. R. Advances in Escherichia coli production of therapeutic proteins. Curr. Opin. Biotechnol. 2001, 12, 195201,  DOI: 10.1016/S0958-1669(00)00199-3
  49. 49
    Karbalaei, M.; Rezaee, S. A.; Farsiani, H. Pichia pastoris: A highly successful expression system for optimal synthesis of heterologous proteins. J. Cell. Physiol. 2020, 235, 58675881,  DOI: 10.1002/jcp.29583
  50. 50
    Hunter, M.; Yuan, P.; Vavilala, D.; Fox, M. Optimization of Protein Expression in Mammalian Cells. Curr. Protoc. Protein Sci. 2019, 95, e77  DOI: 10.1002/cpps.77
  51. 51
    Fox, B. G.; Blommel, P. G. Autoinduction of Protein Expression. Curr. Protoc. Protein Sci. 2009, 56, 5.23.15.23.18,  DOI: 10.1002/0471140864.ps0523s56
  52. 52
    Silverman, A. D.; Karim, A. S.; Jewett, M. C. Cell-free gene expression: an expanded repertoire of applications. Nat. Rev. Genet. 2020, 21, 151170,  DOI: 10.1038/s41576-019-0186-3
  53. 53
    de Carvalho, C. C. C. R. Whole cell biocatalysts: essential workers from Nature to the industry. Microb. Biotechnol. 2017, 10, 250263,  DOI: 10.1111/1751-7915.12363
  54. 54
    Alissandratos, A. In vitro multi-enzymatic cascades using recombinant lysates of E. coli: an emerging biocatalysis platform. Biophys. Rev. 2020, 12, 175182,  DOI: 10.1007/s12551-020-00618-3
  55. 55
    Gräslund, S.; Nordlund, P.; Weigelt, J.; Hallberg, B. M.; Bray, J.; Gileadi, O.; Knapp, S.; Oppermann, U.; Arrowsmith, C.; Hui, R.; Ming, J.; dhe-Paganon, S.; Park, H.-w.; Savchenko, A.; Yee, A.; Edwards, A.; Vincentelli, R.; Cambillau, C.; Kim, R.; Kim, S.-H.; Rao, Z.; Shi, Y.; Terwilliger, T. C.; Kim, C.-Y.; Hung, L.-W.; Waldo, G. S.; Peleg, Y.; Albeck, S.; Unger, T.; Dym, O.; Prilusky, J.; Sussman, J. L.; Stevens, R. C.; Lesley, S. A.; Wilson, I. A.; Joachimiak, A.; Collart, F.; Dementieva, I.; Donnelly, M. I.; Eschenfeldt, W. H.; Kim, Y.; Stols, L.; Wu, R.; Zhou, M.; Burley, S. K.; Emtage, J. S.; Sauder, J. M.; Thompson, D.; Bain, K.; Luz, J.; Gheyi, T.; Zhang, F.; Atwell, S.; Almo, S. C.; Bonanno, J. B.; Fiser, A.; Swaminathan, S.; Studier, F. W.; Chance, M. R.; Sali, A.; Acton, T. B.; Xiao, R.; Zhao, L.; Ma, L. C.; Hunt, J. F.; Tong, L.; Cunningham, K.; Inouye, M.; Anderson, S.; Janjua, H.; Shastry, R.; Ho, C. K.; Wang, D.; Wang, H.; Jiang, M.; Montelione, G. T.; Stuart, D. I.; Owens, R. J.; Daenke, S.; Schütz, A.; Heinemann, U.; Yokoyama, S.; Büssow, K.; Gunsalus, K. C.; Structural Genomics, C.; Architecture et Fonction des Macromolécules, B.; Berkeley Structural Genomics, C.; China Structural Genomics, C.; Integrated Center for, S.; Function, I.; Israel Structural Proteomics, C.; Joint Center for Structural, G.; Midwest Center for Structural, G.; New York Structural Genomi, X. R. C. f. S. G.; Northeast Structural Genomics, C.; Oxford Protein Production, F.; Protein Sample Production Facility, M. D. C. f. M. M.; Initiative, R. S. G. P.; Complexes, S. Protein production and purification. Nat. Methods 2008, 5, 135146,  DOI: 10.1038/nmeth.f.202
  56. 56
    Hughes, R. A.; Ellington, A. D. Synthetic DNA Synthesis and Assembly: Putting the Synthetic in Synthetic Biology. Cold Spring Harbor Perspect. Biol. 2017, 9, a023812,  DOI: 10.1101/cshperspect.a023812
  57. 57
    Baker Dockrey, S. A.; Doyon, T. J.; Perkins, J. C.; Narayan, A. R. H. Whole-cell biocatalysis platform for gram-scale oxidative dearomatization of phenols. Chem. Biol. Drug Des. 2019, 93, 12071213,  DOI: 10.1111/cbdd.13443
  58. 58
    Bai, Y.; Yang, X.; Yu, H.; Chen, X. Substrate and Process Engineering for Biocatalytic Synthesis and Facile Purification of Human Milk Oligosaccharides. ChemSusChem 2022, 15, e202102539  DOI: 10.1002/cssc.202102539
  59. 59
    Börner, T.; Grey, C.; Adlercreutz, P. Generic HPLC platform for automated enzyme reaction monitoring: Advancing the assay toolbox for transaminases and other PLP-dependent enzymes. Biotechnol. J. 2016, 11, 10251036,  DOI: 10.1002/biot.201500587
  60. 60
    Claaßen, C.; Mack, K.; Rother, D. Benchtop NMR for Online Reaction Monitoring of the Biocatalytic Synthesis of Aromatic Amino Alcohols. ChemCatChem. 2020, 12, 11901199,  DOI: 10.1002/cctc.201901910
  61. 61
    Bommarius, A. S. Biocatalysis: A Status Report. Annu. Rev. Chem. Biomol. Eng. 2015, 6, 319345,  DOI: 10.1146/annurev-chembioeng-061114-123415
  62. 62
    Reetz, M. T. What are the Limitations of Enzymes in Synthetic Organic Chemistry?. Chem. Rec. 2016, 16, 24492459,  DOI: 10.1002/tcr.201600040
  63. 63
    Stepankova, V.; Bidmanova, S.; Koudelakova, T.; Prokop, Z.; Chaloupkova, R.; Damborsky, J. Strategies for Stabilization of Enzymes in Organic Solvents. ACS Catal. 2013, 3, 28232836,  DOI: 10.1021/cs400684x
  64. 64
    Guzik, U.; Hupert-Kocurek, K.; Wojcieszyńska, D. Immobilization as a Strategy for Improving Enzyme Properties-Application to Oxidoreductases. Molecules 2014, 19, 89959018,  DOI: 10.3390/molecules19078995
  65. 65
    De Santis, P.; Meyer, L.-E.; Kara, S. The rise of continuous flow biocatalysis – fundamentals, very recent developments and future perspectives. React. Chem. Eng. 2020, 5, 21552184,  DOI: 10.1039/D0RE00335B
  66. 66
    France, S. P.; Lewis, R. D.; Martinez, C. A. The Evolving Nature of Biocatalysis in Pharmaceutical Research and Development. JACS Au 2023, 3, 715735,  DOI: 10.1021/jacsau.2c00712
  67. 67
    Zhang, Y.; Xia, B.; Li, Y.; Lin, X.; Wu, Q. Substrate Engineering in Lipase-Catalyzed Selective Polymerization of d-/l-Aspartates and Diols to Prepare Helical Chiral Polyester. Biomacromolecules 2021, 22, 918926,  DOI: 10.1021/acs.biomac.0c01605
  68. 68
    Turner, N. J. Directed evolution drives the next generation of biocatalysts. Nat. Chem. Biol. 2009, 5, 567573,  DOI: 10.1038/nchembio.203
  69. 69
    Cobb, R. E.; Chao, R.; Zhao, H. Directed evolution: Past, present, and future. AIChE J. 2013, 59, 14321440,  DOI: 10.1002/aic.13995
  70. 70
    Steiner, K.; Schwab, H. Recent advances in rational approaches for enzyme engineering. Comput. Struct. Biotechnol. J. 2012, 2, e201209010  DOI: 10.5936/csbj.201209010
  71. 71
    Fernandes, P. Miniaturization in Biocatalysis. Int. J. Mol. Sci. 2010, 11, 858879,  DOI: 10.3390/ijms11030858
  72. 72
    Bell, E. L.; Finnigan, W.; France, S. P.; Green, A. P.; Hayes, M. A.; Hepworth, L. J.; Lovelock, S. L.; Niikura, H.; Osuna, S.; Romero, E.; Ryan, K. S.; Turner, N. J.; Flitsch, S. L. Biocatalysis. Nat. Rev. Methods Primers 2021, 1, 46,  DOI: 10.1038/s43586-021-00044-z
  73. 73
    Duetz, W. A. Microtiter plates as mini-bioreactors: miniaturization of fermentation methods. Trends Microbiol. 2007, 15, 469475,  DOI: 10.1016/j.tim.2007.09.004
  74. 74
    Diefenbach, X. W.; Farasat, I.; Guetschow, E. D.; Welch, C. J.; Kennedy, R. T.; Sun, S.; Moore, J. C. Enabling Biocatalysis by High-Throughput Protein Engineering Using Droplet Microfluidics Coupled to Mass Spectrometry. ACS Omega 2018, 3, 14981508,  DOI: 10.1021/acsomega.7b01973
  75. 75
    Finnigan, W.; Hepworth, L. J.; Flitsch, S. L.; Turner, N. J. RetroBioCat as a computer-aided synthesis planning tool for biocatalytic reactions and cascades. Nat. Catal. 2021, 4, 98104,  DOI: 10.1038/s41929-020-00556-z
  76. 76
    Altschul, S. F.; Gish, W.; Miller, W.; Myers, E. W.; Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403410,  DOI: 10.1016/S0022-2836(05)80360-2
  77. 77
    Cai, X.-H.; Jaroszewski, L.; Wooley, J.; Godzik, A. Internal organization of large protein families: Relationship between the sequence, structure, and function-based clustering. Proteins: Struct., Funct., Bioinf. 2011, 79, 23892402,  DOI: 10.1002/prot.23049
  78. 78
    Sirota, F. L.; Maurer-Stroh, S.; Li, Z.; Eisenhaber, F.; Eisenhaber, B. Functional Classification of Super-Large Families of Enzymes Based on Substrate Binding Pocket Residues for Biocatalysis and Enzyme Engineering Applications. Front. Bioeng. Biotechnol. 2021, 9, 701120,  DOI: 10.3389/fbioe.2021.701120
  79. 79
    Wilding, M.; Peat, T. S.; Kalyaanamoorthy, S.; Newman, J.; Scott, C.; Jermiin, L. S. Reverse engineering: transaminase biocatalyst development using ancestral sequence reconstruction. Green Chem. 2017, 19, 53755380,  DOI: 10.1039/C7GC02343J
  80. 80
    Gumulya, Y.; Baek, J.-M.; Wun, S.-J.; Thomson, R. E. S.; Harris, K. L.; Hunter, D. J. B.; Behrendorff, J. B. Y. H.; Kulig, J.; Zheng, S.; Wu, X.; Wu, B.; Stok, J. E.; De Voss, J. J.; Schenk, G.; Jurva, U.; Andersson, S.; Isin, E. M.; Bodén, M.; Guddat, L.; Gillam, E. M. J. Engineering highly functional thermostable proteins using ancestral sequence reconstruction. Nat. Catal. 2018, 1, 878888,  DOI: 10.1038/s41929-018-0159-5
  81. 81
    Tamura, K.; Peterson, D.; Peterson, N.; Stecher, G.; Nei, M.; Kumar, S. MEGA5: Molecular Evolutionary Genetics Analysis Using Maximum Likelihood, Evolutionary Distance, and Maximum Parsimony Methods. Mol. Biol. Evol. 2011, 28, 27312739,  DOI: 10.1093/molbev/msr121
  82. 82
    Hall, B. G. Building phylogenetic trees from molecular data with MEGA. Mol. Biol. Evol. 2013, 30, 12291235,  DOI: 10.1093/molbev/mst012
  83. 83
    Yates, A. D.; Achuthan, P.; Akanni, W.; Allen, J.; Allen, J.; Alvarez-Jarreta, J.; Amode, M. R.; Armean, I. M.; Azov, A. G.; Bennett, R.; Bhai, J.; Billis, K.; Boddu, S.; Marugán, J. C.; Cummins, C.; Davidson, C.; Dodiya, K.; Fatima, R.; Gall, A.; Giron, C. G.; Gil, L.; Grego, T.; Haggerty, L.; Haskell, E.; Hourlier, T.; Izuogu, O. G.; Janacek, S. H.; Juettemann, T.; Kay, M.; Lavidas, I.; Le, T.; Lemos, D.; Martinez, J. G.; Maurel, T.; McDowall, M.; McMahon, A.; Mohanan, S.; Moore, B.; Nuhn, M.; Oheh, D. N.; Parker, A.; Parton, A.; Patricio, M.; Sakthivel, M. P.; Abdul Salam, A. I.; Schmitt, B. M.; Schuilenburg, H.; Sheppard, D.; Sycheva, M.; Szuba, M.; Taylor, K.; Thormann, A.; Threadgold, G.; Vullo, A.; Walts, B.; Winterbottom, A.; Zadissa, A.; Chakiachvili, M.; Flint, B.; Frankish, A.; Hunt, S. E.; IIsley, G.; Kostadima, M.; Langridge, N.; Loveland, J. E.; Martin, F. J.; Morales, J.; Mudge, J. M.; Muffato, M.; Perry, E.; Ruffier, M.; Trevanion, S. J.; Cunningham, F.; Howe, K. L.; Zerbino, D. R.; Flicek, P. Ensembl 2020. Nucleic Acids Res. 2020, 48, D682D688,  DOI: 10.1093/nar/gkz966
  84. 84
    Atkinson, H. J.; Morris, J. H.; Ferrin, T. E.; Babbitt, P. C. Using Sequence Similarity Networks for Visualization of Relationships Across Diverse Protein Superfamilies. PloS One 2009, 4, e4345  DOI: 10.1371/journal.pone.0004345
  85. 85
    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, 41694182,  DOI: 10.1021/acs.biochem.9b00735
  86. 86
    Doyon, T. J.; Perkins, J. C.; Baker Dockrey, S. A.; Romero, E. O.; Skinner, K. C.; Zimmerman, P. M.; Narayan, A. R. H. Chemoenzymatic o-Quinone Methide Formation. J. Am. Chem. Soc. 2019, 141, 2026920277,  DOI: 10.1021/jacs.9b10474
  87. 87
    Rodriguez Benitez, A.; Narayan, A. R. H. Frontiers in Biocatalysis: Profiling Function across Sequence Space. ACS Cent. Sci. 2019, 5, 17471749,  DOI: 10.1021/acscentsci.9b01112
  88. 88
    Fisher, B. F.; Snodgrass, H. M.; Jones, K. A.; Andorfer, M. C.; Lewis, J. C. Site-Selective C–H Halogenation Using Flavin-Dependent Halogenases Identified via Family-Wide Activity Profiling. ACS Cent. Sci. 2019, 5, 18441856,  DOI: 10.1021/acscentsci.9b00835
  89. 89
    Schülke, K. H.; Ospina, F.; Hörnschemeyer, K.; Gergel, S.; Hammer, S. C. Substrate Profiling of Anion Methyltransferases for Promiscuous Synthesis of S-Adenosylmethionine Analogs from Haloalkanes. ChemBioChem. 2022, 23, e202100632  DOI: 10.1002/cbic.202100632
  90. 90
    Lachowicz, J. C.; Gizzi, A. S.; Almo, S. C.; Grove, T. L. Structural Insight into the Substrate Scope of Viperin and Viperin-like Enzymes from Three Domains of Life. Biochemistry 2021, 60, 21162129,  DOI: 10.1021/acs.biochem.0c00958
  91. 91
    Tararina, M. A.; Allen, K. N. Bioinformatic Analysis of the Flavin-Dependent Amine Oxidase Superfamily: Adaptations for Substrate Specificity and Catalytic Diversity. J. Mol. Biol. 2020, 432, 32693288,  DOI: 10.1016/j.jmb.2020.03.007
  92. 92
    Gerlt, J. A.; Bouvier, J. T.; Davidson, D. B.; Imker, H. J.; Sadkhin, B.; Slater, D. R.; Whalen, K. L. Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST): A web tool for generating protein sequence similarity networks. Biochim. Biophys. Acta, Proteins Proteomics 2015, 1854, 10191037,  DOI: 10.1016/j.bbapap.2015.04.015
  93. 93
    Kohl, M.; Wiese, S.; Warscheid, B. Cytoscape: Software for Visualization and Analysis of Biological Networks. In Data Mining in Proteomics: From Standards to Applications; Hamacher, M., Eisenacher, M., Stephan, C., Eds.; Methods in Molecular Biology; Humana Press: Totowa, NJ, 2011; pp 291303.
  94. 94
    E. J. Corey, X.-M. C. The Logic of Chemical Synthesis; Wiley, :New York, 1989.
  95. 95
    Bommarius, A. S.; Riebel Bommarius, B. R. Biocatalysis; Wiley-VCH Verlag: Weinheim, Germany, 2004; p 634.
  96. 96
    Whittall, J. Applied biocatalysis; John Wiley & Sons: Nashville, TN, 2020; p 560.
  97. 97
    Turner, N. J.; Humphreys, L. Biocatalysis in Organic Synthesis: The Retrosynthesis Approach; Royal Society of Chemistry: 2018.
  98. 98
    Burns, M.; Martinez, C. A.; Vanderplas, B.; Wisdom, R.; Yu, S.; Singer, R. A. A Chemoenzymatic Route to Chiral Intermediates Used in the Multikilogram Synthesis of a Gamma Secretase Inhibitor. Org. Process Res. Dev. 2017, 21, 871877,  DOI: 10.1021/acs.oprd.7b00096
  99. 99
    Raker, J. R.; Holme, T. A. A Historical Analysis of the Curriculum of Organic Chemistry Using ACS Exams as Artifacts. J. Chem. Educ. 2013, 90, 14371442,  DOI: 10.1021/ed400327b
  100. 100
    Cooper, M. M.; Stowe, R. L.; Crandell, O. M.; Klymkowsky, M. W. Organic Chemistry, Life, the Universe and Everything (OCLUE): A Transformed Organic Chemistry Curriculum. J. Chem. Educ. 2019, 96, 18581872,  DOI: 10.1021/acs.jchemed.9b00401
  101. 101
    Raker, J.; Holme, T.; Murphy, K. The ACS Exams Institute Undergraduate Chemistry Anchoring Concepts Content Map II: Organic Chemistry. J. Chem. Educ. 2013, 90, 14431445,  DOI: 10.1021/ed400175w
  102. 102
    Brummund, J.; Sonke, T.; Müller, M. Process Development for Biocatalytic Oxidations Applying Alcohol Dehydrogenases. Org. Process Res. Dev. 2015, 19, 15901595,  DOI: 10.1021/op500307e
  103. 103
    Wong, C.-H.; Whitesides, G. M. Enzyme-catalyzed organic synthesis: NAD(P)H cofactor regeneration by using glucose-6-phosphate and the glucose-5-phosphate dehydrogenase from Leuconostoc mesenteroides. J. Am. Chem. Soc. 1981, 103, 48904899,  DOI: 10.1021/ja00406a037
  104. 104
    Johnston, M. R.; Makriyannis, A.; Whitten, K. M.; Drew, O. C.; Best, F. A. Biocatalyzed Regioselective Synthesis in Undergraduate Organic Laboratories: Multistep Synthesis of 2-Arachidonoylglycerol. J. Am. Chem. Soc. 2016, 93, 20802083,  DOI: 10.1021/acs.jchemed.6b00225
  105. 105
    Beers, M.; Archer, C.; Feske, B. D.; Mateer, S. C. Using biocatalysis to integrate organic chemistry into a molecular biology laboratory course. Biochem. Mol. Biol. Educ. 2012, 40, 130137,  DOI: 10.1002/bmb.20578
  106. 106
    Fronier, A. Not Voodoo X.4. http://www.chem.rochester.edu/notvoodoo/ (accessed 2023-02-01).
  107. 107
    Chun, S. W.; Narayan, A. R. H. Biocatalytic, Stereoselective Deuteration of α-Amino Acids and Methyl Esters. ACS Catal. 2020, 10, 74137418,  DOI: 10.1021/acscatal.0c01885
  108. 108
    Rogova, T.; Gabriel, P.; Zavitsanou, S.; Leitch, J. A.; Duarte, F.; Dixon, D. J. Reverse Polarity Reductive Functionalization of Tertiary Amides via a Dual Iridium-Catalyzed Hydrosilylation and Single Electron Transfer Strategy. ACS Catal. 2020, 10, 1143811447,  DOI: 10.1021/acscatal.0c03089
  109. 109
    DeHovitz, J. S.; Loh, Y. Y.; Kautzky, J. A.; Nagao, K.; Meichan, A. J.; Yamauchi, M.; MacMillan, D. W. C.; Hyster, T. K. Static to inducibly dynamic stereocontrol: The convergent use of racemic β-substituted ketones. Science 2020, 369, 11131118,  DOI: 10.1126/science.abc9909
  110. 110
    Key, H. M.; Clark, D. S.; Hartwig, J. F. Generation, Characterization, and Tunable Reactivity of Organometallic Fragments Bound to a Protein Ligand. J. Am. Chem. Soc. 2015, 137, 82618268,  DOI: 10.1021/jacs.5b04431
  111. 111
    Huang, J.; Liu, Z.; Bloomer, B. J.; Clark, D. S.; Mukhopadhyay, A.; Keasling, J. D.; Hartwig, J. F. Unnatural biosynthesis by an engineered microorganism with heterologously expressed natural enzymes and an artificial metalloenzyme. Nat. Chem. 2021, 13, 11861191,  DOI: 10.1038/s41557-021-00801-3
  112. 112
    Gu, Y.; Natoli, S. N.; Liu, Z.; Clark, D. S.; Hartwig, J. F. Site-Selective Functionalization of (sp3)C–H Bonds Catalyzed by Artificial Metalloenzymes Containing an Iridium-Porphyrin Cofactor. Angew. Chem., Int. Ed. 2019, 58, 1395413960,  DOI: 10.1002/anie.201907460

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 24 publications.

  1. Meriem Daghmoum, Nazarii Sabat, Meije Lecoq, Thelma Viraize, Xavier Guinchard. Merging Au(I)-Catalysis and Biocatalysis: Practical and Scalable Synthesis of Chiral Amino Building Blocks from Alkynes. ACS Catalysis 2025, 15 (3) , 2484-2491. https://doi.org/10.1021/acscatal.4c07259
  2. Lanting Xiang, Jennifer Solarczek, Victor Krajka, Hui Liu, Lina Ahlborn, Anett Schallmey, Iordania Constantinou. Evaluating the Potential of Microdroplet Flow in Two-Phase Biocatalysis: A Systematic Study. ACS Applied Materials & Interfaces 2025, 17 (3) , 4776-4787. https://doi.org/10.1021/acsami.4c15647
  3. Di Yang, Chang-Hwa Chiang, Taveechai Wititsuwannakul, Charles L. Brooks III, Paul M. Zimmerman, Alison R. H. Narayan. Engineering the Reaction Pathway of a Non-heme Iron Oxygenase Using Ancestral Sequence Reconstruction. Journal of the American Chemical Society 2024, 146 (50) , 34352-34363. https://doi.org/10.1021/jacs.4c08420
  4. Sabine L. Flitsch, Nicholas J. Turner, Zhi Li. Biocatalysis in Asia and the Pacific. JACS Au 2024, 4 (8) , 2713-2714. https://doi.org/10.1021/jacsau.4c00693
  5. Alexander V. Fejzagić, Matthias Schmid, Philipp Kollmus, Hao Wu, Frederic Buono, Marco Santagostino. G-EnzyBeads for Biocatalysis Screening Facilitated by Volumetric Dosing. Organic Process Research & Development 2024, 28 (8) , 3404-3413. https://doi.org/10.1021/acs.oprd.4c00260
  6. Shubhanshu Jain, Felipe Ospina, Stephan C. Hammer. A New Age of Biocatalysis Enabled by Generic Activation Modes. JACS Au 2024, 4 (6) , 2068-2080. https://doi.org/10.1021/jacsau.4c00247
  7. Kuan Zheng, Ran Hong. Exploring a Non-Comfort Zone: Central Chirality-Generating Macrocyclizations. The Journal of Organic Chemistry 2024, 89 (11) , 7355-7365. https://doi.org/10.1021/acs.joc.4c00030
  8. Ze-Long Mei, Cong-Cong Li, Xu Han, Yu-Chen Tian, Shuo-Han Li, Weidong Liu, Ge Qu, Manfred T. Reetz, Zhoutong Sun, Jun-An Ma, Fa-Guang Zhang. Enzymatic Stereodivergent Access to Fluorinated β-Lactam Pharmacophores via Triple-Parameter Engineered Ketoreductases. ACS Catalysis 2024, 14 (8) , 6358-6368. https://doi.org/10.1021/acscatal.4c00945
  9. Sarah E. Champagne, Chang-Hwa Chiang, Philipp M. Gemmel, Charles L. Brooks III, Alison R. H. Narayan. Biocatalytic Stereoselective Oxidation of 2-Arylindoles. Journal of the American Chemical Society 2024, 146 (4) , 2728-2735. https://doi.org/10.1021/jacs.3c12393
  10. Carter N. Stout, Nour M. Wasfy, Fang Chen, Hans Renata. Charting the Evolution of Chemoenzymatic Strategies in the Syntheses of Complex Natural Products. Journal of the American Chemical Society 2023, 145 (33) , 18161-18181. https://doi.org/10.1021/jacs.3c03422
  11. Cheng Fu, Xinying Wang, Zhuoyi Liu, Yulong Li, Yaping Zhu, Wei Zhang. Trypsin-catalyzed asymmetric aldol reactions of isatins with cyclic ketones and the mechanistic insights on activity differences at theoretical level. Journal of Biotechnology 2025, 402 , 14-20. https://doi.org/10.1016/j.jbiotec.2025.03.001
  12. Jia-Qi Zhu, Jia-Neng Xu, Xin-Ru Bian, Peng Cheng, Zhe Dou, Wen-Ting Dai, Shu-Yun Ju, Ya-Jun Wang. Enhancing the stability and activity of enzymes through layer-by-layer immobilization with nanocomposite hydrogel. Biochemical Engineering Journal 2025, 215 , 109604. https://doi.org/10.1016/j.bej.2024.109604
  13. Yizhu Li, Zhigang Zhang, Lei Xiong, Junping Zheng, Tianxiang Zhu, Jingjing Li, Aizhen Lin, Hongtao Liu. A novel salt-adapted bifunctional glucanase/mannanase from Klebsiella pneumoniae and its application in oligosaccharide production. International Journal of Biological Macromolecules 2025, 296 , 139678. https://doi.org/10.1016/j.ijbiomac.2025.139678
  14. Shenhan Xie, Chenyi Zhang, Yihang Dai, Fengxi Li, Lei Wang, Peng Chen, Zhi Wang. Hemoglobin: An efficient oxidase for the green synthesis of quinazoline derivatives by biocatalytic domino strategy. Molecular Catalysis 2025, 570 , 114680. https://doi.org/10.1016/j.mcat.2024.114680
  15. Hui Wang, Bin Gao, Heli Cheng, Shixuan Cao, Xinyi Ma, Yinjuan Chen, Yuxuan Ye. Unmasking the reverse catalytic activity of ‘ene’-reductases for asymmetric carbonyl desaturation. Nature Chemistry 2025, 17 (1) , 74-82. https://doi.org/10.1038/s41557-024-01671-1
  16. Fengming Shi, Bin Chen, Jinhai Yu, Ruiqi Zhu, Yu Zheng, Xiaoqiang Huang. Enantioselective biosynthesis of vicinal diamines enabled by synergistic photo/biocatalysis consisting of an ene-reductase and a green-light-excited organic dye. Chinese Journal of Catalysis 2025, 68 , 223-229. https://doi.org/10.1016/S1872-2067(24)60168-3
  17. Gaurav P. Kudalkar, Virendra K. Tiwari, David B. Berkowitz. Exploiting Archaeal/Thermostable Enzymes in Synthetic Chemistry: Back to the Future?. ChemCatChem 2024, 16 (21) https://doi.org/10.1002/cctc.202400835
  18. Deepthi Pulivarthi, Ravinder Reddy Patlolla, Joshna Kuni, Raveena Gajjala, Santhosh Nayak Kethavath, Linga Banoth, Subba Reddy B V. Enzymatic kinetic resolution of an intermediate alcohol of the drug Tegoprazan using lipase from Thermomyces lanuginosus. Biocatalysis and Biotransformation 2024, 42 (6) , 653-660. https://doi.org/10.1080/10242422.2024.2326421
  19. Peng Zhang, Shuaiqi Meng, Zhongyu Li, Dennis Hirtz, Lothar Elling, Leilei Zhu, Yu Ji, Ulrich Schwaneberg. A comparative molecular dynamics approach guides the tailoring of glycosyltransferases to meet synthetic applications. Green Chemistry 2024, 26 (16) , 9186-9194. https://doi.org/10.1039/D4GC01508H
  20. John M. Woodley. Biocatalysis in microfluidic systems: an experimental basis for data science. Reaction Chemistry & Engineering 2024, 9 (8) , 2028-2033. https://doi.org/10.1039/D3RE00703K
  21. Marianne B. Haarr, Kotchakorn T. Sriwong, Elaine O'Reilly. Commercial Transaminases for the Asymmetric Synthesis of Bulky Amines. European Journal of Organic Chemistry 2024, 27 (23) https://doi.org/10.1002/ejoc.202400257
  22. D. Alonzo Durante-Salmerón, Isabel Fraile-Gutiérrez, Rubén Gil-Gonzalo, Niuris Acosta, Inmaculada Aranaz, Andrés R. Alcántara. Strategies to Prepare Chitin and Chitosan-Based Bioactive Structures Aided by Deep Eutectic Solvents: A Review. Catalysts 2024, 14 (6) , 371. https://doi.org/10.3390/catal14060371
  23. Jelena Lazic, Vuk Filipovic, Lena Pantelic, Jelena Milovanovic, Sandra Vojnovic, Jasmina Nikodinovic-Runic. Late-stage diversification of bacterial natural products through biocatalysis. Frontiers in Bioengineering and Biotechnology 2024, 12 https://doi.org/10.3389/fbioe.2024.1351583
  24. Priya Kamboj, Vikas Tyagi. Enzymatic Synthesis of Indole‐Based Imidazopyridine using α‐Amylase. ChemBioChem 2024, 25 (6) https://doi.org/10.1002/cbic.202300824

JACS Au

Cite this: JACS Au 2023, 3, 8, 2073–2085
Click to copy citationCitation copied!
https://doi.org/10.1021/jacsau.3c00263
Published July 19, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

6946

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Select examples of chemical structures accessed by using biocatalysis. (A) Compounds formed through C–C bond forming reactions. (B) Compounds accessed using C–H hydroxylation reactions. (C) Hydroxylative dearomatization in the total synthesis of azaphilone natural products. (D) Amino acid C–H hydroxylation in the synthesis of manzacidin C. (E) Multienzyme cascade toward the process-scale total synthesis of islatravir.

    Figure 2

    Figure 2. (A) Historical access to enzymes and enzyme products was a time-consuming process. The understanding of biological systems and the lack of enabling technologies make it difficult to efficiently develop new biocatalysts. (B) Example of an early application of biocatalysts in the synthesis of d-amino acids. This process required the use of a specific strain of bacteria to complete the transformation.

    Figure 3

    Figure 3. Accessing biocatalysts with today’s methods. (A) General workflow for producing enzymes from the gene encoding for an enzyme of interest. The various entry points where a scientist could step into the process are highlighted. (B) Enzymes can be used in biocatalytic reactions at various levels of purity.

    Figure 4

    Figure 4. Chemoenzymatic synthesis of Molnupiravir demonstrated by Merck (right) compared to the previous small-molecule route (left).

    Figure 5

    Figure 5. Example of a chemoenzymatic synthesis used in an undergraduate chemistry laboratory course.

    Figure 6

    Figure 6. Examples of chemoenzymatic and enzymatic methods that result from collaborations between organic and biocatalysis research groups.

  • References


    This article references 112 other publications.

    1. 1
      Atanasov, A. G.; Zotchev, S. B.; Dirsch, V. M.; Orhan, I. E.; Banach, M.; Rollinger, J. M.; Barreca, D.; Weckwerth, W.; Bauer, R.; Bayer, E. A.; Majeed, M.; Bishayee, A.; Bochkov, V.; Bonn, G. K.; Braidy, N.; Bucar, F.; Cifuentes, A.; D’Onofrio, G.; Bodkin, M.; Diederich, M.; Dinkova-Kostova, A. T.; Efferth, T.; El Bairi, K.; Arkells, N.; Fan, T.-P.; Fiebich, B. L.; Freissmuth, M.; Georgiev, M. I.; Gibbons, S.; Godfrey, K. M.; Gruber, C. W.; Heer, J.; Huber, L. A.; Ibanez, E.; Kijjoa, A.; Kiss, A. K.; Lu, A.; Macias, F. A.; Miller, M. J. S.; Mocan, A.; Müller, R.; Nicoletti, F.; Perry, G.; Pittalà, V.; Rastrelli, L.; Ristow, M.; Russo, G. L.; Silva, A. S.; Schuster, D.; Sheridan, H.; Skalicka-Woźniak, K.; Skaltsounis, L.; Sobarzo-Sánchez, E.; Bredt, D. S.; Stuppner, H.; Sureda, A.; Tzvetkov, N. T.; Vacca, R. A.; Aggarwal, B. B.; Battino, M.; Giampieri, F.; Wink, M.; Wolfender, J.-L.; Xiao, J.; Yeung, A. W. K.; Lizard, G.; Popp, M. A.; Heinrich, M.; Berindan-Neagoe, I.; Stadler, M.; Daglia, M.; Verpoorte, R.; Supuran, C. T. Natural products in drug discovery: advances and opportunities. Nat. Rev. Drug Discovery 2021, 20, 200216,  DOI: 10.1038/s41573-020-00114-z
    2. 2
      Dandapani, S.; Marcaurelle, L. A. Grand Challenge Commentary: Accessing new chemical space for ’undruggable’ targets. Nat. Chem. Biol. 2010, 6, 861863,  DOI: 10.1038/nchembio.479
    3. 3
      Rotella, D. P. The Critical Role of Organic Chemistry in Drug Discovery. ACS Chem. Neurosci. 2016, 7, 13151316,  DOI: 10.1021/acschemneuro.6b00280
    4. 4
      Grygorenko, O. O.; Volochnyuk, D. M.; Ryabukhin, S. V.; Judd, D. B. The Symbiotic Relationship Between Drug Discovery and Organic Chemistry. Chem. Eur. J. 2020, 26, 11961237,  DOI: 10.1002/chem.201903232
    5. 5
      Pyser, J. B.; Chakrabarty, S.; Romero, E. O.; Narayan, A. R. H. State-of-the-Art Biocatalysis. ACS Cent. Sci. 2021, 7, 11051116,  DOI: 10.1021/acscentsci.1c00273
    6. 6
      Chakrabarty, S.; Romero, E. O.; Pyser, J. B.; Yazarians, J. A.; Narayan, A. R. H. Chemoenzymatic Total Synthesis of Natural Products. Acc. Chem. Res. 2021, 54, 13741384,  DOI: 10.1021/acs.accounts.0c00810
    7. 7
      Zetzsche, L. E.; Yazarians, J. A.; Chakrabarty, S.; Hinze, M. E.; Murray, L. A. M.; Lukowski, A. L.; Joyce, L. A.; Narayan, A. R. H. Biocatalytic oxidative cross-coupling reactions for biaryl bond formation. Nature 2022, 603, 7985,  DOI: 10.1038/s41586-021-04365-7
    8. 8
      Chakrabarty, S.; Wang, Y.; Perkins, J. C.; Narayan, A. R. H. Scalable biocatalytic C–H oxyfunctionalization reactions. Chem. Soc. Rev. 2020, 49, 81378155,  DOI: 10.1039/D0CS00440E
    9. 9
      Clouthier, C. M.; Pelletier, J. N. Expanding the organic toolbox: a guide to integrating biocatalysis in synthesis. Chem. Soc. Rev. 2012, 41, 15851605,  DOI: 10.1039/c2cs15286j
    10. 10
      Sheldon, R. A.; Brady, D.; Bode, M. L. The Hitchhiker’s guide to biocatalysis: recent advances in the use of enzymes in organic synthesis. Chem. Sci. 2020, 11, 25872605,  DOI: 10.1039/C9SC05746C
    11. 11
      Abdelraheem, E. M. M.; Busch, H.; Hanefeld, U.; Tonin, F. Biocatalysis explained: from pharmaceutical to bulk chemical production. React. Chem. Eng. 2019, 4, 18781894,  DOI: 10.1039/C9RE00301K
    12. 12
      Sheldon, R. A.; Brady, D. Broadening the Scope of Biocatalysis in Sustainable Organic Synthesis. ChemSusChem 2019, 12, 28592881,  DOI: 10.1002/cssc.201900351
    13. 13
      Hughes, G.; Lewis, J. C. Introduction: Biocatalysis in Industry. Chem. Rev. 2018, 118, 13,  DOI: 10.1021/acs.chemrev.7b00741
    14. 14
      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.; Halsey, H. M.; Hughes, G. J.; Jo, J.; Joyce, L. A.; Kolev, J. N.; Liang, J.; Maloney, K. M.; Mann, B. F.; Marshall, N. M.; McLaughlin, M.; Moore, J. C.; Murphy, G. S.; Nawrat, C. C.; Nazor, J.; Novick, S.; Patel, N. R.; Rodriguez-Granillo, A.; Robaire, S. A.; Sherer, E. C.; Truppo, M. D.; Whittaker, A. M.; Verma, D.; Xiao, L.; Xu, Y.; Yang, H. Design of an in vitro biocatalytic cascade for the manufacture of islatravir. Science 2019, 366, 12551259,  DOI: 10.1126/science.aay8484
    15. 15
      Loskot, S. A.; Romney, D. K.; Arnold, F. H.; Stoltz, B. M. Enantioselective Total Synthesis of Nigelladine A via Late-Stage C–H Oxidation Enabled by an Engineered P450 Enzyme. J. Am. Chem. Soc. 2017, 139, 1019610199,  DOI: 10.1021/jacs.7b05196
    16. 16
      Chen, K.; Huang, X.; Kan, S. B. J.; Zhang, R. K.; Arnold, F. H. Enzymatic construction of highly strained carbocycles. Science 2018, 360, 7175,  DOI: 10.1126/science.aar4239
    17. 17
      Zhang, X.; King-Smith, E.; Dong, L.-B.; Yang, L.-C.; Rudolf, J. D.; Shen, B.; Renata, H. Divergent synthesis of complex diterpenes through a hybrid oxidative approach. Science 2020, 369, 799806,  DOI: 10.1126/science.abb8271
    18. 18
      Nakamura, H.; Schultz, E. E.; Balskus, E. P. A new strategy for aromatic ring alkylation in cylindrocyclophane biosynthesis. Nat. Chem. Biol. 2017, 13, 916921,  DOI: 10.1038/nchembio.2421
    19. 19
      Schultz, E. E.; Braffman, N. R.; Luescher, M. U.; Hager, H. H.; Balskus, E. P. Biocatalytic Friedel–Crafts Alkylation Using a Promiscuous Biosynthetic Enzyme. Angew. Chem., Int. Ed. 2019, 58, 31513155,  DOI: 10.1002/anie.201814016
    20. 20
      Lau, W.; Sattely, E. S. Six enzymes from mayapple that complete the biosynthetic pathway to the etoposide aglycone. Science 2015, 349, 12241228,  DOI: 10.1126/science.aac7202
    21. 21
      Lowell, A. N.; DeMars, M. D.; Slocum, S. T.; Yu, F.; Anand, K.; Chemler, J. A.; Korakavi, N.; Priessnitz, J. K.; Park, S. R.; Koch, A. A.; Schultz, P. J.; Sherman, D. H. Chemoenzymatic Total Synthesis and Structural Diversification of Tylactone-Based Macrolide Antibiotics through Late-Stage Polyketide Assembly, Tailoring, and C─H Functionalization. J. Am. Chem. Soc. 2017, 139, 79137920,  DOI: 10.1021/jacs.7b02875
    22. 22
      Lukowski, A. L.; Denomme, N.; Hinze, M. E.; Hall, S.; Isom, L. L.; Narayan, A. R. H. Biocatalytic Detoxification of Paralytic Shellfish Toxins. ACS Chem. Biol. 2019, 14, 941948,  DOI: 10.1021/acschembio.9b00123
    23. 23
      Wang, J.; Zhang, Y.; Liu, H.; Shang, Y.; Zhou, L.; Wei, P.; Yin, W.-B.; Deng, Z.; Qu, X.; Zhou, Q. A biocatalytic hydroxylation-enabled unified approach to C19-hydroxylated steroids. Nat. Commun. 2019, 10, 3378,  DOI: 10.1038/s41467-019-11344-0
    24. 24
      Pyser, J. B.; Baker Dockrey, S. A.; Benítez, A. R.; Joyce, L. A.; Wiscons, R. A.; Smith, J. L.; Narayan, A. R. H. Stereodivergent, Chemoenzymatic Synthesis of Azaphilone Natural Products. J. Am. Chem. Soc. 2019, 141, 1855118559,  DOI: 10.1021/jacs.9b09385
    25. 25
      Zwick, C. R.; Renata, H. Remote C–H Hydroxylation by an α-Ketoglutarate-Dependent Dioxygenase Enables Efficient Chemoenzymatic Synthesis of Manzacidin C and Proline Analogs. J. Am. Chem. Soc. 2018, 140, 11651169,  DOI: 10.1021/jacs.7b12918
    26. 26
      Lukowski, A. L.; Liu, J.; Bridwell-Rabb, J.; Narayan, A. R. H. Structural basis for divergent C–H hydroxylation selectivity in two Rieske oxygenases. Nat. Commun. 2020, 11, 2991,  DOI: 10.1038/s41467-020-16729-0
    27. 27
      Wu, S.; Snajdrova, R.; Moore, J. C.; Baldenius, K.; Bornscheuer, U. T. Biocatalysis: Enzymatic Synthesis for Industrial Applications. Angew. Chem., Int. Ed. 2021, 60, 88119,  DOI: 10.1002/anie.202006648
    28. 28
      McIntosh, J. A.; Benkovics, T.; Silverman, S. M.; Huffman, M. A.; Kong, J.; Maligres, P. E.; Itoh, T.; Yang, H.; Verma, D.; Pan, W.; Ho, H.-I.; Vroom, J.; Knight, A. M.; Hurtak, J. A.; Klapars, A.; Fryszkowska, A.; Morris, W. J.; Strotman, N. A.; Murphy, G. S.; Maloney, K. M.; Fier, P. S. Engineered Ribosyl-1-Kinase Enables Concise Synthesis of Molnupiravir, an Antiviral for COVID-19. ACS Cent. Sci. 2021, 7, 19801985,  DOI: 10.1021/acscentsci.1c00608
    29. 29
      Bornscheuer, U. T.; Buchholz, K. Highlights in Biocatalysis – Historical Landmarks and Current Trends. Eng. Life Sci. 2005, 5, 309323,  DOI: 10.1002/elsc.200520089
    30. 30
      Buchner, E. Alkoholische Gährung ohne Hefezellen. Berichte der deutschen chemischen Gesellschaft 1897, 30, 117124,  DOI: 10.1002/cber.18970300121
    31. 31
      Heckmann, C. M.; Paradisi, F. Looking Back: A Short History of the Discovery of Enzymes and How They Became Powerful Chemical Tools. ChemCatChem. 2020, 12, 60826102,  DOI: 10.1002/cctc.202001107
    32. 32
      Whitesides, G. M. Applications of Cell-Free Enzymes in Organic Synthesis. In Ciba Foundation Symposium 111 - Enzymes in Organic Synthesis; Pitman: London, 1985; pp 7696.
    33. 33
      Olivieri, R.; Fascetti, E.; Angelini, L.; Degen, L. Microbial transformation of racemic hydantoins to d-amino acids. Biotechnol. Bioeng. 1981, 23, 21732183,  DOI: 10.1002/bit.260231002
    34. 34
      Liu, Y.; Zhu, L.; Qi, W.; Yu, B. Biocatalytic production of D-p-hydroxyphenylglycine by optimizing protein expression and cell wall engineering in Escherichia coli. Appl. Microbiol. Biotechnol. 2019, 103, 88398851,  DOI: 10.1007/s00253-019-10155-z
    35. 35
      Buchholz, K. A breakthrough in enzyme technology to fight penicillin resistance─industrial application of penicillin amidase. Appl. Microbiol. Biotechnol. 2016, 100, 38253839,  DOI: 10.1007/s00253-016-7399-6
    36. 36
      Wicks, C.; Hudlicky, T.; Rinner, U. Morphine alkaloids: History, biology, and synthesis. In The Alkaloids: Chemistry and Biology; Knölker, H.-J., Ed.; Academic Press: 2021; Vol. 86, Ch. 2, pp 145342.
    37. 37
      Gulland, J. M.; Robinson, R. Constitution of codeine and thebaine. Mem. Proc. Manchester Lit. Philos. Soc. 1925, 69, 7986
    38. 38
      Armstrong, E. F. Enzymes: A Discovery and its Consequences. Nature 1933, 131, 535537,  DOI: 10.1038/131535a0
    39. 39
      Mohan, R. S.; Mejia, M. P. Environmentally Friendly Organic Chemistry Laboratory Experiments for the Undergraduate Curriculum: A Literature Survey and Assessment. J. Chem. Educ. 2020, 97, 943959,  DOI: 10.1021/acs.jchemed.9b00753
    40. 40
      Heather, J. M.; Chain, B. The sequence of sequencers: The history of sequencing DNA. Genomics 2016, 107, 18,  DOI: 10.1016/j.ygeno.2015.11.003
    41. 41
      Baxevanis, A. D. Using Genomic Databases for Sequence-Based Biological Discovery. Mol. Med. 2003, 9, 185192,  DOI: 10.1007/BF03402130
    42. 42
      GenBank and WGS Statistics. https://www.ncbi.nlm.nih.gov/genbank/statistics/ (accessed 2023-02–01).
    43. 43
      The UniProt Consortium UniProt: the universal protein knowledgebase. Nucleic Acids Res. 2017, 45, D158D169,  DOI: 10.1093/nar/gkw1099
    44. 44
      LeProust, E. M.; Peck, B. J.; Spirin, K.; McCuen, H. B.; Moore, B.; Namsaraev, E.; Caruthers, M. H. Synthesis of high-quality libraries of long (150mer) oligonucleotides by a novel depurination controlled process. Nucleic Acids Res. 2010, 38, 25222540,  DOI: 10.1093/nar/gkq163
    45. 45
      Gibson, D. G.; Young, L.; Chuang, R.-Y.; Venter, J. C.; Hutchison, C. A.; Smith, H. O. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 2009, 6, 343345,  DOI: 10.1038/nmeth.1318
    46. 46
      Loftie-Eaton, W.; Heinisch, T.; Soskine, M.; Champion, E.; Godron, X.; Ybert, T. Novel Variants of Endonuclease V and Uses Thereof. WO2022/090057, 2022.
    47. 47
      Moustafa, K.; Makhzoum, A.; Trémouillaux-Guiller, J. Molecular farming on rescue of pharma industry for next generations. Crit. Rev. Biotechnol. 2016, 36, 840850,  DOI: 10.3109/07388551.2015.1049934
    48. 48
      Swartz, J. R. Advances in Escherichia coli production of therapeutic proteins. Curr. Opin. Biotechnol. 2001, 12, 195201,  DOI: 10.1016/S0958-1669(00)00199-3
    49. 49
      Karbalaei, M.; Rezaee, S. A.; Farsiani, H. Pichia pastoris: A highly successful expression system for optimal synthesis of heterologous proteins. J. Cell. Physiol. 2020, 235, 58675881,  DOI: 10.1002/jcp.29583
    50. 50
      Hunter, M.; Yuan, P.; Vavilala, D.; Fox, M. Optimization of Protein Expression in Mammalian Cells. Curr. Protoc. Protein Sci. 2019, 95, e77  DOI: 10.1002/cpps.77
    51. 51
      Fox, B. G.; Blommel, P. G. Autoinduction of Protein Expression. Curr. Protoc. Protein Sci. 2009, 56, 5.23.15.23.18,  DOI: 10.1002/0471140864.ps0523s56
    52. 52
      Silverman, A. D.; Karim, A. S.; Jewett, M. C. Cell-free gene expression: an expanded repertoire of applications. Nat. Rev. Genet. 2020, 21, 151170,  DOI: 10.1038/s41576-019-0186-3
    53. 53
      de Carvalho, C. C. C. R. Whole cell biocatalysts: essential workers from Nature to the industry. Microb. Biotechnol. 2017, 10, 250263,  DOI: 10.1111/1751-7915.12363
    54. 54
      Alissandratos, A. In vitro multi-enzymatic cascades using recombinant lysates of E. coli: an emerging biocatalysis platform. Biophys. Rev. 2020, 12, 175182,  DOI: 10.1007/s12551-020-00618-3
    55. 55
      Gräslund, S.; Nordlund, P.; Weigelt, J.; Hallberg, B. M.; Bray, J.; Gileadi, O.; Knapp, S.; Oppermann, U.; Arrowsmith, C.; Hui, R.; Ming, J.; dhe-Paganon, S.; Park, H.-w.; Savchenko, A.; Yee, A.; Edwards, A.; Vincentelli, R.; Cambillau, C.; Kim, R.; Kim, S.-H.; Rao, Z.; Shi, Y.; Terwilliger, T. C.; Kim, C.-Y.; Hung, L.-W.; Waldo, G. S.; Peleg, Y.; Albeck, S.; Unger, T.; Dym, O.; Prilusky, J.; Sussman, J. L.; Stevens, R. C.; Lesley, S. A.; Wilson, I. A.; Joachimiak, A.; Collart, F.; Dementieva, I.; Donnelly, M. I.; Eschenfeldt, W. H.; Kim, Y.; Stols, L.; Wu, R.; Zhou, M.; Burley, S. K.; Emtage, J. S.; Sauder, J. M.; Thompson, D.; Bain, K.; Luz, J.; Gheyi, T.; Zhang, F.; Atwell, S.; Almo, S. C.; Bonanno, J. B.; Fiser, A.; Swaminathan, S.; Studier, F. W.; Chance, M. R.; Sali, A.; Acton, T. B.; Xiao, R.; Zhao, L.; Ma, L. C.; Hunt, J. F.; Tong, L.; Cunningham, K.; Inouye, M.; Anderson, S.; Janjua, H.; Shastry, R.; Ho, C. K.; Wang, D.; Wang, H.; Jiang, M.; Montelione, G. T.; Stuart, D. I.; Owens, R. J.; Daenke, S.; Schütz, A.; Heinemann, U.; Yokoyama, S.; Büssow, K.; Gunsalus, K. C.; Structural Genomics, C.; Architecture et Fonction des Macromolécules, B.; Berkeley Structural Genomics, C.; China Structural Genomics, C.; Integrated Center for, S.; Function, I.; Israel Structural Proteomics, C.; Joint Center for Structural, G.; Midwest Center for Structural, G.; New York Structural Genomi, X. R. C. f. S. G.; Northeast Structural Genomics, C.; Oxford Protein Production, F.; Protein Sample Production Facility, M. D. C. f. M. M.; Initiative, R. S. G. P.; Complexes, S. Protein production and purification. Nat. Methods 2008, 5, 135146,  DOI: 10.1038/nmeth.f.202
    56. 56
      Hughes, R. A.; Ellington, A. D. Synthetic DNA Synthesis and Assembly: Putting the Synthetic in Synthetic Biology. Cold Spring Harbor Perspect. Biol. 2017, 9, a023812,  DOI: 10.1101/cshperspect.a023812
    57. 57
      Baker Dockrey, S. A.; Doyon, T. J.; Perkins, J. C.; Narayan, A. R. H. Whole-cell biocatalysis platform for gram-scale oxidative dearomatization of phenols. Chem. Biol. Drug Des. 2019, 93, 12071213,  DOI: 10.1111/cbdd.13443
    58. 58
      Bai, Y.; Yang, X.; Yu, H.; Chen, X. Substrate and Process Engineering for Biocatalytic Synthesis and Facile Purification of Human Milk Oligosaccharides. ChemSusChem 2022, 15, e202102539  DOI: 10.1002/cssc.202102539
    59. 59
      Börner, T.; Grey, C.; Adlercreutz, P. Generic HPLC platform for automated enzyme reaction monitoring: Advancing the assay toolbox for transaminases and other PLP-dependent enzymes. Biotechnol. J. 2016, 11, 10251036,  DOI: 10.1002/biot.201500587
    60. 60
      Claaßen, C.; Mack, K.; Rother, D. Benchtop NMR for Online Reaction Monitoring of the Biocatalytic Synthesis of Aromatic Amino Alcohols. ChemCatChem. 2020, 12, 11901199,  DOI: 10.1002/cctc.201901910
    61. 61
      Bommarius, A. S. Biocatalysis: A Status Report. Annu. Rev. Chem. Biomol. Eng. 2015, 6, 319345,  DOI: 10.1146/annurev-chembioeng-061114-123415
    62. 62
      Reetz, M. T. What are the Limitations of Enzymes in Synthetic Organic Chemistry?. Chem. Rec. 2016, 16, 24492459,  DOI: 10.1002/tcr.201600040
    63. 63
      Stepankova, V.; Bidmanova, S.; Koudelakova, T.; Prokop, Z.; Chaloupkova, R.; Damborsky, J. Strategies for Stabilization of Enzymes in Organic Solvents. ACS Catal. 2013, 3, 28232836,  DOI: 10.1021/cs400684x
    64. 64
      Guzik, U.; Hupert-Kocurek, K.; Wojcieszyńska, D. Immobilization as a Strategy for Improving Enzyme Properties-Application to Oxidoreductases. Molecules 2014, 19, 89959018,  DOI: 10.3390/molecules19078995
    65. 65
      De Santis, P.; Meyer, L.-E.; Kara, S. The rise of continuous flow biocatalysis – fundamentals, very recent developments and future perspectives. React. Chem. Eng. 2020, 5, 21552184,  DOI: 10.1039/D0RE00335B
    66. 66
      France, S. P.; Lewis, R. D.; Martinez, C. A. The Evolving Nature of Biocatalysis in Pharmaceutical Research and Development. JACS Au 2023, 3, 715735,  DOI: 10.1021/jacsau.2c00712
    67. 67
      Zhang, Y.; Xia, B.; Li, Y.; Lin, X.; Wu, Q. Substrate Engineering in Lipase-Catalyzed Selective Polymerization of d-/l-Aspartates and Diols to Prepare Helical Chiral Polyester. Biomacromolecules 2021, 22, 918926,  DOI: 10.1021/acs.biomac.0c01605
    68. 68
      Turner, N. J. Directed evolution drives the next generation of biocatalysts. Nat. Chem. Biol. 2009, 5, 567573,  DOI: 10.1038/nchembio.203
    69. 69
      Cobb, R. E.; Chao, R.; Zhao, H. Directed evolution: Past, present, and future. AIChE J. 2013, 59, 14321440,  DOI: 10.1002/aic.13995
    70. 70
      Steiner, K.; Schwab, H. Recent advances in rational approaches for enzyme engineering. Comput. Struct. Biotechnol. J. 2012, 2, e201209010  DOI: 10.5936/csbj.201209010
    71. 71
      Fernandes, P. Miniaturization in Biocatalysis. Int. J. Mol. Sci. 2010, 11, 858879,  DOI: 10.3390/ijms11030858
    72. 72
      Bell, E. L.; Finnigan, W.; France, S. P.; Green, A. P.; Hayes, M. A.; Hepworth, L. J.; Lovelock, S. L.; Niikura, H.; Osuna, S.; Romero, E.; Ryan, K. S.; Turner, N. J.; Flitsch, S. L. Biocatalysis. Nat. Rev. Methods Primers 2021, 1, 46,  DOI: 10.1038/s43586-021-00044-z
    73. 73
      Duetz, W. A. Microtiter plates as mini-bioreactors: miniaturization of fermentation methods. Trends Microbiol. 2007, 15, 469475,  DOI: 10.1016/j.tim.2007.09.004
    74. 74
      Diefenbach, X. W.; Farasat, I.; Guetschow, E. D.; Welch, C. J.; Kennedy, R. T.; Sun, S.; Moore, J. C. Enabling Biocatalysis by High-Throughput Protein Engineering Using Droplet Microfluidics Coupled to Mass Spectrometry. ACS Omega 2018, 3, 14981508,  DOI: 10.1021/acsomega.7b01973
    75. 75
      Finnigan, W.; Hepworth, L. J.; Flitsch, S. L.; Turner, N. J. RetroBioCat as a computer-aided synthesis planning tool for biocatalytic reactions and cascades. Nat. Catal. 2021, 4, 98104,  DOI: 10.1038/s41929-020-00556-z
    76. 76
      Altschul, S. F.; Gish, W.; Miller, W.; Myers, E. W.; Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403410,  DOI: 10.1016/S0022-2836(05)80360-2
    77. 77
      Cai, X.-H.; Jaroszewski, L.; Wooley, J.; Godzik, A. Internal organization of large protein families: Relationship between the sequence, structure, and function-based clustering. Proteins: Struct., Funct., Bioinf. 2011, 79, 23892402,  DOI: 10.1002/prot.23049
    78. 78
      Sirota, F. L.; Maurer-Stroh, S.; Li, Z.; Eisenhaber, F.; Eisenhaber, B. Functional Classification of Super-Large Families of Enzymes Based on Substrate Binding Pocket Residues for Biocatalysis and Enzyme Engineering Applications. Front. Bioeng. Biotechnol. 2021, 9, 701120,  DOI: 10.3389/fbioe.2021.701120
    79. 79
      Wilding, M.; Peat, T. S.; Kalyaanamoorthy, S.; Newman, J.; Scott, C.; Jermiin, L. S. Reverse engineering: transaminase biocatalyst development using ancestral sequence reconstruction. Green Chem. 2017, 19, 53755380,  DOI: 10.1039/C7GC02343J
    80. 80
      Gumulya, Y.; Baek, J.-M.; Wun, S.-J.; Thomson, R. E. S.; Harris, K. L.; Hunter, D. J. B.; Behrendorff, J. B. Y. H.; Kulig, J.; Zheng, S.; Wu, X.; Wu, B.; Stok, J. E.; De Voss, J. J.; Schenk, G.; Jurva, U.; Andersson, S.; Isin, E. M.; Bodén, M.; Guddat, L.; Gillam, E. M. J. Engineering highly functional thermostable proteins using ancestral sequence reconstruction. Nat. Catal. 2018, 1, 878888,  DOI: 10.1038/s41929-018-0159-5
    81. 81
      Tamura, K.; Peterson, D.; Peterson, N.; Stecher, G.; Nei, M.; Kumar, S. MEGA5: Molecular Evolutionary Genetics Analysis Using Maximum Likelihood, Evolutionary Distance, and Maximum Parsimony Methods. Mol. Biol. Evol. 2011, 28, 27312739,  DOI: 10.1093/molbev/msr121
    82. 82
      Hall, B. G. Building phylogenetic trees from molecular data with MEGA. Mol. Biol. Evol. 2013, 30, 12291235,  DOI: 10.1093/molbev/mst012
    83. 83
      Yates, A. D.; Achuthan, P.; Akanni, W.; Allen, J.; Allen, J.; Alvarez-Jarreta, J.; Amode, M. R.; Armean, I. M.; Azov, A. G.; Bennett, R.; Bhai, J.; Billis, K.; Boddu, S.; Marugán, J. C.; Cummins, C.; Davidson, C.; Dodiya, K.; Fatima, R.; Gall, A.; Giron, C. G.; Gil, L.; Grego, T.; Haggerty, L.; Haskell, E.; Hourlier, T.; Izuogu, O. G.; Janacek, S. H.; Juettemann, T.; Kay, M.; Lavidas, I.; Le, T.; Lemos, D.; Martinez, J. G.; Maurel, T.; McDowall, M.; McMahon, A.; Mohanan, S.; Moore, B.; Nuhn, M.; Oheh, D. N.; Parker, A.; Parton, A.; Patricio, M.; Sakthivel, M. P.; Abdul Salam, A. I.; Schmitt, B. M.; Schuilenburg, H.; Sheppard, D.; Sycheva, M.; Szuba, M.; Taylor, K.; Thormann, A.; Threadgold, G.; Vullo, A.; Walts, B.; Winterbottom, A.; Zadissa, A.; Chakiachvili, M.; Flint, B.; Frankish, A.; Hunt, S. E.; IIsley, G.; Kostadima, M.; Langridge, N.; Loveland, J. E.; Martin, F. J.; Morales, J.; Mudge, J. M.; Muffato, M.; Perry, E.; Ruffier, M.; Trevanion, S. J.; Cunningham, F.; Howe, K. L.; Zerbino, D. R.; Flicek, P. Ensembl 2020. Nucleic Acids Res. 2020, 48, D682D688,  DOI: 10.1093/nar/gkz966
    84. 84
      Atkinson, H. J.; Morris, J. H.; Ferrin, T. E.; Babbitt, P. C. Using Sequence Similarity Networks for Visualization of Relationships Across Diverse Protein Superfamilies. PloS One 2009, 4, e4345  DOI: 10.1371/journal.pone.0004345
    85. 85
      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, 41694182,  DOI: 10.1021/acs.biochem.9b00735
    86. 86
      Doyon, T. J.; Perkins, J. C.; Baker Dockrey, S. A.; Romero, E. O.; Skinner, K. C.; Zimmerman, P. M.; Narayan, A. R. H. Chemoenzymatic o-Quinone Methide Formation. J. Am. Chem. Soc. 2019, 141, 2026920277,  DOI: 10.1021/jacs.9b10474
    87. 87
      Rodriguez Benitez, A.; Narayan, A. R. H. Frontiers in Biocatalysis: Profiling Function across Sequence Space. ACS Cent. Sci. 2019, 5, 17471749,  DOI: 10.1021/acscentsci.9b01112
    88. 88
      Fisher, B. F.; Snodgrass, H. M.; Jones, K. A.; Andorfer, M. C.; Lewis, J. C. Site-Selective C–H Halogenation Using Flavin-Dependent Halogenases Identified via Family-Wide Activity Profiling. ACS Cent. Sci. 2019, 5, 18441856,  DOI: 10.1021/acscentsci.9b00835
    89. 89
      Schülke, K. H.; Ospina, F.; Hörnschemeyer, K.; Gergel, S.; Hammer, S. C. Substrate Profiling of Anion Methyltransferases for Promiscuous Synthesis of S-Adenosylmethionine Analogs from Haloalkanes. ChemBioChem. 2022, 23, e202100632  DOI: 10.1002/cbic.202100632
    90. 90
      Lachowicz, J. C.; Gizzi, A. S.; Almo, S. C.; Grove, T. L. Structural Insight into the Substrate Scope of Viperin and Viperin-like Enzymes from Three Domains of Life. Biochemistry 2021, 60, 21162129,  DOI: 10.1021/acs.biochem.0c00958
    91. 91
      Tararina, M. A.; Allen, K. N. Bioinformatic Analysis of the Flavin-Dependent Amine Oxidase Superfamily: Adaptations for Substrate Specificity and Catalytic Diversity. J. Mol. Biol. 2020, 432, 32693288,  DOI: 10.1016/j.jmb.2020.03.007
    92. 92
      Gerlt, J. A.; Bouvier, J. T.; Davidson, D. B.; Imker, H. J.; Sadkhin, B.; Slater, D. R.; Whalen, K. L. Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST): A web tool for generating protein sequence similarity networks. Biochim. Biophys. Acta, Proteins Proteomics 2015, 1854, 10191037,  DOI: 10.1016/j.bbapap.2015.04.015
    93. 93
      Kohl, M.; Wiese, S.; Warscheid, B. Cytoscape: Software for Visualization and Analysis of Biological Networks. In Data Mining in Proteomics: From Standards to Applications; Hamacher, M., Eisenacher, M., Stephan, C., Eds.; Methods in Molecular Biology; Humana Press: Totowa, NJ, 2011; pp 291303.
    94. 94
      E. J. Corey, X.-M. C. The Logic of Chemical Synthesis; Wiley, :New York, 1989.
    95. 95
      Bommarius, A. S.; Riebel Bommarius, B. R. Biocatalysis; Wiley-VCH Verlag: Weinheim, Germany, 2004; p 634.
    96. 96
      Whittall, J. Applied biocatalysis; John Wiley & Sons: Nashville, TN, 2020; p 560.
    97. 97
      Turner, N. J.; Humphreys, L. Biocatalysis in Organic Synthesis: The Retrosynthesis Approach; Royal Society of Chemistry: 2018.
    98. 98
      Burns, M.; Martinez, C. A.; Vanderplas, B.; Wisdom, R.; Yu, S.; Singer, R. A. A Chemoenzymatic Route to Chiral Intermediates Used in the Multikilogram Synthesis of a Gamma Secretase Inhibitor. Org. Process Res. Dev. 2017, 21, 871877,  DOI: 10.1021/acs.oprd.7b00096
    99. 99
      Raker, J. R.; Holme, T. A. A Historical Analysis of the Curriculum of Organic Chemistry Using ACS Exams as Artifacts. J. Chem. Educ. 2013, 90, 14371442,  DOI: 10.1021/ed400327b
    100. 100
      Cooper, M. M.; Stowe, R. L.; Crandell, O. M.; Klymkowsky, M. W. Organic Chemistry, Life, the Universe and Everything (OCLUE): A Transformed Organic Chemistry Curriculum. J. Chem. Educ. 2019, 96, 18581872,  DOI: 10.1021/acs.jchemed.9b00401
    101. 101
      Raker, J.; Holme, T.; Murphy, K. The ACS Exams Institute Undergraduate Chemistry Anchoring Concepts Content Map II: Organic Chemistry. J. Chem. Educ. 2013, 90, 14431445,  DOI: 10.1021/ed400175w
    102. 102
      Brummund, J.; Sonke, T.; Müller, M. Process Development for Biocatalytic Oxidations Applying Alcohol Dehydrogenases. Org. Process Res. Dev. 2015, 19, 15901595,  DOI: 10.1021/op500307e
    103. 103
      Wong, C.-H.; Whitesides, G. M. Enzyme-catalyzed organic synthesis: NAD(P)H cofactor regeneration by using glucose-6-phosphate and the glucose-5-phosphate dehydrogenase from Leuconostoc mesenteroides. J. Am. Chem. Soc. 1981, 103, 48904899,  DOI: 10.1021/ja00406a037
    104. 104
      Johnston, M. R.; Makriyannis, A.; Whitten, K. M.; Drew, O. C.; Best, F. A. Biocatalyzed Regioselective Synthesis in Undergraduate Organic Laboratories: Multistep Synthesis of 2-Arachidonoylglycerol. J. Am. Chem. Soc. 2016, 93, 20802083,  DOI: 10.1021/acs.jchemed.6b00225
    105. 105
      Beers, M.; Archer, C.; Feske, B. D.; Mateer, S. C. Using biocatalysis to integrate organic chemistry into a molecular biology laboratory course. Biochem. Mol. Biol. Educ. 2012, 40, 130137,  DOI: 10.1002/bmb.20578
    106. 106
      Fronier, A. Not Voodoo X.4. http://www.chem.rochester.edu/notvoodoo/ (accessed 2023-02-01).
    107. 107
      Chun, S. W.; Narayan, A. R. H. Biocatalytic, Stereoselective Deuteration of α-Amino Acids and Methyl Esters. ACS Catal. 2020, 10, 74137418,  DOI: 10.1021/acscatal.0c01885
    108. 108
      Rogova, T.; Gabriel, P.; Zavitsanou, S.; Leitch, J. A.; Duarte, F.; Dixon, D. J. Reverse Polarity Reductive Functionalization of Tertiary Amides via a Dual Iridium-Catalyzed Hydrosilylation and Single Electron Transfer Strategy. ACS Catal. 2020, 10, 1143811447,  DOI: 10.1021/acscatal.0c03089
    109. 109
      DeHovitz, J. S.; Loh, Y. Y.; Kautzky, J. A.; Nagao, K.; Meichan, A. J.; Yamauchi, M.; MacMillan, D. W. C.; Hyster, T. K. Static to inducibly dynamic stereocontrol: The convergent use of racemic β-substituted ketones. Science 2020, 369, 11131118,  DOI: 10.1126/science.abc9909
    110. 110
      Key, H. M.; Clark, D. S.; Hartwig, J. F. Generation, Characterization, and Tunable Reactivity of Organometallic Fragments Bound to a Protein Ligand. J. Am. Chem. Soc. 2015, 137, 82618268,  DOI: 10.1021/jacs.5b04431
    111. 111
      Huang, J.; Liu, Z.; Bloomer, B. J.; Clark, D. S.; Mukhopadhyay, A.; Keasling, J. D.; Hartwig, J. F. Unnatural biosynthesis by an engineered microorganism with heterologously expressed natural enzymes and an artificial metalloenzyme. Nat. Chem. 2021, 13, 11861191,  DOI: 10.1038/s41557-021-00801-3
    112. 112
      Gu, Y.; Natoli, S. N.; Liu, Z.; Clark, D. S.; Hartwig, J. F. Site-Selective Functionalization of (sp3)C–H Bonds Catalyzed by Artificial Metalloenzymes Containing an Iridium-Porphyrin Cofactor. Angew. Chem., Int. Ed. 2019, 58, 1395413960,  DOI: 10.1002/anie.201907460