ACS Publications. Most Trusted. Most Cited. Most Read
Noncanonical Amino Acids in Biocatalysis
My Activity
  • Open Access
Review

Noncanonical Amino Acids in Biocatalysis
Click to copy article linkArticle link copied!

Open PDF

Chemical Reviews

Cite this: Chem. Rev. 2024, 124, 14, 8740–8786
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.chemrev.4c00120
Published July 3, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

In recent years, powerful genetic code reprogramming methods have emerged that allow new functional components to be embedded into proteins as noncanonical amino acid (ncAA) side chains. In this review, we will illustrate how the availability of an expanded set of amino acid building blocks has opened a wealth of new opportunities in enzymology and biocatalysis research. Genetic code reprogramming has provided new insights into enzyme mechanisms by allowing introduction of new spectroscopic probes and the targeted replacement of individual atoms or functional groups. NcAAs have also been used to develop engineered biocatalysts with improved activity, selectivity, and stability, as well as enzymes with artificial regulatory elements that are responsive to external stimuli. Perhaps most ambitiously, the combination of genetic code reprogramming and laboratory evolution has given rise to new classes of enzymes that use ncAAs as key catalytic elements. With the framework for developing ncAA-containing biocatalysts now firmly established, we are optimistic that genetic code reprogramming will become a progressively more powerful tool in the armory of enzyme designers and engineers in the coming years.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society

Special Issue

Published as part of Chemical Reviews virtual special issue “Noncanonical Amino Acids”.

1. Introduction

Click to copy section linkSection link copied!

Nature has evolved proteins with a diverse array of structures and functions using a standard set of twenty amino acid building blocks, as defined by the universal genetic code. Enzymes are a subset of proteins that accelerate the biochemical processes needed to sustain life. These biological catalysts use complex networks of active site residues and well-defined substrate binding pockets to process challenging transformations with unrivalled efficiencies and specificities. (1) In some cases, enzymes recruit additional functional components such as metal ions or organic cofactors to expand upon the limited chemical diversity contained within the standard amino acid side chains. (2)
Over the past decades the fields of biocatalysis and enzymology have benefitted greatly from increasingly sophisticated methods for protein engineering, which have given us unprecedented control over protein sequence, structure and function. For example, site directed mutagenesis has been extensively used to study the roles of individual amino acids in catalysis, advancing in our understanding of enzyme mechanisms. (3) More extensive engineering of proteins has been enabled by methods such as directed evolution (4,5) and computational (re)design, (6−8) leading to the development of biocatalysts with improved stability, selectivity, efficiency and substrate range. (9,10) These approaches have even allowed access to enzymes with mechanisms and catalytic functions that are unknown in nature. (11,12)
Although powerful, most enzyme engineering approaches only make use of nature’s standard alphabet of twenty canonical amino acids (cAAs). These standard amino acids are limited in their chemical diversity, which ultimately restricts our control of protein structure and function. To address this limitation, recent years have seen the emergence of powerful genetic code reprogramming methods that allow introduction of new functional elements into proteins as noncanonical amino acid (ncAA) side chains. (13,14) These methods can be divided into global replacement strategies, where one of the twenty standard amino acids is replaced by a noncanonical structural analogue, or site selective genetic code expansion (GCE) strategies where proteins are synthesized from 21 or more amino acids. To date, these methods have been used to encode hundreds of structurally diverse ncAAs into proteins for diverse applications (Figure 1). (15,16) Genetic code reprogramming has allowed introduction of new spectroscopic handles (17−19) and targeted replacement of individual atoms or functional groups, (20,21) providing new insights into how enzymes operate. These methods have also been used to develop engineered biocatalysts with augmented properties. (22−26) By combining genetic code reprogramming with directed evolution, it has also been possible to embed new modes of catalysis into proteins that would be difficult to access within the constraints of the genetic code. (27,28)

Figure 1

Figure 1. NcAAs discussed in this review. (A) NcAAs incorporated via selective pressure incorporation (SPI), expressed protein ligation (EPL), or solid-phase peptide synthesis (SPPS). (B) NcAAs incorporated by GCE. The orthogonal translation system(s) used to incorporate each ncAA are listed. For several ncAAs, multiple incorporation techniques are discussed in this review, and these are also listed. DAP is incorporated as a precursor featuring a photocleavable group, which matures to DAP upon irradiation at 365 nm. 4-NH2Phe is incorporated as 4-AzPhe, which is then chemically reduced in situ to form 4-NH2Phe.

In this review article, we discuss the different approaches available to incorporate ncAAs into proteins and illustrate how these amino acids have been used for diverse applications in enzyme design, engineering, and characterization.

2. Enabling Technologies

Click to copy section linkSection link copied!

2.1. Post-Translational Protein Modifications

New functionality can be introduced into proteins post-translationally using a variety of chemical or enzymatic methods. (29,30) The thiol group of Cys is commonly targeted for protein modification, usually via alkylation or disulphide bond exchange reactions, due to its high nucleophilicity and relative scarcity in proteins. (31) Although achieving specific and selective modification is often challenging, post-translational protein functionalization is highly versatile and has allowed incorporation of diverse nonproteogenic elements. (32−35) Covalent modification of Cys residues to introduce new catalytic elements has provided artificial enzymes for a range of chemical transformations. For example, flavin analogues have been installed into the active site pocket of the cysteine protease, papain, to generate biocatalysts with oxidoreductase activity. (36−38) Ruthenium and rhodium complexes have also been introduced into protein scaffolds via Cys functionalization, to provide enzymes for ring closing metathesis (39) and hydroformylations, (40−42) respectively.

2.2. Solid-Phase Peptide Synthesis

Functionalized peptides and proteins can be produced using solid-phase peptide synthesis (SPPS), involving sequential coupling of protected amino acids on a solid support. (43) This approach is technically challenging and time-consuming but offers flexibility, allowing synthesis of user-defined peptide sequences containing a variety of canonical and artificial building blocks, including those that are toxic to cells or poorly compatible with the cellular translation machinery (e.g., D-amino acids). (44−47) Peptides synthesized using SPPS are typically restricted to sequences of less than 100 residues; however larger proteins can be constructed by combining SPPS with a variety of ligation strategies including native chemical ligation (NCL) and expressed protein ligation (EPL). (48−54) NCL offers greater flexibility with respect to the number and distribution of ncAAs that can be introduced into the polypeptide chain; however, producing large proteins is both costly and technically challenging. In contrast, EPL allows for more facile production of large modified proteins but is typically limited to the introduction of ncAA-containing synthetic sequences at the C-terminus. These synthetic or semisynthetic methods have been used to create polymer-modified glycoprotein mimetics, (55−58) proteins with modified backbones (59−61) and defined patterns of posttranslational modifications, (62−64) split-proteins, (65) proteins containing fluorescent probes (66−68) and proteins with altered catalytic properties. (47,69,70)

2.3. Selective Pressure Incorporation

Selective pressure incorporation (SPI) exploits the promiscuity of native aminoacyl tRNA synthetases (aaRSs) to aminoacylate a canonical tRNA with a close structural analogue of its cognate cAA, leading to sense codon reassignment (Figure 2). This method results in global replacement of one of the twenty standard amino acids by a ncAA. SPI relies on the use of natural or engineered auxotrophic expression strains that are deficient in a target amino acid. (71−73) A seminal study by Cowie and Cohen reported the quantitative substitution of Met for selenomethionine (SeMet) in proteins produced by the Met auxotroph Escherichia coli (E. coli) ML304d. (74) In this case, the auxotrophic strain displayed normal exponential growth when Met was swapped for SeMet in the growth media, installing SeMet throughout every protein in the bacterial proteome. Other auxotrophic strains have also been developed, enabling the global replacement of Trp with 4-fluorotryptophan (75) (4-FTrp) and valine with α-aminobutyric acid. (76) However, for many ncAAs, global incorporation throughout the host proteome is detrimental to cell growth. To address this challenge, auxotrophic strains are often cultured in defined media containing a low concentration of the cAA, which becomes depleted during cell growth. The ncAA is then supplemented into the media at an appropriate growth state when protein expression is induced. Alternatively, the cells can be harvested prior to induction and resuspended in fresh media containing ncAA but lacking the cAA. To further expand the range of ncAAs that can be incorporated using SPI, the substrate promiscuity of canonical aaRSs has been extended using protein engineering. (76−79)

Figure 2

Figure 2. SPI of ncAAs. SPI employs an auxotrophic expression system to globally replace a target canonical amino acid (cAA) with a close structural analogue. An endogenous aaRS loads its cognate tRNA with the ncAA which is incorporated into proteins. Created with BioRender.com.

SPI has enabled the successful incorporation of a wide range of ncAAs into recombinant proteins for diverse applications (Figure 1A). For example, proteins containing heavy atom analogues of cAAs are routinely used in protein X-ray crystallography to aid experimental phasing. (80−84) Isosteric amino acid analogues have been used to investigate protein folding, stability and activity. (85−90) SPI has also been used to install spectroscopic probes (91−95) and biorthogonal handles for protein conjugations. (77,79,96−101) As detailed in the sections below, this approach has also been widely used to improve enzyme stability and activity.

2.4. Genetic Code Expansion

GCE enables the site selective incorporation of ncAAs in vivo or in vitro, in response to a reassigned codon. In this way proteins containing 21 or more amino acids can be biosynthesized.

2.4.1. In Vitro Genetic Code Expansion

Peptides and proteins containing ncAAs have been ribosomally synthesized in vitro using cell free extracts (e.g., E. coli lysates) or reconstituted protein synthesis systems (e.g., PURExpress). (104−106) Here, tRNAs precharged with the target ncAA are supplemented into an in vitro expression system and decoded by the ribosome in response to a repurposed codon (Figure 3). Early work on in vitro genetic code reprogramming focused on chemically modifying aminoacyl-tRNAs. (107,108) For example, conversion of a Phe-loaded tRNA to the corresponding α-hydroxy acid enabled production of proteins with polyester backbones. (108)

Figure 3

Figure 3. Strategies for the generation of ncAA-loaded tRNAs employ either chemoenzymatic methods (top left, PDB: 2C5U (102)) or Flexizymes (bottom left, PDB: 3CUN (103)). These ncAA-tRNAs can then be incorporated into a polypeptide chain using cell-free expression (CFE) systems (right). Created with BioRender.com.

The development of chemo-enzymatic strategies for synthesizing aminoacyl-tRNAs has provided a more general approach that has expanded the range of ncAAs that can be incorporated into proteins. Here, an N-protected amino acid is first chemically coupled to di(adenosine 5′-)diphosphate or a 5′-phospho-2′-deoxyribocytidylriboadenosine (pdCpA) dinucleotide at the 3′-hydroxyl group. The resulting aminoacylated nucleotide fragment is then ligated to a truncated tRNA using T4 RNA ligase and deprotected to yield the final aminoacyl tRNA. (104,109) Alternatively, aminoacylation ribozymes, which are RNA-based catalysts, can be used to directly charge tRNAs with ncAAs. (110) In particular, Flexizymes have proven to be valuable catalysts for mediating transacylation of activated amino acid esters onto the 3′-hydroxyl group of tRNAs. (111,112) By mixing Flexizymes with tRNA and the respective amino acid ester, nearly any tRNA can be loaded with almost any desired ncAA (Figure 3). (105) To date, efficient and precise ribosomal translation of proteins containing diverse ncAAs has been achieved. (113) Combining Flexizyme genetic code reprogramming technology with mRNA display has proven especially valuable for discovery of macrocyclic peptide binders in drug development. (114) Although powerful, one drawback of these in vitro methods is that they cannot be easily scaled due to the requirement to supply super stoichiometric quantities of aminoacyl-tRNAs as well as the high costs associated with reconstituted in vitro translation systems. These challenges have so far limited the use of these technologies in biocatalysis due to the relatively large quantities of protein required for most applications.

2.4.2. In Vivo Genetic Code Expansion

In vivo genetic expansion methods provide a more scalable approach to the synthesis of ncAA-containing proteins. These methods rely on an orthogonal aminoacyl-tRNA synthetase–tRNA pair to direct the incorporation of an ncAA in response to a repurposed codon introduced into the gene of interest (Figure 4). (115) Of the codons available, the amber codon (UAG) is most commonly used, although reassigned sense codons and quadruplet codons have also been exploited. (116,117) The aaRS/tRNA pair used for ncAA incorporation must be orthogonal in the context of all endogenous aaRS/tRNA pairs within the host organism. This orthogonality is most commonly achieved by importing a heterologous aaRS/tRNA pair from a different domain of life. (118) If necessary, orthogonality can be improved through aaRS and/or tRNA engineering. Finally, the substrate specificity of the heterologous aaRS is engineered so that it uniquely recognizes the ncAA of interest. (118) Although several methods have been reported to engineer aaRSs, (119−123) this is most often achieved by subjecting large active site libraries to a two-stage selection protocol that links cell viability to aaRS activity and specificity. (124) A positive selection step, in which cell survival is dependent upon suppression of an amber codon introduced into the chloramphenicol acetyl transferase (CAT) gene, selects for aaRSs that aminoacylate the suppressor tRNA with the target amino acid. A subsequent negative selection step then removes aaRSs that incorporate cAAs. The positively selected clones are transformed into cells containing a gene encoding a toxic barnase protein with amber mutations at permissive sites. The cells are then grown in the absence of the ncAA, and the presence of aaRS variants that are able to utilize cAAs will result in cell death. The activity and specificity of surviving clones can be quantified using a fluorescence screening assay which relies on suppression of an amber codon introduced into green fluorescent protein (GFP). (125)

Figure 4

Figure 4. Positive and negative selection processes can be used to engineer orthogonal aaRS-tRNA pairs to improve incorporation efficiency and/or specificity. The engineered aaRS catalyzes an aminoacylation reaction between its cognate tRNA and ncAA, with the ncAA added to the growing polypeptide chain during translation in response to a repurposed codon (e.g., the amber stop codon, UAG). Created with BioRender.com.

Using this approach, hundreds of structurally and functionally diverse ncAAs can now be incorporated into proteins (Figure 1B). (126,127) For example, a large number of aromatic ncAAs have been incorporated into proteins produced in E. coli using engineered TyrRS-tRNATyr pairs derived from Methanocaldococcus jannaschii (MjTyrRS-tRNATyr). (128) The structural diversity of encodable ncAAs has been further expanded following the discovery of PylRS-tRNAPyl pairs, which naturally suppress amber codons to incorporate ‘the 22nd amino acid’ pyrrolysine (Pyl, O) in methanogenic archaea. (127,129,130) Due to a high degree of active site plasticity and orthogonality in E. coli, yeast and mammalian cells, PylRS systems have emerged as the most versatile platform for GCE in vivo. (127,131)

3. Probing Enzyme Mechanisms with Noncanonical Amino Acids

Click to copy section linkSection link copied!

Site directed mutagenesis has proven to be an invaluable tool in enzymology, providing a versatile strategy for identifying catalytically important residues. However, standard mutations of active site residues can sometimes lead to unintended structural changes that can complicate interpretation of structure–activity relationships. Furthermore, mutation of key catalytic residues often dramatically reduces activity, which provides important but limited mechanistic insights. The incorporation of ncAAs can open new avenues to study complex biochemical mechanisms by allowing more nuanced perturbations to enzyme structure and function. (14,20,22,132) Such modifications can be used to introduce spectroscopic probes, enable covalent trapping of intermediates and fine-tune active site environments by replacement of specific atoms or functional groups.

3.1. Incorporation of Spectroscopic Handles and Biophysical Probes

Strategic incorporation of ncAAs with distinct spectroscopic and biophysical properties has been used to study enzyme function and structure through a range of techniques, including NMR, (133,134) EPR, (135) IR, (136) and X-ray crystallography. (137) Additionally, ncAAs with fluorescent sidechains such as dansyl-Ala (138) or (7-hydroxycoumarin-4-yl)ethylglycine (Hco), (20) or with biorthogonal handles to which fluorescent molecules can be conjugated, (139−141) have been introduced into peptides and proteins to monitor folding and conformational changes, (142−144) for example in superoxide dismutase, (138) myoglobin (Mb), (145) deubiquitinases, (146) and T4 lysozyme. (147)

3.1.1. Noncanonical Amino Acids to Facilitate Protein Structure Determination

Global incorporation of SeMet through SPI has enabled great advances in protein structure determination by X-ray crystallography, through additional anomalous signals derived from the presence of heavy atoms within the crystal lattice. (148) Genetic incorporation of 4-iodophenylalanine (4-IPhe) (137) or 3-iodotyrosine (3-ITyr) (149) gives a site selective way to introduce an alternative heavy atom to facilitate such single-wavelength anomalous diffraction (SAD) experiments and allow structure determination with less diffraction data. Additionally, a metal-chelating ncAA, (8-hydroxyquinolin-3-yl)alanine (3-HqAla), provided a binding site for Zn(II) within the crystal lattice of O-acetylserine sulfhydrylase, which enabled SAD phasing to determine protein structure. (150)
Recently, X-ray free-electron lasers (XFELs) have allowed structural characterization of transient reaction intermediates using time-resolved serial femtosecond crystallography. Light is a convenient trigger for time-resolved experiments, and several natural photoenzymes have been studied with XFELs. (151−157) Although not yet applied to catalysis, GCE can be used to introduce artificial phototriggers into proteins, such as photocaged AAs or triplet sensitizers. The Wang laboratory used GCE to encode (S)-2-amino-3-(7-fluoro-9-oxo-9H-xanthen-2-yl)propanoic acid (FXO), an ncAA that has a strongly absorbing xanthone side chain that undergoes a photocrosslinking when placed in the hydrophobic pocket of a human liver fatty acid binding protein. (158) Time-resolved crystallography experiments were used to monitor conformational changes in the xanthone side chain between 10 and 300 ns after irradiation. In principle, this approach could be extendable to studying mechanisms of designed enzymes with triplet photosensitizers (vide infra). (159)

3.1.2. Nuclear Magnetic Resonance Studies with Noncanonical Amino Acids

Protein nuclear magnetic resonance (NMR) spectroscopy is a widely used technique that detects signals from spin-active nuclei, and can provide information about reaction kinetics or conformational changes upon ligand binding. To study proteins via NMR, proteins can be expressed in 15N- and/or 13C-enriched media, resulting in uniform labeling of the entire protein. NcAAs can be used to simplify the interpretation of protein NMR data by enabling site-specific isotope labeling or by introducing NMR-active nuclei into proteins, which can be used as sensitive probes that can report on changes to the local conformation and/or electronic environment upon, for example, ligand binding. (160,161) This approach can prove especially valuable for analysis of large proteins, where interpretation of spectra can be complicated. NcAA incorporation has been widely used to selectively install isotopic labels and new spin active nuclei to facilitate protein NMR analysis. In this section we present illustrative examples of where ncAAs have been used to study enzymes through NMR. The reader is directed to excellent reviews for a more comprehensive overview. (133,162)
One early report from 1968 studied staphylococcal nuclease by replacing 14 of the amino acids with their deuterated analogues. (160) The simplified NMR spectra facilitated analysis of the effects of metal ion and inhibitor binding. Site-specific isotopic enrichment was first achieved using a chemically aminoacylated suppressor tRNA to introduce a single 13C-labeled Ala82 into T4 lysozyme in response to a TAG codon. (134) This enabled isolation of the signal from Ala82 in 13C NMR spectra and allowed for observation of changes in shift dispersion upon denaturation. More recently, an engineered tRNAPyl/G1PylRS pair was used to encode 15N-/13C-labeled 7-azatryptophan (7-AzaTrp) into a protease from the Zika virus, establishing a method for achieving site-specific peak assignments in Heteronuclear Single Quantum Coherence (HSQC) spectra. (163)
Three different ncAAs (O-trifluoromethoxytyrosine (O-CF3Tyr), 13C-/15N-labeled O-MeTyr, and 15N-labeled O-nitrobenzyltyrosine (O-NBTyr)) were introduced into eleven positions of the active site of the thioesterase domain of human fatty acid synthase. The differential chemical shifts in HSQC and 19F NMR spectra in response to ligand binding confirmed the ligand binding site, and gave information on protein motion. (164) Similarly, 13C-O-MeTyr was used to provide evidence for the existence of multiple, ligand-dependent conformational states of the cytochrome P450 CYP119. (165)
NcAAs can also be used to introduce spin-active nuclei that do not naturally occur in proteins. Due to the intrinsic NMR sensitivity of 19F and the lack of fluorine-containing cAAs, ncAAs such as 5-fluorotryptophan (5-FTrp), 6-FTrp, 3-FTyr, and 4-FPhe have provided useful spectroscopic handles for studying protein structure and function. (166−175)
SPI was used to introduce trifluoromethionine (F3Met) and difluoromethionine (F2Met) at positions 1, 14 and 107 of lysozyme from bacteriophage λ (LaL). (174,176) Of three sites of ncAA incorporation, the F2Met14 residue in the core of the protein showed the largest chemical shift difference between the two fluorine signals, likely due to restricted side chain movement. Upon binding of an oligosaccharide inhibitor, the two F2Met14 signals diverge even further, suggesting a conformational change to the enzyme in a region closer to one of the two fluorine atoms. (177) In a similar manner, replacement of Met with F3Met in DNA polymerase I from Thermus aquativus (KlenTaq) enabled 19F-NMR studies to study enzyme dynamics. (178) Individual 19F resonances could be resolved and the spectra could be used to distinguish between the free enzyme, the binary complex in the presence of a DNA primer, and the ternary complex with a ddNTP bound. This method has also been used to study conformational changes upon ligand binding at the dimeric interface of nitroreductase and histidinol dehydrogenase. (168) The 19F spectra of nitroreductase containing 4-CF3Phe at position 124 showed a single peak corresponding to the ncAA, which shifted upfield upon binding of an inhibitor due to conformational changes.
Fluorine NMR signals show scalar 19F-19F coupling when in proximity. Orton et al. demonstrated these effects are observable within the hydrophobic core of the E. coli protein cis–trans prolyl-isomerase B by positioning multiple fluorinated ncAAs. (179) Genetic incorporation of two 4-CF3Phe or O-CF3Tyr residues gave observable 19F-19F Total Correlation Spectroscopy (TOCSY) and Double Quantum Filtered Correlation Spectroscopy (DQF-COSY) cross peaks with high sensitivity. These studies suggest that through space 19F-19F couplings between fluorine substituted ncAAs can offer a sensitive tool for studying protein structure and function.
NcAAs can also be used to introduce functional groups that occupy distinct chemical shift regions within 1H NMR spectra. For example, Ekanayake et al. developed an engineered aminoacyl-tRNA synthetase to incorporate Νε-(((trimethylsilyl)methyl)-carbamoyl)lysine (TMSNLys) for study of the dimeric SARS-CoV-2 main protease. (133) 1H NMR signals from the trimethylsilyl (TMS) group of TMSNLys and a second ncAA, Nε-(((trimethylsilyl)-methoxy)carbonyl)lysine (TMSLys) appear at ∼ 0 ppm, a clear spectral region with few features from aqueous buffer or protein environments. Distinct peak shifts were observed upon binding of two ligands, a cyclic peptide inhibitor and Calpeptin. Interestingly, the expected perturbations were not observed with one putative allosteric ligand, pelitinib, suggesting that this ligand does not bind in the region previously proposed in the literature.

3.1.3. Incorporation of Noncanonical Amino Acids for Electron Paramagnetic Resonance Spectroscopy

Electron paramagnetic resonance (EPR) spectroscopy is a powerful biophysical technique for characterizing species with unpaired electrons, such as transition metals complexes or organic radicals. (180) Site-selective spin labeling has been achieved by reacting Cys residues with nitroxide-labeled reagents, although this method relies on the introduction and/or removal of cysteine residues that might have functional significance. (181) As an alternative, GCE can be used to directly encode paramagnetic moieties as amino acid side chains (135) or ncAAs with bioorthogonal handles to which spin labels can be efficiently conjugated. (182−185)
Direct genetic incorporation of a spin-labeled amino acid ((S)-2-amino-6-(1-oxy-2,2,5,5-tetramethylpyrroline-3-carboxamido)hexanoic acid, OPCLys) was achieved using an engineered tRNAPyl/PylRS pair. (135) Purified GFP with the spin-active ncAA at position 39 gives a nitroxide-characteristic EPR signal with a spectral shape indicative for low N–O• mobility, in contrast to that of the free amino acid that shows a spectrum characteristic to that of a fast-tumbling small molecule. Interestingly, quantification of this signal reveals that the intensity is approximately 50% of that expected, likely due to some instability of nitroxide radicals in vivo.
A higher degree of spin-labeling is achievable through post-translational derivatization of ncAAs. Genetic incorporation of 4-acetylphenylalanine (4-AcPhe) has been used to introduce nitroxide spin labels into proteins by acid-catalyzed derivatization of the ncAA side chain with a hydroxylamine nitroxide, first in the model protein T4 lysozyme. (186) This approach has been extended to include alternative ncAA-spin probe combinations. By incorporating ncAAs with either azide or alkyne functional groups, Cu(I)-catalyzed azide-alkyne cycloadditions (CuAACs) or strain-promoted azide-alkyne cycloadditions (SPAACs) can be used to introduce spin labels into proteins. (183,187−191) For example, Gd(III) moieties were conjugated to two sites in the Zika virus NS2B-NS3 protease using 4-azidophenylalanine (4-AzPhe). (192) Double electron–electron resonance (DEER) measurements in the presence and absence of a known inhibitor for this protease revealed similar distance distributions, suggesting the commonly used linked construct of this protease remains in the closed conformation, irrespective of ligand binding.

3.1.4. Noncanonical Amino Acids for Protein Infrared Spectroscopy

Infrared spectroscopy (IR) is a well-established method of study protein structure and function. (193,194) The IR spectra of proteins have a transparent window between 1800 and 2500 cm–1, which can be exploited by incorporating CD (carbon-deuterium), CN, (195−199) SCN, (200) N3, (201−204) and NO2 (205) bonds into proteins to introduce distinct spectroscopic features. In this way, deuterated Met has been used as a vibrational probe to study a range of proteins including myeloperoxidase, (206) dihydrofolate reductase, (207) and plastocyanin. (208)
The addition of ncAAs containing cyano- (19) and nitro- (205) groups to the genetic code affords a method for site specific introduction of vibrational probes, (209) that are sensitive to hydrogen bonding, packing interactions, and protein conformation. (210) The Boxer laboratory introduced 4-cyanophenylalanine (4-CNPhe), 3-cyanophenylalanine (3-CNPhe), and S-cyanohomocysteine (S-CNHcs) into the active site of Ribonuclease S (RNaseS) through SPPS of peptide fragments (residues 1–20) that, when combined with a recombinantly expressed protein fragment (residues 21–124), gave active RNase S variants with no significant changes to enzyme structure or activity. (136) Vibrational Stark Effect (VSE) experiments using each ncAA were used to calibrate the sensitivity of the nitrile stretch to external electric fields, allowing a quantitative interpretation of the frequency shifts, providing a measure of the local electrostatic fields (see section 3.3.2). An expansion of this method used 13C-labeled 4-CNPhe to assess solvatochromic effects and account for hydrogen bonding to the aromatic nitriles, allowing an estimate of the average total electrostatic field within an enzyme active site. (211)
Genetically encoded 3-azido-Tyr (3-N3Tyr) served as a 2-dimensional IR probe in the study of the metalloenzyme DddK, an iron-dependent enzyme that catalyzes the conversion of dimethylsulfoniopropionate to dimethylsulfide. (212) Standard substitutions of Tyr64 abolished catalytic function, however substitution to 3-N3Tyr was well-tolerated and allowed measurement of fs–ps time scale water dynamics. The addition of even low concentrations of the denaturation reagent guanidinium hydrochloride was shown to decrease water confinement in the active site and substantially reduce catalytic activity without affecting the wider protein structure. The authors proposed that Tyr64 is activated as a catalytic base through deprotonation by a neighboring ordered water molecule, and disruption of this process by addition of the denaturation reagent causes loss of catalytic activity.

3.2. Covalent Trapping of Reactive Intermediates and Transient Complexes

NcAAs have been used to covalently trap close analogues of catalytic intermediates and catalytically relevant protein–protein complexes to facilitate their characterization. This approach has provided new structural insights into transient reaction intermediates and can be used to interrogate biosynthetic pathways in vivo.

3.2.1. Capturing Acyl-Enzyme Intermediates

Many enzymes such as Ser hydrolases, Cys proteases, and ubiquitinases proceed via the formation of transient acyl-enzyme (ester or thioester) intermediates. Replacement of the catalytic serine/cysteine residue with 2,3-diaminopropionic acid (DAP) gave a powerful new method of trapping these intermediates with nonhydrolyzable amide bonds (Figure 5A). (213) The structural similarity between DAP and Cys/Ser complicates its direct incorporation through cellular translation. Instead, the Chin lab developed an engineered tRNAPyl/PylRS pair to incorporate a photocaged derivative of DAP that can be deprotected post-translation upon irradiation with light. This approach was used for structural analysis of valinomycin synthetase (Figure 5B,C), a member of the nonribosomal peptide synthetase family. Structural characterization of a covalently linked substrate-enzyme complex by X-ray crystallography revealed a large conformational change to a ‘lid’ region upon substrate acyl-enzyme complex formation.

Figure 5

Figure 5. DAP incorporation into Valinomycin synthetase. (A) Genetically encoded (2S)-2-amino-3-([(2-[1-(6-nitrobenzo[d][1,3]dioxol-5-yl)ethyl]thio)ethoxy)carbonyl] ncAA is photodeprotected by irradiation at 365 nm to give DAP, which forms stable acyl-enzyme intermediates with an amide bond that is resistant to hydrolysis. (B) The active site of Valinomycin synthetase (protein shown as a gray cartoon, PDB: 6ECE (213)) with a noncanonical DAP nucleophile in position 2463 (atom-colored sticks, brown carbons) bound to a dodecadepsipeptide substrate (atom colored sticks, blue carbons). (C) Large structural differences are observed in the lid region of Valinomycin synthetase when bound to a dodecadepsipeptidyl intermediate (gray cartoon, PDB: 6ECE (213)) in comparison to a tetradepsipeptidyl intermediate (blue cartoon, PDB: 6ECD (213)).

Subsequent work from the Chin laboratory described the use of DAP incorporation to create ‘substrate traps’ in mammalian cells. (214) Proteases and other hydrolases equipped with DAP nucleophiles were purified from live cells with substrate fragments bound via stable amide bonds, thus allowing the annotation of proteins of previously unknown function.

3.2.2. Photo-Cross-Linking to Map Protein–Protein Interactions

Benzophenones are powerful photocross-linking agents that have been widely used as photophysical probes. (215) The addition of 4-benzoylphenylalanine (BpAla) to the genetic code using an engineered tRNATyr/MjTyrRS pair allowed introduction of this functional motif into proteins with unparalleled site selectivity. (216) This approach has enabled mapping of transient protein–protein interactions through covalent photocrosslinking, (217) including complexes between glutathione S-transferase (GST) dimers, (218) SecA Atpase and the SecYEG translocon, (219) the transcriptional activator Gal4 and the inhibitor protein Gal80, (220) the large subunit of rubisco and its chaperone RbcX2, (221) acyl carrier proteins and ketosynthases, (222,223) and vaccinia H1-related (VHR) phosphatase dimers. (224) In one notable example, this approach was used to identify a new preinitiation complex in the mechanism of a human mitochondrial RNA polymerase (mtRNAP). (225) Mammalian transcription apparatus contains mtRNAP, a transcription initiation factor-like protein TFB2M, and a major nucleoid protein TFAM. Upon incorporation of BpAla into the C-terminal region of TFAM, photoinduced cross-linking to mtRNAP was observed, dependent on DNA-binding, independent of the major nucleoid protein TFB2M. These data suggest the mtRNAP mechanism involves formation of an initial complex involving mtRNAP, TFAM and promoter DNA.

3.3. Modulating Noncovalent Interactions and Intermediate Lifetimes

GCE allows for selective and targeted substitution of individual functional groups or atoms within an enzyme active site. These molecular edits can then be correlated with changes in catalytic activity and reaction mechanism, to build detailed structure–activity relationships that are not possible to achieve with standard mutagenesis methods. (226,227) In this section, we illustrate examples of how ncAAs have be used to tune the pKa and/or redox potential of selected residues, (228) modulate hydrogen bonding networks (229) or other noncovalent interactions, (230) and to perturb the lifetimes of reactive intermediates. (231)

3.3.1. Tuning the pKa and/or Reduction Potential of Key Residues

Using ncAAs to modulate the pKa and/or reduction potential of key functional residues has provided new insights into the mechanisms of a diverse set of enzymes.
Variants of RNase A with the catalytic His, His12 and His119, replaced by 4-fluorohistidine (4-FHis) were synthesized by stepwise peptide ligation, to give single and double mutants of the enzyme. (48) The pH-rate profiles of each variant for the cleavage of uridyl-3′,5′-adenosine were different to that of the wild type (WT) due to the lower pKa of 4-FHis cf. His. RNase A His12(4-FHis) has a broader pH profile and is able to effect catalysis at ∼ 2 pH units lower than the WT, consistent with its proposed role as a catalytic base. In contrast, replacement of His119, which is thought to serve as a catalytic acid, reduced activity substantially, likely due to the more acidic 4-FHis existing in an inactive nonprotonated state.
Replacement of an active site His of a porcine pancreatic phospholipase A2 (PLA) by 1,2,4-triazole-3-alanine (Taz) using SPI yields a catalytically active enzyme with an altered pH-rate profile to the WT. (232) WT PLA has a pH optimum of 6, with no activity at pH 3. However, at both pH 6 and pH 3 PLA His48Taz is active, at approximately a 5-fold lower rate in comparison to the WT maximum. The mechanism of phospholipase relies on water activation by basic His48. The lower pKa of Taz compared with His ensures that it remains in its active, nonprotonated state at lower pH and is able to activate the nucleophilic water. In contrast, replacement of the only remaining His in a His31Asn/His137Asn double mutant of LaL by Taz had limited effect on the pH-rate profile. (233)
To tune the pKa of an active site tyrosine in the DNA polymerase KlenTaq, 2,3,5-trifluorotyrosine (2,3,5-F3Tyr) was introduced at position 671 using GCE. (234) Structural analysis of KlenTaq in complex with a DNA template containing an abasic site shows Tyr671 occupies the position of the missing base in the DNA template, forming a hydrogen bond to the N3 of purine bases to guide incorporation of dATP and dGTP. (235) Replacement of Tyr671 with 2,3,5-F3Tyr671 using GCE dramatically reduced activity toward single nucleotide incorporations at abasic sites, showing how even subtle perturbations to hydrogen bonding interactions can impact enzyme activity.
Class I ribonucleotide reductases (RNRs) catalyze the conversion of nucleotides to 2′-deoxynucleotides through a mechanism that involves generation of a transient thiyl radical and long-range, reversible radical transfer between subunits α and β (Figure 6A,B). (236) Early docking studies suggested the diferric-Tyr122 radical cofactor in the β-subunit is ∼ 35 Å from the oxidized residue Cys439 in the α subunit, with the proposed radical transfer pathway following Tyr122β → W49β → Tyr365β → Tyr731α → Tyr730α → Cys439α. (237) To study the RNR mechanism, the oligomeric state of the enzyme, and to conclusively identify the redox active residues involved in radical transfer, the Stubbe laboratory has used ncAA incorporation to great effect (Figure 6). (236) Mutation to canonical residues along the electron transfer pathway abolishes radical transfer, however, site specific introduction of fluorinated tyrosines (FnTyr, n = 2–4), (238,239) 3-aminotyrosine (3-NH2Tyr), (240,241) and 3,4-dihydroxyphenylalanine (DOPhe, introduced by EPL methods) (242,243) yields redox-competent variants of the two subunits (α and β). Replacing a Tyr residue along the radical pathway with 3-NH2Tyr or DOPhe traps the radical at that position, as it is unable to oxidize the next amino acid in the chain. (240) UV–vis and EPR spectroscopies were used to quantify the presence of trapped DOPhe• or 3-NH2Tyr• radical signals. The stoichiometry of radical generation gave evidence for the active form of RNR requiring two α/β pairs (i.e., one radical equivalent per α2β2).

Figure 6

Figure 6. Mechanistic studies on RNRs using ncAAs have shed light on the electron transfer pathway and enabled structural characterization of the active form of the multimer. (A) A cryogenic electron microscopy structure of RNR (PDB: 6W4X (231)) in its active α2β2 form was captured using a 2,3,5-F3Tyr122 mutation. The protein chains are shown as cartoons, and GDP and TPP are shown as red and gray spheres, respectively. (B) The mechanism of RNRs, which catalyze the conversion of nucleoside di- and triphosphates to deoxynucleotides. (236) TR = thioredoxin. (C) DEER experiments provided information on the relative distances between the Tyr122 radical in the unreacted α/β pair and radicals on an N3NDP mechanistic inhibitor or radicals trapped on 3-NH2Tyr.

DEER spectroscopy has been used to measure the distance between a trapped 3-NH2Tyr• radical (generated at positions 356 in the β subunit or 730 and 731 in the α subunit) and a Tyr122 radical in the adjacent subunit pair (Figure 6C). (238) Similarly, inter-radical distances have been derived between 3-NO2Tyr122• and Tyr356•, 3-NO2Tyr122•, or Tyr731• in adjacent subunit pairs. (244) Additionally, the distance between Tyr122 and a nitrogen-centered radical covalently attached to Cys439 was measured using a mechanistic inhibitor, 2-azido-2-deoxycytidine diphosphate. The distances derived from DEER experiments are shown with red arrows in Figure 6C. Replacement of Tyr122 with 2,3,5-F3-Tyr or 3-NO2Tyr similarly generates 0.5 equivalents of radicals per α/β pair, further suggesting that RNR activity is reliant on formation of an α2β2 complex. (244) The incorporation of 2,3,5-F3-Tyr at position 122, in combination with a Glu52Gln mutation, enabled structural characterization of the active α2β2 form by cryo-electron microscopy. (231) The 2,3,5-F3-Tyr122 substitution allowed trapping of the radical at Tyr356 along the radical transfer pathway leading to tighter subunit affinity.
Noncanonical Tyr analogues were also used to investigate the role of a key catalytic residue in the Fe(II) and 2-oxoglutarate-dependent enzyme, Verruculogen synthase (FtmOx1). (245) FtmOx1 installs an endoperoxide between C21 and C27 of fumitremorgin B using a high energy ferryl intermediate to break an allylic C21–H bond, followed by dioxygen capture, addition of the C21-O–O• radical to the C27 alkene, and hydrogen atom transfer (HAT) to the resulting C26• radical (Figure 7). Replacement of both active site Tyr residues (68 and 244) with ncAA analogue 3-FTyr facilitated unambiguous identification of Tyr86 as the hydrogen donor to C26•, with Tyr224 playing no essential role. Substitution of Tyr68 caused a blue shift in the sharp Tyr• absorption feature observed by stopped-flow absorption spectroscopy (maximum signal intensity observed at 0.35 s). X-band EPR spectra of samples freeze-quenched after 0.35 s exhibit large hyperfine coupling from 19F ortho to the phenolic oxygen radical that are not seen with the WT enzyme. These effects were not replicated in Tyr244 variants, confirming the role of Tyr68. Substitution of Tyr68 or Tyr244 with both 3-NH2Phe or 3-ClTyr preserved catalytic function. The radical species observed with 4-NH2Phe has a UV–vis absorbance spectrum in agreement with previously observed 4-methylaniline radicals and an EPR signal with additional hyperfine splittings from both 14N and the remaining proton.

Figure 7

Figure 7. Mechanistic proposal for the FtmOx1-catalyzed hydrogen atom transfer from Tyr68 to C26•.

The photoenzyme protochlorophyllide oxidoreductase (POR) catalyzes the reduction of protochlorophyllide to chlorophyllide. (246) Recently, Taylor et al. replaced a key conserved tyrosine in the POR active site, which had previously been proposed to act as a proton donor in POR photocatalysis, with fluorinated Tyr analogues. (247) Tyr fluorination led to a reduction in catalytic activity and impaired substrate binding. However, time-resolved laser spectroscopy reveals that the rate of proton transfer is almost identical in the WT and 3-FTyr193 variants. These data suggest that Tyr193 is unlikely to serve as a proton donor but instead is important for substrate binding.
S-adenosylmethionine (SAM)-dependent methyltransferases are known to form strong carbon-oxygen hydrogen bonds (CH···O) from an active site Tyr residue to their substrate SAM. To understand the importance of this interaction, Horowitz et al. incorporated 4-NH2Phe in place of Tyr335 in the active site of lysine methyltransferase SET7/9. (248) This mutation was shown to be structurally conservative by X-ray crystallography, but reduced SAM binding affinity by ∼ 10,000 fold (to ∼ 1 mΜ). In contrast, SET7/9 Tyr355(4-NH2Phe) was able to bind the coproduct S-adenosyl-L-homocysteine with a similar affinity to that of the WT methyltransferase (∼900 μM), suggesting that the Tyr hydroxyl plays an important role in discriminating substrate vs product binding. The kcat of SET7/9 Tyr355(4-NH2Phe) is 35-fold lower than that of the WT enzyme.

3.3.2. Modulating Electric Fields

Enzymes can apply strong electric fields to activate substrates in their active sites. VSE spectroscopy can be used to measure such effects, provided a suitable spectroscopic handle is available in the substrate or the protein. (249−251) Ketosteroid isomerase (KSI) catalyzes a rate-limiting double bond migration of steroid substrates, which contain an IR-visible carbonyl bond (Figure 8A). (252) The Boxer laboratory used VSE spectroscopy to observe the large electric field applied to the carbonyl of a product analogue (Figure 8C) in the KSI active site. (229) This active site has an oxyanion hole consisting of Asp(H)103 and Tyr16, which forms part of an extended hydrogen-bonding network with Tyr32 and Tyr57 (Figure 8B). (253) Using GCE, each of the three active site Tyr residues were substituted with 3-chlorotyrosine (3-ClTyr). The impact of these substitutions on the local electric field was measured using VSE spectroscopy and correlated with the changes in catalytic activity compared to the WT enzyme. (254) The results show a linear relationship between the electric field applied to the carbonyl and the ΔG, suggesting that the role of the Tyr16 hydroxyl and the adjacent hydrogen bonding network is to tune the electrostatic potential for efficient KSI catalysis.

Figure 8

Figure 8. 3-ClTyr incorporation into Ketosteroid Isomerase (KSI) to tune the active site electric field. (A) The mechanism of KSI. (B) The active site of KSI (PDB: 5KP1 (254)) with the ncAA 3-ClTyr in the active site, shown with orange carbons. The protein backbone is shown as a gray cartoon. Active site residues and the substrate and transition state analogue equilenin are shown as atom-colored sticks, with gray and blue carbons, respectively. (C) The product analogue 19-nortestosterone used for VSE experiments.

3.3.3. Noncanonical Amino Acids to Modulate Cation−π Interactions

Terpene synthases catalyze complex carbocation-based cyclization and rearrangement cascades to give a diverse set of terpenoid natural products. Aromatic ncAAs have been used to modulate cation−π interactions in intermediates in these enzymes.
NcAAs with a range of electron-withdrawing substituents were used to tune cation intermediate stabilization in aristolochene synthase. (255) Replacement of a key active site residue Trp344 by Phe, 4-ClPhe, 4-CF3Phe, and 4-NO2Phe revealed a correlation between the electronic properties of the aromatic side chains and catalytic activity, with more electron-rich substituents leading to higher total turnovers, implicating residue 344 in stabilization of a key cationic intermediate.
A combination of canonical and noncanonical substitutions were used to interrogate cation-π interactions formed by two active site Phe residues in prokaryotic squalene-hopene cyclase (SHC), which mediates the conversion of squalene to hopene via a C22-hopanyl cation intermediate. (256) Interestingly, substitution of Phe365 or Phe605 with the more electron-rich aromatics O-MeTyr, Tyr, and Trp, intended to increase the strength cation−π interactions, increases the catalytic rate at low temperature but leads to a rate reduction at higher temperatures, plausibly due to increased conformational disorder upon introduction of larger aromatic sidechains. To further investigate the effect of the cation-π binding energies on catalytic rates, Phe365 and Phe605 were systematically replaced by 4-FPhe, 3,4-F2Phe, and 3,4,5-F3Phe, which are of similar van der Waals radii to Phe. The catalytic rate decreases with increasing fluorination at either position in a manner proportional to the predicted cation-π binding energies of the aromatic side chains.

3.3.4. Tuning Metal Coordination Environments

Introduction of noncanonical metal coordinating residues into proteins allows the electronic and steric properties of catalytic metal centers to be fine-tuned, providing new structure–activity relationships that cannot be obtained with standard amino acid substitutions.
Nickel-dependent superoxide dismutases (NiSODs) have a metal coordination environment comprised of two Cys ligands, the N-terminal amine, the imidazole of His1, and the backbone amidate of Cys2 (Figure 9). (257) The catalytic nickel center can exist as a five-coordinate pyramidal Ni(III) or a four-coordinate planar Ni(II) species, which lacks the imidazole ligand of His1. To investigate the role of the backbone amidate ligand, a semisynthetic variant of NiSOD with the His1-Cys2 amide replaced by a secondary amine was prepared by NCL (Figure 9). (258) This backbone modification leads to a ∼ 100 fold reduction in the catalytic rate. X-ray absorption near edge structure (XANES) and EPR spectroscopies reveal that the modified variant exists almost exclusively in the four-coordinate Ni(II) state, ca. 50:50 Ni(II)/Ni(III) mixture observed with WT preparations. These experiments suggest that the redox potential of the WT enzyme is finely tuned by the coordination environment to allow facile access to catalytically important Ni(II) and Ni(III) states.

Figure 9

Figure 9. Active site of WT NiSOD (left) and a variant with a secondary amine backbone substitution (right).

The metal coordination environment of [NiFe]-dependent hydrogenase was modified by systematically replacing each of the four coordinating Cys ligands by Sec using allo-tRNAs and selenocysteine synthases. (259,260) In all cases, the catalytic rate of Sec-containing hydrogenases were lower than the WT, retaining between 3 and 14% of the WT activity. However, replacement of the nonbridging Cys576 ligand, which coordinates the nickel cofactor at a position cis to the putative H2 binding site, led to increased oxygen tolerance. The authors suggest that the increased size of the Se atom may prevent O2 binding to the vacant metal coordination site, or alternatively that the more nucleophilic Sec attacks reactive oxygen intermediates to generate a selenoxide that is easily reduced to release H2O. Interestingly, Sec-containing thioredoxin (TR) has also been shown to be more oxygen tolerant than its Cys-containing counterpart. (261) In another study, Wu et al. capitalized on the redox properties of selenocysteine to develop an artificial glutathione peroxidase by replacing the catalytic serine of subtilisin with Sec using a two-step chemical method. Interestingly this Sec modified subtilisin also displays acyl transferase activity. (262)
The identity of the metal-coordinating axial ligand varies across heme enzyme families and is known to play a critical role in controlling catalytic function. (263,264) As such, there is considerable interest in understanding the relationships between proximal ligand electron donation, the structures and reactivities of the Fe(IV)═O intermediates compounds I and II, and overall catalytic function (Figure 10). (265) Natural cytochrome P450s contain Cys-ligated heme iron centers that can perform challenging oxidations of unactivated C–H bonds, chemistry which is inaccessible with histidine-ligated heme peroxidases, even though both enzyme classes use similar Fe(IV)═O intermediates.

Figure 10

Figure 10. Electron donation to the iron center affects ferryl reactivity. (top) Cytochrome P450s are capable of hydrogen atom abstraction by the intermediate Compound I. Increased electron donation through an ncAA selenolate ligand increases the rate compared to WT P450. (bottom) Heme peroxidase compound II is reduced through proton coupled electron transfer. His to MeHis substitution decreases the electron donation to the ferryl intermediate and reduces its proton affinity, slowing the rate of compound II reduction.

The role of the cysteine axial ligand has been explored through its targeted replacement by selenocysteine in a variety of cytochrome P450 enzymes, including P450cam, CYP125, and CYP119. (266−270) Upon mutating the axial Cys to Sec, spectroscopic features and spin states of P450s are similar to the WT enzymes, with minor red-shifts in the Soret maxima and changes in the Raman spectra that demonstrate that Sec is a more electron donating ligand than Cys. It was subsequently shown that the compound I state of Sec-ligated CYP119 was able to cleave C–H bonds more rapidly than the thiolate-ligated WT, providing a direct link between ligand strength and ferryl reactivity. (269) This increased reactivity can be attributed to generation of a more basic compound I state in the Sec-ligated enzyme, increasing the driving force for H-atom abstraction.
A similar approach has been used to derive relationships between ligand strength and ferryl heme reactivity in heme peroxidases. These enzymes have an “imidazolate-like” axial ligand, that is more electron donating than a typical histidine, due to a hydrogen bond between the noncoordinating Nδ atom and a conserved aspartate in the proximal pocket. GCE was used to replace the axial His175 of cytochrome c peroxidase (CcP) by a less-electron donating MeHis, which lacks the proximal hydrogen bonding interaction to Asp235. (271,272) This substitution was shown to be structurally conservative by X-ray crystallography and resulted in a ∼ 20-fold reduction in kcat for ferrous cytc oxidation with negligible changes in KM. In contrast to the situation encountered in P450s, ligand replacement in CcP had minimal impact on the reactivity of compound I (an oxidized neutral ferryl heme and a Trp191 radical cation). Instead, His175MeHis substitution caused a 10-fold reduction in the rate of proton-coupled electron transfer to compound II. This trend can be attributed to weaker electron donation from the axial MeHis ligand affording an electron deficient ferryl-oxygen with reduced proton affinity.
In a subsequent study, Trp51 in the distal pocket of CcP was replaced by a noncanonical 3-(benzothienyl)alanine (STrp), which is structurally similar to Trp but lacks the N–H moiety that forms a hydrogen bond to the ferryl oxygen in compounds I and II. (230) Removal of this hydrogen bond stabilized compound I, but leads to a more basic and reactive compound II state. This increased compound II reactivity manifests in a > 60-fold increase in peroxidase activity toward the small molecule substrate guaiacol (2-methoxyphenol), but interestingly has minimal effect on the rate of oxidation of the biological redox partner, cytc.

3.4. Mimics of Post-Translational Modifications

The structures and functions of native proteins are often tailored through PTMs. In many cases, these covalent modifications cannot easily be recapitulated in recombinant proteins, and as a result their functional significance can be difficult to elucidate. (273) GCE offers a powerful approach to probe the role of PTMs, through selective introduction of ncAAs that are structural mimics of post-translationally modified residues. (274−276)

3.4.1. Lysine and Tyrosine Modifications

Lys residues are subject to a range of reversible PTMs for regulating enzyme function, protein–protein interactions, and cellular localization. (277) The genetic encoding of Nε-lysine derivatives using engineered tRNAPyl/Pyl-tRNA synthetase pairs (278−283) have been used to great effect in the study of individual PTM function in histones (284−289) and enzymes throughout the citric acid cycle. (290−293) For example, acetylation of Lys295 decreases citrate synthase activity by 10-fold, whereas acetylation of Lys238 leads to a 2-fold activity increase. (294) GCE was also used to incorporate Nε-acetylLys (AcLys) into the selenoprotein thioredoxin reductase, to investigate the effect of the PTM on enzyme activity and regulation. (295) A series of thioredoxin reductase variants with acetylated Lys at positions 141, 200, and 307 were all shown to have increased activity compared with the parent enzyme. The authors propose that AcLys incorporation destabilizes inactive multimeric states and favors the active dimeric form of the enzyme.
Dual incorporation of phosphoserine (Sep) (296) and AcLys was used to mimic the coexisting acetylation and phosphorylation PTMs present in native malate dehydrogenase (MLDH). (297) Lysine acetylation at position 140 increased enzyme activity by 3.5-fold, while a single Ser280Sep mutation reduced activity to ca. 30% of the WT enzyme. Simultaneous incorporation of both ncAAs restored the MLDH activity to WT levels, suggesting the two PTMs work in synergy to moderate the activity. Structural modeling of the homotetramer reveals positions 140 and 280 are in proximity at the dimeric interface, opposite a hydrophobic surface region of the protein, suggesting that the PTMs serve to modulate surface charge and stabilize the oligomeric state.
Stop codon suppression (SCS) was used to incorporate Nε-threonyllysine (ThrLys) to study the effect of this recently discovered PTM on Aurora Kinase A activity. (298) Threonylation of Lys162 completely inhibits kinase activity toward a synthetic heptapeptide, with computational modeling suggesting that the modified side chain occupies the ATP binding site. This inhibitory effect can be reversed by the deacetylase Sirtuin 3, which removes the threonylated group from the Lys162.
Human glycolysis enzymes have been found to be heavily lactylated, (299) with lactylation of fructose-bisphosphate aldolase A at the active site Lys147 being a particularly prevalent protein modification. GCE was used to replace Lys147 by a noncanonical lactyllysine (LacLys), which resulted in inhibition of enzyme activity. These results led the authors to propose a lactylation-dependent negative feedback loop in glycolysis, whereby enzymes upstream of an overactivated glycolysis pathway are inhibited leading to reduced glycolytic flux and decreased lactate levels.
Tyrosine PTMs can also be studied using ncAA analogues. 4-carboxymethyl-Phe (4-CmPhe) was used as a mimic of phosphotyrosine to study how the PTM regulates function in the SAM-dependent enzyme protein arginine methyltransferase 1 (PRMT1). (275) Tyr291(4-CmPhe) substitution did not affect the kinetic parameters for methylation of histone H4, but caused a large increase in the KM for two peptides that mimic either the histone H4 tail region or the site of methylation. These data suggest that the phosphorylation of Tyr291 plays a role in the modulation of the substrate specificity of PRMT1 in vivo.

3.4.2. Noncanonical Amino Acids to Mimic Post-Translational Cross-Links

Thioether-bonded tyrosine–cysteine cross-links (Tyr-Cys cross-links) are a PTM common to a range of metalloenzymes, including galactose oxidase, (300) cytochrome c nitrite reductase, (301) and cysteine dioxygenase. (302) To explore the role of this PTM, a model of T. nitratireducens cytochrome c nitrite reductase (TvNiR) was developed in sperm whale Mb with either Tyr or 2-amino-3-(4-hydroxy-3-(methylthio)phenyl)propanoic acid (3-MeS-Tyr) at position 33, to mimic the Tyr-Cys cross-link of the WT nitrite reductase. (303) Hydroxylamine reduction activity increased 4-fold with 3-MeSTyr in place of Tyr33. These activity improvements can be attributed to the 3-methylthio modification lowering the pKa and reduction potential of the active site Tyr, leading to an increased rate of proton coupled electron transfer to the substrate.
NcAA incorporation can also be used to study Tyr-Cys cross-link formation, which is often challenging due to the rapidity of the oxidative cross-linking reaction. The nonheme iron enzyme cysteine dioxygenase (CDO) features a Tyr157-Cys93 cross-link adjacent to the iron center which increases catalytic efficiency by around 10-fold. (304) Site-specific substitution of Tyr157 for 3,5-F2-Tyr and subsequent crystal growth under anaerobic conditions enabled acquisition of a noncross-linked structure while retaining Cys93 (Figure 11). (305) Interestingly, exposure of CDO 3,5-F2-Tyr193 crystals to oxygen resulted in cross-link formation and cleavage of one of the C–F bonds of the ncAA. (306) The positioning of the precross-linked residues in the CDO 3,5-F2-Tyr193 structure suggests that the first target for oxidation during the cross-linking reaction is Cys93, given its proximity to the dioxygen binding site of the metal ion.

Figure 11

Figure 11. Anaerobic X-ray crystal structures of the active sites of Human Cysteine Dioxygenase (CDO, PDB: 6N43 (306)) and CDO Tyr157F2-Tyr (PDB: 6BPR (306)) in complex with the substrate cysteine and NO. CDO and CDO Tyr157F2-Tyr are shown as cartoons in blue and gray, respectively, with key active site residues and the substrate cysteine shown as atom-colored sticks with blue and gray carbon atoms. The noncanonical F2-Tyr157 is shown with orange carbon atoms.

OvoA is a nonheme iron enzyme that catalyzes the oxidative coupling of histidine and cysteine substrates, but also possesses ∼ 10% competing cysteine dioxygenase activity despite lacking a Tyr-Cys cross-link. Interestingly, replacement of active site Tyr417 with 3-MeSTyr alters the product distribution of the enzyme, with 30% of the Cys substrate oxidized to the corresponding sulfinic acid. (307)
The functional importance of post-translational Tyr-His cross-links can also be interrogated using ncAAs. Heme copper oxidases (HCOs) selectively catalyze the four electron reduction of molecular oxygen to water without releasing reactive oxygen species (ROS). (308) The copper center is coordinated by three His residues, one of which forms a cross-link from the Nε atom to the C6 of a neighboring Tyr residue. To better understand the mechanistic significance of this conserved cross-link, (309) Liu et al. incorporated (S)-2-amino-3-(4-hydroxy-3-(1H-imi-dazol-1-yl)phenyl)propanoic acid (ImiTyr) at position 33 in a previously reported functional HCO model based on Mb (vide supra). (310) The resulting enzyme performed over 1000 cycles of oxygen reduction with less than 6% ROS produced, demonstrating 8-fold higher H2O/ROS selectivity and 3 times as many turnovers as the previous HCO model that lacks a Tyr-imidazole cross-link. (311) A related study investigated the effect of Tyr33 pKa in the HCO model by incorporating a series of Tyr analogues by GCE. (312) The rate of O2 reduction was found to be inversely proportional to the pKa of the phenolic proton, consistent with a mechanism in which Tyr33 acts as a proton donor to facilitate O–O cleavage. Interestingly, introduction of 3-MeTyr, which has a similar pKa to Tyr but a lower reduction potential, also gave an increase in rate compared to the Tyr33 variant. (313)

4. Augmenting Function

Click to copy section linkSection link copied!

Augmenting the functions and properties of existing enzymes is a cornerstone of modern biocatalysis. The ability to improve and control enzyme activity, selectivity, and tolerance to process conditions is vital in enabling the full industrial potential of biocatalysts to be realized. Though a large degree of enzyme optimization can be attained using cAAs, the expanded structural and chemical diversity offered by ncAAs can provide new avenues to access optimized biocatalysts. Many enzyme properties can be augmented using ncAAs, including stability and reusability, substrate and product selectivity, catalytic efficiency, and regulation of activity.

4.1. Stability and Immobilization

Enzymes deployed in industrial processes must often withstand elevated temperatures and organic solvents, and are more cost-effective when recovered and reused across multiple reaction cycles. (314,315) NcAAs can be an effective tool to improve enzyme stability and solvent tolerance, and can open up new biorthogonal chemistries to immobilize enzymes in defined orientations on solid supports. A variety of approaches can be used to achieve stabilization and immobilization, reflecting the diversity of ncAAs available and the different attributes of SPI and SCS methodologies. This section reviews examples of enzyme stabilization achieved through ncAA-mediated noncovalent interactions, covalent cross-links within and between macromolecules, and site-specific immobilization on a range of solid supports.

4.1.1. Stabilization via Noncovalent Interactions

NcAAs can be used to stabilize enzymes by strengthening favorable noncovalent interactions without disrupting tertiary structure. Incorporation of halogenated ncAAs has often been shown to increase enzyme thermostability, (316−318) an effect attributed to dipolar interactions or the formation of halogen bonds between the positive σ-hole of the halogen atom and a neighboring lone pair or negatively charged group. (319) Halogenated ncAAs that are isostructural to their canonical counterparts are often well-accommodated in enzymes, with the similarity between C–F and C–H bond lengths being particularly advantageous in this regard. (25,320) Enzyme stabilization using less conservative ncAA modifications has also been investigated, most commonly employing aliphatic side chain substitutions designed to increase hydrophobic interactions. (321,322) The examples discussed below illustrate how ncAA incorporation can increase thermostability and solvent tolerance across a range of enzyme classes, and in some cases that these improvements can also correlate with enhanced catalytic performance.
Thermostable and solvent-tolerant lipases are widely used in many chemical industries. (323,324) Budisa et al. globally substituted 5-FTrp, 3-FTyr or 4-fluorophenylalanine (4-FPhe) for their respective cAAs in Candida antarctica lipase B (CalB) via SPI. (325) Despite some moderate changes in secondary structure (as observed by circular dichroism), all three variants retained lipase activity, and exhibited an increased shelf life. In particular, the 4-FPhe variant was 50% more active than the WT enzyme following storage for several months at 4 °C. A similar study globally incorporated 6-FTrp, 4-FPhe and 4-fluoroproline (4-FPro) simultaneously into Thermoanaerobacter thermohydrosulfuricus lipase (TTL), substituting almost 10% of the residues with fluorinated analogues. (326) Surprisingly, this extensive fluorination had minimal effects on secondary structure or enzyme activity, with the modified enzyme retaining around 60% of the WT activity. A subsequent study investigated the effect of separate global substitution of ten different ncAAs into TLL, including fluorinated Phe, Tyr and Pro, as well as norleucine and azidohomoalanine. (327) All variants evaluated retained at least some activity, with the 3-fluorophenylalanine (3-FPhe) variant displaying 25% greater efficiency than the WT. Interestingly, this variant also exhibited an expanded substrate scope, with a broader range of triacylglycerol chain lengths (C2-C18) accepted compared to the WT (C6-C10). A further study incorporated 13 different ncAA analogues into TTL and investigated the solvent and surfactant tolerance of the resulting variants. (321) In many cases, these substituted variants displayed an increased tolerance to organic solvents compared to the WT, with some exhibiting higher activity in solvent compared to in aqueous buffer. For example, SPI of 4-R-fluoroproline (4-R-FPro) displayed a 4.5-fold increase in activity in tert-butanol and 3.9-fold in acetonitrile. Many of the variants were also more resistant to denaturation, with 4-aminotryptophan (4-NH2Trp) and 3-FPhe substitutions enabling 70% of activity to be retained after incubation in 0.5 M guanidinium chloride, conditions which led to complete inactivation of the WT enzyme.
Transaminases are important biocatalysts due to their ability to produce enantiopure amines, a motif found in a large proportion of pharmaceutical intermediates and chemical commodities. (328,329) SPI of 3-FTyr into ω-transaminase from Vibrio fluvialis JS17 resulted in a soluble enzyme with 20% higher transaminase activity compared to the WT. (317) Thermostability was also enhanced, with a two hour 70 °C incubation reducing the activity of the fluorinated variant to 36%, compared to only 3.3% for the WT. The fluorinated enzyme also demonstrated increased solvent tolerance, with 90% of activity retained in 50% v/v DMSO cosolvent, whereas the WT exhibited a comparable loss in activity in less than 10% DMSO. Importantly, substrate specificity and enantioselectivity were unaffected by 3-FTyr incorporation, enabling the modified transaminase to catalyze the production of enantiopure (S)-1-phenylethylamine on a preparative scale in 89% isolated yield. A later study demonstrated in vivo biosynthesis of 2-fluorotyrosine (2-FTyr) via transgenic tyrosine phenol lyase and subsequent incorporation into two ω-transaminases from Sphaerobacter thermophilus. (330) The resulting enzymes exhibited high levels of 2-FTyr incorporation and increased thermostability compared to their parent enzymes. Again, substrate selectivity was preserved, and in a high temperature kinetic resolution of racemic α-methylbenzylamine the fluorinated variants yielded the R enantiomer of the amine to > 99% ee, compared to under 50% for the WT enzymes.
In some cases, global replacement of residues with halogenated analogues can simultaneously introduce both beneficial and deleterious mutations into an enzyme. Votchitseva et al. globally incorporated 3-FTyr into a phosphotriesterase (PTE), (331) an enzyme class which has garnered interest for their ability to degrade organophosphate-based pesticides and chemical warfare agents for the purposes of bioremediation. (332,333) The modified PTE exhibited an extended pH tolerance down to pH 5.5, plausibly due to the lower pKa of 3-FTyr compared to Tyr, as well as increased thermostability, with preincubation at 60 °C having minimal effect on paraoxon hydrolysis activity compared to ∼ 90% inactivation of the WT. However, under optimum conditions the catalytic efficiency of the variant was ∼ 400-fold lower than WT, plausibly due to Tyr309 substitution in the leaving group binding pocket, highlighting how ncAA incorporation at residue-level specificity can simultaneously introduce desirable and undesirable effects at different sites. A similar study globally replaced all 32 Pro residues in KlenTaq DNA polymerase with 4-R-FPro. (334) Despite the large number of substitutions made and the highly dynamic nature of polymerases during catalysis, the fluorinated variant retained DNA extension activity, exhibiting 94% of the WT dNTP consumption rate and a similar fidelity. However, thermostability was somewhat reduced, with residual activity only half that of the WT after a one hour incubation at 95 °C. Crystallization of the fluorinated polymerase (335) revealed that 4-R-FPro has a tendency to adopt an exo puckered conformation, forcing some positions to adopt this conformation rather than their WT endo pucker (Figure 12A), which may have contributed to the reduction in stability. Interestingly, the authors noted that the fluorinated variant exhibited enhanced crystallizability compared to the WT, with crystals formed under a broader range of conditions. Several surface 4-R-FPro positions were identified where the fluorine atoms were in close proximity to neighboring symmetry chains, potentially forming new crystallization contacts.

Figure 12

Figure 12. NcAA-mediated noncovalent interactions influence enzyme stability. (A) SPI of 4-R-FPro in KlenTaq DNA polymerase switches many Pro puckers from endo to exo, as illustrated by the substitution of Pro555 (left, gray carbons) to 4-R-FPro555 (right, orange carbons) (PDB: 4DLG, 4DLE (335)). (B) Evolutionary trajectory of TFLeu-incorporating CAT (orange bars) starting from WT CAT (gray bar) against the half-life of enzyme inactivation at 60 °C. (C) Structures of T4 lysozyme with canonical Tyr18 (left, gray carbons) and noncanonical 3-ClTyr18 (right, orange carbons). Glu11 and Gly28 backbone atoms shown (white carbons). Halogen bond between Gly28 backbone oxygen and 3-ClTyr18 chlorine atom indicated with a dashed line (PDB: 1L63, (340) 5V7E (339)).

Enzyme engineering can be used to overcome deleterious effects arising from global ncAA incorporation. Panchenko et al. investigated the effects of global leucine replacement with 5′,5′,5′-trifluoroleucine (TFLeu) in CAT. (336) The fluorinated enzyme retained 76% of WT activity but was much less thermostable, exhibiting a 27-fold reduction in half-life at 60 °C. TFLeu-containing CAT was subsequently evolved by screening acetyltransferase activity after preincubation of lysates at 60 °C for one hour. (90) After two rounds, a triple mutant was identified which exhibited comparable thermostability and activity to the nonfluorinated WT CAT (Figure 12B). Introduction of the same mutations into the nonfluorinated enzyme had a much smaller effect on half-life, demonstrating the specificity of the mutations for adapting the enzyme to TFLeu incorporation. Another study globally incorporated 4-FPhe into dimeric S5 PTE, leading to improved activity and stability. (316) The fluorinated enzyme displayed improved catalytic efficiency toward paraoxon and 2-napthylacetate hydrolysis (2.9-fold and 4.7-fold respectively). Preincubation at 65 °C led to complete inactivation of the WT, while the 4-FPhe variant retained up to 41% activity. Differential scanning calorimetry revealed that 4-FPhe incorporation increased the Tm of the two endothermic transitions by 1–2 °C. More notably, for samples subjected to a second heating scan, the WT showed no transitions whereas the two transitions were retained for the fluorinated variant, suggesting that the modified enzyme could refold after thermal denaturation. However, the authors noted the fluorinated variant gave substantially poorer yield of soluble protein compared with the WT enzyme, suggesting that not all 4-FPhe substitutions were beneficial. In a follow-up study, all phenylalanine positions in the PTE were interrogated in silico with Rosetta modeling software. (337) The positions were all replaced with 4-FPhe, then singly mutated to each of the 19 other cAA identities, to discover mutations which gave a better predicted energy score than 4-FPhe. Through this investigation, 4-FPhe104Ala was identified as potentially beneficial to enzyme stability, as it relieved a steric clash arising from 4-FPhe substitution at the dimer interface. Expression of this 4-FPhe-F104A mutant doubled soluble protein yield and increased the thermal unfolding transition temperatures. Paraoxon hydrolase activity was retained at similar levels to WT PTE, and the mutant was more stable under storage conditions, with 66% residual activity recorded after seven days compared to < 50% for WT PTE both with and without 4-FPhe incorporation. Interestingly, introducing the F104A substitution into otherwise unmodified PTE dramatically decreased stability and activity, demonstrating successful prediction of a 4-FPhe context-specific stabilizing mutation by the Rosetta protocol.
SCS offers a more targeted approach to ncAA-mediated enzyme stabilization, by enabling incorporation only at positions where the ncAA confers a benefit. Ohtake et al. systematically replaced each of the 15 Tyr residues in microbial transglutaminase (MTG) with 3-ClTyr via suppression of the UAG stop codon. (318) The single mutants at positions 20 and 62 exhibited increased residual activity relative to the WT following preincubation at 60 °C, while the other mutations tested were either neutral or deleterious. Testing combinations of the beneficial and neutral mutations revealed a triple mutant with a 5.1-fold longer half-life than the WT as well as a 1.4-fold increase in activity. The researchers then demonstrated simultaneous incorporation of an α-hydroxy acid analogue of Nε-allyloxycarbonyllysine (ALOLysOH) immediately after an N-terminal inhibitory peptide domain via sense codon reassignment. This generated a thermostabilized MTG variant which undergoes automaturation in basic conditions due to hydrolysis of the backbone ester bond.
Another study utilized 3-bromotyrosine (3-BrTyr) for the stabilization of azoreductase (AzoR). (338) Each Tyr residue in AzoR was individually replaced with 3-BrTyr, revealing three positions where bromination was beneficial to thermostability. Simultaneous substitution at these three sites yielded a variant with a 13-fold longer half-life than the WT enzyme at 78 °C. The authors also subjected GST to a similar engineering strategy, substituting Tyr residues with 3-ClTyr. Systematic identification and combination of beneficial substitutions generated a variant with four 3-ClTyr incorporation sites that exhibited 79% residual activity after a 10 min preincubation at 60 °C, compared to < 5% for WT GST.
SCS can also be used to install single substitutions explicitly designed to introduce new stabilizing interactions. By supplementing an in vitro translation reaction with chemically aminoacylated suppressor tRNAs, Mendel and co-workers substituted Leu133 in the hydrophobic core of T4 lysozyme with a variety of ncAA analogues chosen to sample a range of sizes and shapes for the aliphatic side chain without disrupting neighboring residues. (322) They correctly predicted that enzyme thermostability would correlate with ncAA side chain size, with norvaline and ethylglycine reducing Tm, and 2-amino-4-methylhexanoic acid and 2-amino-3-cyclopentylpropanoic acid increasing it by up to 4.3 °C, due to their larger side chains establishing more extensive hydrophobic interactions. In a different study, Scholfield et al. engineered a halogen bond into T4 lysozyme by substituting Tyr18 with 4-iodophenylalanine (4-IPhe). (320) Crystal structures of the modified and WT enzyme evidenced the formation of a halogen bond between the iodine atom and the backbone carbonyl of Glu11, coupled with displacement of the aromatic Tyr ring from its WT position. Replacement of the iodine with a methyl group did not result in the same displacement. However, 4-IPhe incorporation slightly decreased overall enzyme stability due to disruption of favorable hydrogen bonds made by the 4-hydroxyl of the WT tyrosine. A subsequent study incorporated 3-ClTyr at the same site. (339) This modification preserved hydrogen bond interactions made by the hydroxyl group and introduced a new halogen bond between the chlorine and the carbonyl oxygen of G28 (Figure 12C), leading to a 1.0 °C increase in Tm and a 15% increase in enzymatic activity at 40 °C compared to the WT.

4.1.2. Stabilization via Covalent Cross-Linking

Disulfide bridge formation through oxidative linking of Cys side chain thiols has long been known to enhance enzyme stability, (341,342) and engineering new disulfide bridges into enzymes has proven an effective method of stabilization. (343,344) However, disulfide links between canonical Cys residues are subject to geometric and chemical constraints that limit the range of enzymes in which these linkages can be successfully installed. (345) The use of noncanonical analogues can overcome some of these constraints, enabling more facile and diverse deployment of covalent cross-linking for enzyme stabilization (Figure 13A). Liu et al. developed noncanonical Tyr analogues derivatized with aliphatic thiols through the para hydroxyl motif, enabling formation of longer disulfide cross-links. (346) SCS was used to incorporate the thiol-containing ncAAs in a library of N-truncated TEM-1 β-lactamase where one random codon per gene was replaced with TAG. The N-terminal truncation was previously shown to destabilize the β-lactamase such that it could no longer confer antibiotic resistance to the host cell above 37 °C. (347) Growth of library-transformed cells in the presence of the lactam antibiotic ampicillin at 40 °C was used to select for active, stabilized enzyme mutants. A single hit was identified containing a disulfide bridge formed between a Cys residue at position 65 and the thiol ncAA O-(4-mercaptobutyl)tyrosine (SbuTyr) at position 184, which increased the Tm by 9 °C compared to the parent enzyme. Analysis of WT crystal structures revealed the cross-link was formed across the hinge region of two half-domains, which had previously been identified as a hotspot for stabilizing mutations. (348) Interestingly, the distance between the two β carbons of residues 65 and 184 was 11.1 Å, far exceeding the maximum distance of ∼ 5.5 Å for a canonical Cys disulfide. (345)

Figure 13

Figure 13. Covalent cross-links mediated by ncAAs. (A) Cross-links generated between cAAs (black) and ncAAs (orange). Cross-linking bonds shown in gray. Top left: canonical Cys-Cys cross-link. Top right: Cys-SbuTyr cross-link. Middle left: Cys-BpAla cross-link. Middle right: amino group-4-NCSPhe cross-link. Bottom left: Cys-O-2-BeTyr cross-link. Bottom right: Cys-4-CaaPhe cross-link. (B) Structures of Cys-O-2-BeTyr cross-link (left) and Cys-4-CaaPhe cross-link (right) in Mb(H64V,V68A), with Tm increases given by one and two cross-links indicated. ncAAs shown with orange carbons and Cys with white carbons (PDB: 7SPE, 7SPH (351)).

The use of ncAAs can also improve the chemical stability of covalent cross-links. Moore and co-workers developed redox-stable thioether “staples” formed between Cys thiols and O-2-bromoethyltyrosine (O-2-BeTyr) residues (349) (Figure 13B). The RosettaMatch algorithm was used to identify positions in a cyclopropanation-competent Mb variant which could accommodate Cys-O-2-BeTyr cross-links in a strain-free configuration, as well as compensatory mutations of surrounding residues to maximize favorable interactions around the staples. Out of nine designs experimentally evaluated, five successfully formed thioether cross-links, all of which exhibited Tm increases ranging from 3.9 to 10.0 °C over the parent enzyme. Substitution of O-2-BeTyr with isosteric but noncross-linking O-propargyltyrosine (O-PaTyr) lowered the Tm values to below WT, confirming the thioether cross-links were essential for the observed stability increases. Combination of the two most stabilized variants resulted in a doubly stapled design with a 16.8 °C higher Tm than the parent enzyme. Impressively, this variant also retained high levels of activity and stereoselectivity for styrene cyclopropanation, as well as exhibiting enhanced organic solvent tolerance. A different study conducted a similar RosettaMatch protocol on a multimodular pullulanase. (350) Installation of Cys-O-2-BeTyr cross-links in the N/C-terminal domains increased Tm by up to 7.0 °C, though hydrolase activity was adversely affected in the majority of designs. Compensatory mutations designed to improve packing around the cross-links yielded a stabilized variant with 44% higher activity than the WT. A later study investigated the use of alternative cross-linking ncAAs. (351) Substitution of O-2-BeTyr in the aformentioned Mb variants with phenylalanine derivatives containing a chloroacetamido (4-CaaPhe), acrylamido (4-AaPhe) or vinylsulfonamido (4-VsaPhe) group led to successful cross-link formation and increased Tm values relative to the WT enzyme. Notably, 4-CaaPhe incorporation led to even larger stability improvements than O-2-BeTyr, with up to a 16.4 °C increase in Tm given by a single Cys-4-CaaPhe cross-link, and a 26.2 °C increase for two cross-links. The 4-CaaPhe-stabilized mutants were shown to retain the same levels of cyclopropanation activity and stereoselectivity as the parent enzyme, with experimental crystal structures showing good agreement with the design models (Figure 13B). Intriguingly, 4-CaaPhe incorporation was also able to rescue some designs which had previously failed to form cross-links using O-2-BeTyr, suggesting a higher stapling efficiency and greater tolerance of 4-CaaPhe to different local environments.
NcAA-mediated cross-links can also be used to stabilize interactions between protein monomers. Homodimeric homoserine O-succinyltransferase (MetA) catalyzes an essential step in methionine biosynthesis in E. coli, but possesses only mesophilic thermostability which leads to inhibited cell growth above 44 °C. (352) Li et al. exploited this property by screening a BpAla scanning library of MetA variants for increased growth of cells lacking native MetA at 44 °C. (353) A consensus Phe21BpAla mutation emerged which conferred substantially faster growth on the host cells. Isolation of the mutant enzyme revealed a remarkable 21 °C increase in Tm over the WT. Position 21 is located near to the dimer interface, suggesting the possibility of a cross-link spanning the two subunits. Mutation of nucleophilic residues in the region facing Phe21 across the interface revealed that Cys90 was essential to the increase in stability, with a Cys90Ser substitution reverting Tm back to that of the WT. Further evidence of a Cys-BpAla cross-link was provided via trapping of the adduct with β-mercaptoethanol and 13C NMR studies. An alternative enzyme-stabilizing covalent linkage strategy was developed by Xuan et al.. (354) 4-Isothiocyanate phenylalanine (4-NCSPhe) was shown to react with Lys side chains, forming covalent thiourea adducts. The salt bridge formed between Lys17 and Asp123 in Mb was replaced with a thiourea cross-link via incorporation of 4-NCSPhe at position 123, which increased the Tm by 4.8 °C compared to the WT. A subsequent study utilized 4-NCSPhe to create stabilizing cross-links in MetA. (355) Using the same elevated temperature cell growth assay described above, a Phe264(4-NCSPhe) mutant was identified which gave a dramatic 24 °C increase in Tm over the WT. Though initially only the monomeric species was observed after SDS-PAGE analysis, it was found that incubation of the mutant at 37 °C led to the appearance of a dimeric species, suggesting that cross-link formation was temperature dependent. The increased thermostability corresponded with increased transferase activity at higher temperatures, with over 60% residual activity recorded at 60 °C, compared to negligible activity of the WT above 50 °C. Proteolysis and mass spectrometry (MS) fragment analysis revealed a temperature-dependent thiourea cross-link formed between 4-NCSPhe264 and N-terminal Pro2. Inspection of a WT crystal structure revealed that these two positions are over 30 Å apart across the dimer interface and are partially buried by neighboring loops, suggesting that thermally induced conformational changes are necessary to bring the two positions into closer contact, rationalizing why cross-linking only occurs at higher temperatures.
Enzyme stability and solvent tolerance can also be enhanced via ncAA-mediated cross-linking to solubilizing polymers. Deiters et al. reported the cross-linking of alkyne-derivatized polyethylene glycol (PEG) to 4-azidophenylalanine (4-AzPhe) incorporated at position 33 in superoxide dismutase. (356) The copper-catalyzed coupling reaction yielded singly PEGylated enzyme with 70–85% conversion, and with no significant loss in activity compared to the WT. A similar approach was used to generate singly PEGylated CalB. (357) Global substitution of the five methionine residues in the enzyme for azidohomoalanine (AzAla) yielded a variant with a single solvent-exposed azido group, as only one of the methionine residues was located on the enzyme surface. Covalent cross-linking to PEG via a CuAAC reaction gave a monofunctionalised species which displayed similar activity to that of the nonfunctionalized enzyme. Another study generated alkyne-decorated polymersomes, which were then cross-linked to AzAla-containing CalB. (358) The enzyme-decorated polymersomes displayed lipase activity, and could also be recovered from the aqueous phase via filtration, providing a facile method of enzyme recycling. Teeuwen et al. utilized the CuAAC reaction to cross-link the same AzAla CalB variant to the alkyne group of homopropargylglycine-containing elastin-like polypeptides (ELPs), which can be used to control enzyme solubility and aggregation. (359) The cross-linked species was purified from unreacted enzyme by taking advantage of the temperature-dependent phase transition of the ELPs, which reversibly form aggregates above a certain temperature. The residual lipase activity of the enzyme-elastin conjugate was found to be ca. 50% of the nonconjugated AzAla CalB variant. Enzymes containing ncAAs can also be cross-linked to solubilizing polymers using the copper-free SPAAC reaction. Debets et al. prepared dibenzocyclooctyne-functionalized PEG, which was then incubated with AzAla CalB for three hours. (360) Full conversion to the PEGylated enzyme conjugate was achieved, with the majority of the enzyme found to di-PEGylated, suggesting that the higher reactivity associated with the strain-promoted reaction enabled cross-linking to buried as well as surface AzAla residues. In a more recent study, Wilding et al. developed a CFE screening system to assess the effects of SPAAC-mediated PEGylation at six different positions in T4 lysozyme via site-specific incorporation of 4-AzPhe. (361) Interestingly, PEGylation efficiency did not correlate well with the solvent accessibility or hydrophobicity of each site, as previously assumed. In most positions, PEGylation led to slight increases in Tm with minimal impacts on activity. Interestingly however, PEGylation at position 91, which is located in an unstructured loop, was detrimental to both thermal stability and enzyme activity, despite this position previously being identified as optimal for lysozyme immobilization. (362)

4.1.3. Immobilization

A well-established strategy for enhancing enzyme stability is via immobilization on a solid support. (363) Immobilisation can increase enzyme thermostability, solvent tolerance and catalyst lifetime, as well as enabling facile biocatalyst recovery and reuse. (315) However, nonspecific immobilization methods, such as covalent cross-linking of surface amino groups with glutaraldehyde, can substantially decrease effective activity by immobilizing the enzymes in unproductive orientations where the active site is occluded, preventing substrate access (25,364) (Figure 14A). Site-specific ncAA incorporation enables precise immobilization of enzyme molecules in productive orientations (Figure 14B), with a range of biorthogonal cross-linking chemistries now available (Figure 14C). In some cases, ncAA-mediated immobilization can also be used to enhance enzyme activity by facilitating rapid electron exchange with functionalized electrodes. (365)

Figure 14

Figure 14. NcAA-mediated enzyme immobilization. (A) Schematic representation of nonspecific enzyme immobilization, mediated by cross-linking at multiple reactive surface residues (gray circles), resulting in multiple enzyme orientations relative to the solid support, as well as enzyme–enzyme cross-linking leading to multilayer immobilization. (B) Schematic representation of site-specific enzyme immobilization, mediated by a ncAA (orange circles) incorporated site specifically, resulting in a monolayer with a single defined enzyme orientation. (C) Immobilization chemistries utilizing ncAAs (orange). From top to bottom: CuAAC, SPAAC, DOPhe–amine coupling, tetrazine-sTCO Diels–Alder cycloaddition, 3-NH2Tyr-acryloyl Diels–Alder cycloaddition, Glaser–Hay alkynyl coupling, and 4-SHPhe-BODIPY coupling.

Azide-alkyne cycloadditions are an efficient method to generate biorthogonal cross-links for enzyme immobilization. Lim et al. site-specifically incorporated 4-ethynylphenylalanine (4-EthPhe) at position 43 of murine dihydrofolate reductase (mDHFR). (366) Optimisation of the CuAAC reaction conditions enabled efficient cross-linking while maximizing residual enzyme activity. Conjugation of 4-EthPhe to a biotin-PEG3-azide linker enabled immobilization of the enzyme to a streptavidin-coated plate which displayed DHFR activity after washing, while a control immobilization where the biotin linker was omitted resulted in an inactive plate. Another study investigated the effects of different immobilization sites on T4 lysozyme activity. (362) Site-specific incorporation of 4-propargyloxyphenylalanine (4-PaPhe) enabled enzyme immobilization onto azide-decorated magnetic beads via CuAAC. 4-PaPhe incorporation at the mouth of the active site led to a substantial 43% reduction in activity upon immobilization, whereas incorporation at more distal sites only resulted in 19–24% reductions. Additionally, one of these variants, Leu91 4-PaPhe, exhibited higher activity than the WT enzyme immobilized via nonspecific epoxy cross-linking. Enzyme lifetime was also improved, with the immobilized Leu91 4-PaPhe variant retaining 79% activity after three freeze–thaw cycles, compared to 42% residual activity for the epoxy-immobilized WT, and < 27% for nonimmobilized controls. A comparable trend was observed after incubation in 2 M urea. In a similar study, Wang et al. individually substituted five Tyr residues for 4-AzPhe in Geobacillus sp. lipase and immobilized the mutants on a cyclooctyne-decorated support via SPAAC. (367) Triolein hydrolysis activity of the mutants increased by 5–37% upon immobilization, apart from a variant cross-linked at a position adjacent to the active site, which exhibited a slight decrease in activity. All of the immobilized mutants compared favorably with the glutaraldehyde-immobilized WT, displaying up to 6.8-fold higher activity. Additionally, the SPAAC-linked mutants were modestly more thermostable than the randomly cross-linked WT, with up to 1.2-fold higher residual activity recorded after a 29 h incubation at 50 °C. Li et al. also utilized 4-AzPhe, immobilizing aldehyde ketone reductase (AKR) variants to a bicyclononyne-functionalized resin. (368) Variants containing 4-AzPhe at one of five positions were tested, all of which displayed 5–16% increased activity upon immobilization and were also more thermostable than the free unmodified enzyme. Impressively, combining all five 4-AzPhe substitutions into one variant enhanced thermostability further, with over 70% residual activity recorded after 24 h incubation at 60 °C, compared to 5% for the free enzyme.
Other biorthogonal chemistries can also be used to immobilize ncAA-substituted enzymes. DOPhe can be oxidized with sodium periodate to form an orthoquinone intermediate, which can then react with an amino-functionalized coupling partner. (369) Deepankumar and co-workers site-specifically incorporated DOPhe into an ω-transaminase and demonstrated successful immobilization on chitosan and polystyrene beads. (370) They combined this approach with global incorporation of 4-R-FPro, which substantially increased thermostability, doubling the half-life of the enzyme at 70 °C, as well as modestly improving tolerance of a range of organic solvents. Immobilisation of the stabilized variant reduced activity by less than 10% and enabled efficient biocatalyst recovery, with almost complete preservation of enzyme activity following 10 cycles of a 12 h kinetic resolution process. Another study utilized the inverse-electron-demand Diels–Alder (IEDDA) reaction between 1,2,4,5-tetrazines and strained trans-cyclooctenes (sTCOs) for immobilization of thermostable carbonic anhydrase II (tsCA). (371) This conjugation strategy was chosen for its high rate constant (∼72,000 M–1 s–1) which allowed a reduction in immobilization time, thereby minimizing nonspecific adsorption to the solid support. A previously developed tetrazine-containing ncAA, Tet2.0, (372) was incorporated into tsCA separately at three surface positions using SCS, chosen to orient the active site toward bulk solvent, toward the solid support, or parallel to the support. The cross-linking reaction between these variants and sTCO-decorated magnetic beads was found to be efficient enough to allow substoichiometric quantities of enzyme to be used, enabling precise control of the amount of enzyme immobilized on the beads. The orientation of the immobilized enzyme was found to control hydrolysis activity, with the solvent-facing variant retaining 90% activity compared to preimmobilization, the parallel-facing variant 75%, and the support-facing variant 60%. The researchers also demonstrated successful enzyme-limited immobilization onto an sTCO-functionalized flat surface, demonstrating the versatility of the technique. Switzer et al. used Glaser-Hay alkynyl coupling to immobilize a 4-PaPhe-substituted carboxylesterase to an alkyne-decorated Sepharose resin. (373) Four surface Tyr residues were identified and individually substituted with 4-PaPhe, with immobilization of all four variants yielding catalytically active resins. Solvent tolerance was greatly improved by immobilization, with the most active variant displaying increased activity in neat THF compared to neat aqueous buffer, whereas activity of the free WT enzyme was reduced by over 60% in a 1:1 mixture of buffer and THF. The immobilized carboxylesterase also displayed impressive recyclability and lifetime, with almost no activity lost after three years of storage and 18 reaction cycles, compared to the free WT which only retained 20% activity after 6 months of storage.
NcAA-mediated immobilization can also enhance direct electron transfer (DET) between redox enzymes and electrodes, enabling more efficient electrocatalysis, current generation, and reaction monitoring. Ray et al. site-specifically incorporated 3-aminotyrosine (3-NH2Tyr) at a surface position in Mb, which was used to immobilize the protein on an acryloyl-derivatized gold electrode via oxidation of the ncAA to the 2-iminoquinone and subsequent Diels–Alder cycloaddition. (374) As a control, WT Mb was randomly immobilized using a carbodiimide reagent. Atomic force microscopy (AFM) characterization revealed that the ncAA-immobilized protein formed a monolayer with an average feature height of 5 nm, whereas the WT was immobilized in an uneven, multilayer arrangement with an average feature height of 175 nm. Cyclic voltammetry measurements indicated DET could occur between the electrode and both forms of immobilized Mb. Electrocatalytic oxidation of thioanisole was then investigated, with randomly immobilized WT Mb exhibiting 87% conversion and the 3-NH2Tyr-immobilized Mb a comparable 81%, despite the far fewer number of enzyme molecules immobilized in the monolayer arrangement. A different study utilized the SPAAC reaction to immobilize 4-AzPhe-containing mutants of a small laccase to cyclooctyne-derivatized carbon nanotubes. (365) During electro-biocatalytic oxidation of 2,6-dimethoxyphenol, an electron transfer efficiency of 28.7% was recorded when 4-AzPhe was incorporated at position 47, compared to 1.9–7.5% at three other surface positions, and 1.3–3.4% when the enzyme was randomly immobilized by cross-linking surface amino groups to succinimidyl ester-decorated nanotubes. The success of this immobilization position was attributed to its proximity to a water channel connecting the catalytic copper cluster in the active site with the enzyme surface, which may have facilitated electron transfer to the electrode. The 4-AzPhe-immobilized enzyme also exhibited a longer lifetime, with catalytic current decreasing by only 13.6% after 8 days of storage at room temperature, compared to 47.0% for the randomly immobilized control. In another study, Xia and co-workers developed an orthogonal aaRS to enable site-selective incorporation of 4-thiolphenylalanine (4-SHPhe) into tryptophan oxidase (TrpOx). (375) This allowed the enzyme to be derivatized with a boron-dipyrromethene (BODIPY) moiety via a nucleophilic substitution reaction with the ncAA thiol. BODIPY molecules can bind to carbon nanotubes by forming strong π-π interactions, enabling immobilization of BODIPY-conjugated enzymes. AFM characterization of nanotubes incubated with 4-SHPhe-modified enzyme revealed 51.5% coverage when BODIPY was present and only 12.3% when it was absent, with the residual coverage attributed to nonspecific adsorption. On addition of the substrate tryptophan, an oxidative current was detected only from the electrode incubated with both BODIPY and enzyme, demonstrating the immobilized enzyme was functional and that site-specific immobilization was required for efficient electron transfer.

4.2. Improving Enzyme Selectivity

The introduction of ncAAs into enzymes can be used to reshape substrate binding pockets, leading to altered substrate profiles, product distributions and stereoselectivities.

4.2.1. Altering Substrate Profiles

The introduction of ncAAs has been used to alter the substrate preference of PikC, a P450 enzyme that catalyzes a C(sp3)–H oxidation step in the biosynthesis of macrolide antibiotics. (376) PikC substrates require an amino-sugar motif appended during the previous biosynthetic step by the glycosyltransferase DesVII, but it was demonstrated that installation of an ncAA could alleviate this requirement (Figure 15). Surveying a range of aromatic ncAAs at several positions around the substrate binding pocket revealed the mutation His238(4-AcPhe) enabled the site-selective hydroxylation of two different aglycone substrates (10-deoxymethonolide and narbonolide), differentiated from their respective final antibiotic by only the amino-sugar moiety (Figure 15, right). Oxidation site-selectivity for 10-deoxymethonolide could be improved further to 19:1 (methynolide:neomethynolide) by transplanting two known mutations (Glu85Gln and Glu94Gln) into the His238(4-AcPhe) construct. By contrast, the WT PikC oxidizes the glycosylated analogue of 10-deoxymethonolide in a ∼ 1:1 ratio. Structural analyses and docking studies suggested the changes in substrate specificity induced by 4-AcPhe incorporation were due to the acetyl motif supplying hydrogen-bonding and hydrophobic interactions that are normally provided by the desosamine motif in the native substrates. Additional hydrogen-bonding interactions were observed between the C5-hydroxyl group in narbonolide and the acetyl group. Interestingly, the His238(4-AcPhe) mutation also led to improvements in activity and site-selectivity toward the native substrate. The ability to reprogram biosynthetic sequences by altering substrate specificities could have applications in structural diversification for analogue discovery and enabling combinatorial biosynthesis.

Figure 15

Figure 15. Introduction of 4-AcPhe into PikC, a CYP450 enzyme, enabled biosynthetic reprogramming through allowing C(sp3)–H oxidation to occur in the absence of an amino-sugar moiety (brown).

4.2.2. Altering Product Distributions

Fasan and co-workers have investigated the effect of installing ncAAs into a P450 on oxidation regioselectivity. (377) Significant changes in oxidation product distribution were achieved when 4-AcPhe or O-BnTyr were incorporated at one of eleven positions located around the active site or proximal to the prosthetic heme ligand in P450BM3 (Figure 16). For (S)-ibuprofen methyl ester, the Ala78(4-AcPhe) mutation shifted the oxidation product ratio further toward the C1′ oxidation product (1a), whereas Ala181O-BnTyr led to a reversal in regioselectivity, furnishing C2′ oxidation product (1b) as the major species (15:85 1a:1b). For (+)-nootkatone, a new oxidation product, 2c, that was not formed by the WT enzyme was observed; the Ala78(4-AcPhe) mutant formed this as the major product (32:68 2a:2c). The Ala82(4-AcPhe) mutant reduced the native epoxidation activity and concomitantly increased C(sp3)–H oxidation activity; 2a and 2b were formed in a 38:62 ratio respectively. In these cases, canonical substitutions to Tyr at positions 78 and 181 did not give rise to altered product distributions, demonstrating how substrate binding pockets can be more extensively remodelled using ncAAs compared with using cAAs alone on the regioselectivity of P450 oxidations. (377)

Figure 16

Figure 16. Incorporation of ncAAs at various positions within P450BM3 alters the oxidation product distributions for (S)-ibuprofen-OMe and (+)-nootkatone substrates.

4.2.3. Improving Stereoselectivity

NcAA incorporation has been used to improve the stereoselectivity of Acinetobacter baylyi DKR. (378) The WT DKR used in this study catalyzes the reduction of 2-chloro-2-phenylethanone with modest selectivity (9% ee) in favor of the (R)-enantiomer. Mutating Trp222 to smaller cAAs led to a switch in enantio-preference to the (S)-enantiomer, suggesting that steric bulk at position 222 could be a determinant of facial selectivity for hydride delivery. A number of ncAAs with bulky aromatic side chains were incorporated at this position, with the largest improvement in stereoselectivity achieved with O-tBuTyr (up to 34% ee), the bulkiest ncAA tested.
An expanded genetic code has also been used to improve the diastereoselectivity of Pseudomonas alcaligenes lipase during menthol propionate hydrolysis through enhancing substrate specificity for L-menthol propionate over seven other stereoisomers present in the mixture. (379) Guided by MD simulations, a number of substituted aromatic ncAAs were installed at different positions around the active site. The most improved mutant identified was Ala253(2-BrPhe) which selectively catalyzed L-menthol propionate hydrolysis with 94% conversion and > 95% diastereomer selectivity. In another study, Kourist and co-workers investigated the effect of five different ncAAs at ten positions around the substrate binding pocket of the hydrolase Pseudomonas fluorescens esterase. (380) A split-GFP assay was employed to rapidly assess folding of the fifty possible variants, leading to the identification of several sites that are intolerant of ncAA incorporation. The activity and selectivity of the folded, soluble variants were assessed in the kinetic resolution of ethyl 3-phenylbutyrate. While conversions were lower than those achieved with the WT enzyme, enantioselectivity was improved from 27 to 68% ee with a Phe162(2-NapAla) variant, while an Ile224(4-AzPhe) substitution led to a switch in enantio-preference.

4.3. Improving Kinetic Parameters

In addition to improving enzyme stability and selectivity, the introduction of one or multiple ncAAs has been shown to improve enzyme kinetic parameters in effects not replicable with cAAs. Substitutions using both SPI and GCE have been shown to increase kcat, increase total turnover number (TTN), and improve substrate binding.

4.3.1. Noncanonical Ligands in Artificial Metalloenzymes

Heme peroxidases employ a conserved His residue as their axial iron ligand, which forms a strong hydrogen bond from the γ N–H to a neighboring aspartate, an interaction known to increase the electron donating capability of His (see section 3.3.4). Introduction of 3-methylhistidine (MeHis) by GCE into the axial position of an engineered ascorbate peroxidase (APX2) disrupts this hydrogen bond and surprisingly leads to a modest increase in the catalytic efficiency of guaiacol (2-methoxyphenol) oxidation (Figure 17A). (381) Remarkably, this structurally conservative substitution also greatly increases the TTN achieved prior to enzyme deactivation (31,300 and 6,200 for APX2 MeHis and APX2, respectively, Figure 17B).

Figure 17

Figure 17. Peroxidases with MeHis proximal ligands. (A) An overlay of the crystal structures of APX2 (PDB: 1OAG (382)) and APX2 MeHis163 (PDB: 5L86 (381)). Key active site residues and the heme are shown as atom-colored sticks with gray and blue carbons, respectively. MeHis is shown with brown carbons. (B) TTN achieved by APX2 and APX2 MeHis. (C) The catalytic efficiency toward guaiacol (2-methoxyphenol) oxidation for Mb variants and horseradish peroxidase (HRP). (D) An overlay of the crystal structures of Mb (PDB: 1A6K (383)) and Mb MeHis93 (PDB: 5OJ9 (384)). The protein backbones are shown as cartoons, and key active site residues and the heme are shown as atom-colored sticks with gray and blue carbons, respectively. MeHis93 is shown with brown carbon atoms.

The oxygen binding protein Mb lacks the proximal pocket aspartate, which is thought to be one contributing factor to the weak peroxidase activity displayed by this protein. (385) Inspired by the improved catalytic activity of APX2, the proximal ligand (His93) in sperm whale Mb was mutated to MeHis, causing a rotation of the imidazole plane (Figure 17D) and a ∼ 4-fold increase in the kcat of guaiacol oxidation (1.4 vs 6.6 s–1). (384) This increased activity was further improved by both rational mutation and directed evolution, to give a final variant with a catalytic efficiency for guaiacol oxidation that is 1,140-fold higher than WT Mb and in the order of natural peroxidases (Figure 17C).
Axial ligand substitutions in Mb have also been shown to improve nonbiological carbene transfer chemistry. Mutation of the axial iron binding His93 to MeHis in an engineered Mb (Mb*) obviates the requirement for a priming reductant and allows for efficient carbene transfer under aerobic conditions, which contrasts with the oxygen sensitivity of the His93-containing Mb*. (386) These properties were attributed to increased electrophilicity of the Fe(III) center, promoting direct attack of the substrate ethyl diazoacetate (EDA) to produce the carbenoid adduct. This intermediate was captured and characterized through X-ray crystallography and found to form an “unusual” bridging Fe(III)–C–N(pyrrole) geometry (Figure 18A), that the authors suggest is a reaction intermediate in Mb* MeHis93 and in the His-containing enzyme. (386)

Figure 18

Figure 18. Biocatalytic cyclopropanations by Mb* MeHis93. (A) The bridged ion carbenoid intermediate observed by X-ray crystallography (PDB: 6F17 (386)). A 2FO–FC map contoured at 1.5 σ is shown around the bridged carbenoid intermediate and the iron atom. (B) The cyclopropanation reaction catalyzed by engineered Mbs. (C) The non-native cofactor and MeHis ligand used to expand the scope of biocatalytic cyclopropanations. (388)

Pott et al. described the introduction of a wider range of noncanonical ligands into Mb, namely 5-thiazoylalanine (5-ThzAla), 4-thiazoylalanine (4-ThzAla) and 3-(3-thienyl)alanine (3-ThiAla), and found both Mb*(MeHis) and Mb*(5-ThzAla) led to improved oxygen tolerance cf. Mb*. (387) This augmented activity extended to abiological N–H insertion reactions, with MeHis93 and 5-ThzAla93. Introduction of 4-ThzAla as a ligand in Mb*, was found to give the most effective catalyst for S-H insertion reaction of EDA and thiophenol, plausibly due to the higher redox potential of the heme increasing the rate of radical recombination of the intermediate thiol radical with the iron-carbenoid.
Noncanonical ligand substitutions have also been used to expand the scope of biocatalytic cyclopropanations. Installing a non-native iron porphyrin cofactor and a noncanonical axial MeHis ligand into the Mb* variant of Mb afforded a selective artificial metalloenzyme that could catalyze cyclopropanation reactions with electron deficient alkenes (Figure 18B,C). (388) Mechanistic studies indicate that this expanded scope is enabled by radical-type carbene transfer reactivity from the combined effect of the non-native cofactor and axial ligand. In a subsequent study, a wider range of noncanonical ligands were explored at position 93, including 4-NH2Phe and 3-(3′-pyridyl)-Ala (3-PyrAla), albeit with reductions in TTN and/or selectivity compared to the MeHis-containing variant. (389)
Replacement of one key His ligand (His264) in the zinc metalloenzyme mannose-6-phosphate isomerase (ManA) by MeHis gave a functional enzyme that permitted growth of an E. coli strain lacking endogenous ManA activity, albeit at a lower growth rate. (390) Directed evolution of ManA His264MeHis led to the generation of an organism whose growth was strictly dependent on the presence of ncAA, providing a possible strategy for the biocontainment of recombinant organisms.
NcAAs have also been introduced to the distal pocket of heme proteins to tune the properties of the heme cofactor and augment catalytic function. Mutation of the distal pocket His64 of Mb to DOPhe led to a 54- and 10-fold increase in catalytic rate toward benzaldehyde oxidation and thioanisole sulfoxidation, respectively. (391) These improvements were proposed to arise from increased stabilization of the ferryl intermediate compound I from additional hydrogen bonding interactions in the His64DOPhe mutant, with the ncAA performing the hydrogen bonding role of the distal pocket His/Arg residues that are conserved in natural peroxidases.

4.3.2. Noncanonical Amino Acids to Tune Enzyme–Substrate Interactions

A range of cAAs and ncAAs were introduced into the Phe124 position of the prodrug activator nitroreductase, and the activity toward the prodrug CB1954 was found to increase 30-fold with 4-NO2Phe. (392) This residue forms π-stacking interactions with the aryl ring of the prodrug substrate, and previous studies interrogating this position had identified its importance to activity. (393) Interestingly, no correlation between ring electron density and catalytic rate was observed. Instead, it was hypothesized that a polarized aromatic ring at position 124 aids in π-stacking with the polarized aromatic substrate.
Mutation of the active site residue Tyr309 in a bacterial PTE to Hco increased the rate of hydrolysis of the insecticide paraoxon (∼8-fold improvement in kcat). (394) This improvement was suggested to arise from electrostatic repulsion between the negatively charged 4-nitrophenolate product and the anionic ncAA side chain to facilitate the rate-limiting product release step.

4.3.3. Noncanonical Amino Acids to Introduce Conformational Changes

Conformational changes resulting from ncAA incorporation have been reported to augment biocatalyst properties by stabilizing structural elements, improving substrate packing or increasing the rate of product release. Pagar et al. described the systematic incorporation of three ncAAs, 4-MePhe, 4-CF3Phe, and BpAla, into each of the positions 33, 86, and 88 in an (R)-selective transaminase, (R)-ATA, (395) which was developed for industrial sitagliptin production. (396) Phe88(4-MePhe), Phe88(4-CF3Phe), and Phe88BpAla variants all showed increased catalytic activity, the most active variant contained BpAla and gave a 15-fold improvement in activity with 1-phenylpropan-1-amine. Introduction of a Phe86Ala mutation into the BpAla containing variant afforded a Phe86Ala/Phe88BpAla double mutant with 30% higher residual activity after 55 °C incubation, which was used to perform kinetic resolution of 1-phenylpropan-1-amine with benzaldehyde as an amine acceptor.
In another study, the kcat of a transketolase engineered to accept aromatic aldehydes was improved by 2-fold with a Tyr385(4-CNPhe) substitution, with reduced substrate inhibition. (397) An alternative substitution, Tyr385(4-NH2Phe), increased the Tm by 5 °C. Structural modeling suggested this was due to the stabilization of an active site helix through an intersubunit hydrogen bond with Gly262.
SPI of 5-FTrp in place of four canonical Trp residues within the M1–1 isoenzyme of rat GST led to a 4-fold increase in kcat for the arylation of glutathione using 1-chloro-2,4-dinitrobenzene as a substrate, a reaction that is thought to be limited by the rate of product release. (398) Introduction of the noncanonical residues were shown by X-ray crystallography to introduce subtle conformational disorder into the active site, plausibly increasing the rate of product release. Interestingly, the kcat was unchanged using phenanthrene 9,10-oxide and 4-phenyl-3-buten-2-one as substrates, where the chemical step is rate limiting.
SPI was also used to replace the four native Phe residues in the restriction endonucleause PvuII with 2-FPhe, 3-FPhe, or 4-FPhe. (399) Of the three fluorinated Phe analogues, only 3-FPhe increased activity, by approximately 2-fold compared with the WT enzyme. The mutated residues were located distal to the catalytic center and DNA binding sites, with the observed activity increases attributed to conformational changes upon ncAA incorporation.
The specific activity of the lipase TTL toward 4-nitrophenyl palmitate hydrolysis increased upon substitution of 11 methionine residues with norleucine (Nle) using SPI. Activity increases were also observed upon substitution of Asp221 with BpAla by GCE. (400) Notably, simultaneous incorporation of both Nle and BpAla into TTL also gave an engineered variant with higher activity than the WT, albeit with slightly reduced activity compared with TTL containing only the Met to Nle substitutions. Similarly, SPI of Nle into 13 Met residue positions in the heme domain of cytohrome P450BM3 increased the rate of peroxygenase catalysis ∼ 2-fold. (401)
In a modified directed evolution experiment, ten ncAAs Phe analogues were introduced at random positions within TEM-1 β-lactamase using SCS. (402) The resulting library were evaluated using a cell viability assay, leading to identification of a Val216(4-acrylamido-phenylalanine) (4-AcrPhe) variant that increased kcat 14-fold with only a slightly increase in KM. Structural analysis of both apo enzymes and cephalexin acyl-enzyme intermediates by X-ray crystallography revealed conformational changes to key active site residues that likely lower the barrier to transacylation in the ncAA containing mutant.

4.3.4. Building Artificial Dimers Through Noncanonical Amino Acid Tethering

Genetic incorporation of the ncAA 4-AzPhe (403) introduces a ‘clickable’ functional group into proteins, which has been used to form artificial heterodimers using SPAAC reactions with bifunctional linkers. (404) Lim et al. used this approach to improve the cofactor shuttling and increase the efficiency of a two-enzyme cascade for D-mannitol production. (405) Introduction of 4-AzPhe into selected sites of formate dehydrogenase and mannitol dehydrogenase (MNDH) created bioorthogonal handles for SPAAC conjugation to either a heterobifunctional linker harboring a tetrazine handle, or an alternative linker with a cyclooctene handle. The tetrazine and cyclooctene moieties could then undergo an IEDDA reaction to give the artificial heterodimer, with the spatial relationship between the two proteins defined by the site of ncAA incorporation and the length of the linkers (Figure 19). Subsequent work described how the relative orientation of the two enzymes’ active sites affected catalysis. (406) With the active sites in closer proximity, D-mannitol production was 60% higher in comparison to the artificial heterodimer with active sites orientated further from one another.

Figure 19

Figure 19. Introduction of 4-AzPhe into selected sites of formate dehydrogenase (FDH) and mannitol dehydrogenase (MNDH) created bioorthogonal handles for SPAAC conjugation to either a heterobifunctional linker harboring a tetrazine handle or an alternative linker with a cyclooctene handle (PDB: 3WR5, (407) 1LJ8 (408)). FDH and MNDH are shown as gray and blue cartoons, respectively. The sites of 4-AzPhe incorporation are shown as red spheres.

4.4. Regulation of Enzyme Activity

NcAAs have been used to install artificial regulatory elements into biocatalysts to enable spatiotemporal control over enzyme activity in vivo or in vitro.

4.4.1. Chemical and Photochemical Decaging

Genetic code expansion has been used to install protected analogues of functional amino acids into enzymes. These caged amino acids offer an effective strategy to inhibit enzyme activity, either through the steric hindrance of ligand binding sites or by masking key catalytic groups, with facile restoration of activity achievable through cleavage of the obstructing group using chemical (409−411) or photochemical (412−424) methods.
Light offers a minimally invasive stimulus that can be used to regulate protein function both in vitro and in vivo. The most frequently employed photocaging moieties are based upon O-nitrobenzyl (O-NB) groups, which can be applied to cap hydroxy, carboxy, thiol or amino groups and are easily cleaved upon irradiation with long-wave UV light. (425,426) Chemical methods including both SPPS and the PTM of amino acid functional groups have proven effective in the photocaging of proteins. However, these techniques are usually restricted to smaller peptides or preferentially target reactive surface residues. One such example involved the photocaging of lysozyme through the reaction of solvent exposed Lys side chains with a photocleavable PEG reagent. (412) PEGylation of the enzyme resulted in a surrounding polymer layer that prevented interaction with substrate and inhibited enzyme activity, which could only be restored upon irradiation-induced cleavage of the bound PEG polymers.
Alternatively, GCE provides a more targeted approach to install new regulatory elements, and has been used to incorporate photocaged analogues of Tyr, Lys, Ser and Cys into enzyme active sites. (413−424) An early example from the Schultz lab reported the replacement of an active site Tyr503 with O-NBTyr in β-galactosidase to afford an inhibited variant that could be activated upon irradiation with 365 nm light to recover 67% of WT activity. (414) This approach was extended to the development of orthogonal translation components for incorporating photocaged tyrosine analogues in eukaryotic hosts, (427) which has enabled photochemical control of enzymes such Cre recombinases, (428) proteases, (422) or nucleases. (417) Zinc-finger nucleases (ZFNs) have been developed for the sequence-specific scission of double-stranded DNA to enable facile gene editing. (429−432) O-NBTyr was installed into ZFN to develop a photochemically activatable restriction enzyme. (417) Rational replacement of active site Tyr471 by O-NBTyr occludes the active site, preventing binding of the DNA substrate, meaning that phosphodiester cleavage was only observed after brief irradiation with 365 nm light. Crucially, this photocaging strategy is entirely compatible with further engineering of the ZFN, allowing for additional modifications to the dimer interface and introduction of additional mutations to further augment nuclease activity. (433) Photocaged analogues of Tyr have also enabled the regulation of polymerase activity. In the widely used Thermus aquaticus (Taq) polymerase, Tyr671 plays an important role in positioning the DNA template and the incoming dNTP. (434,435) Replacement of Tyr671 with O-NBTyr led to an inactivated variant of Taq polymerase, where the bulky O-NB group is believed to occupy dNTP binding cavity (Figure 20). Upon release of the O-NB caging group via short irradiation with 365 nm light, polymerase activity was restored to 71% of that of the WT. (415) This photocaged polymerase enabled development of a UV-inducible hot-start PCR method, a strategy which can reduce nonspecific DNA amplification and increase sensitivity and specificity. (436) Since various monomeric DNA and RNA polymerases also rely on an analogous active site Tyr, (437) recombinant replacement of this residue with O-NBTyr could provide a general approach to photochemical regulation of polymerase activity in these enzymes. To illustrate this broader applicability, Chou et al. engineered a photocaged variant of a bacteriophage T7 RNA polymerase (T7RNAP) with Tyr639 substituted with O-NBTyr. (416) As T7RNAP is orthogonal to all endogenous prokaryotic and eukaryotic RNA polymerases, expression of genes of interest under T7 promoters in bacterial and mammalian cells could be precisely controlled via light activation. Spatiotemporal activation of the polymerase was demonstrated through the engineering of E. coli to produce the photocaged T7RNAP alongside either GFP or luciferase reporter genes. In both cases, light irradiation successfully initiated RNA polymerization leading to downstream gene expression, where fluorescence intensity correlated with the duration of UV exposure. A more recent study targeted Lys631 for genetic replacement with methyl-2-nitropiperonyllysine (MNPLys), producing an additional light activatable variant of T7RNAP. (420)

Figure 20

Figure 20. Introduction of a photocaged ncAA into a DNA polymerase through GCE occludes the active site, preventing the diffusion of nucleotides for extension. Brief irradiation with UV light cleaves the O-NB moiety to reveal the catalytic Tyr and restore polymerase activity. Created with BioRender.com.

CRISPR/Cas9 is a powerful technology for genome editing that is widely used in medicine and biotechnology. The Deiters lab sought to add an optically controllable element to the Cas9 enzyme using GCE. (421) A combination of Ala and ncAA scanning experiments revealed the catalytic importance of Lys886 for Cas9 activity, so this residue was targeted for substitution with a photocaged analogue using SCS. The presence of MNPLys in the protein active site abolished DNA cleavage and nicking activity. Enzyme activity could be fully restored upon photodecaging of Lys886 using 365 nm light. This light-activated Cas9 system was successfully used for endogenous gene silencing, leading to reduced expression of the CD71 gene and 50% reduction of the CD71 transmembrane transferrin receptor on the cell surface.
Photocaged ncAAs have also found application for studying and modulating kinase activity. MNPLys was incorporated in place of the near-universally conserved Lys97 in the ATP binding domain of the MAP kinase MEK1, inhibiting its activity by blocking ATP binding and preventing phosphorylation. (419) Following photodecaging of Lys97, receptor-independent light activation of a designed subnetwork of the Raf/MEK/ERK signaling pathway was achieved in live mammalian cells, enabling study of the kinetics of individual steps in the transduction cascade.
The chemiluminescence activity of Renilla luciferase (RLuc) can also be controlled directly through the genetic replacement of a catalytic Cys residue with the photocaged analogue PCys using a previously engineered MbPylCKRS. (423,127) Following successful incorporation of the caged Cys, the activity of RLuc could be initiated with brief UV irradiation resulting in a > 150-fold increase in chemiluminescence activity over the caged enzyme. Protease activity can also be regulated via photocaging of catalytic Cys nucleophiles. An early study achieved genetic incorporation of O-NBCys in place of the active site Cys163 of capase-3 in yeast. (413) In lysate assays of the mutant enzyme, protease activity was only detectable after photodecaging upon irradiation with UV light. Interestingly, while expression of WT capase is toxic to yeast cells, expression of the photocaged variant was not detrimental to cell growth. This methodology was extended to protease expression in E. coli through the engineering of a PylRS/tRNACUA pair to encode photocaged Cys variants, allowing the development of light-activatable TEV protease variant. (422)

4.4.2. Azobenzene Photoswitches

Azobenzene derivatives have found diverse application in the reversible photocontrol of biomolecules, owing to their efficient light-induced E/Z isomerization which result in structurally disparate isomers (Figure 21). (438−447) Notably, a variety of azobenzene-based photoswitches can now be selectively incorporated into proteins using GCE. (448−456) In one study, incorporation of an azobenzene switch enabled photochemical control over an allosteric activation mechanism in imidazole glycerol phosphate synthase (ImGPS). (453) The HisH glutaminase subunit of this bienzyme complex is allosterically stimulated by binding of a regulatory ribonucleotide to the HisF cyclase subunit. Light-mediated isomerization (356 nm to induce E to Z isomerization and 420 nm for Z to E) of 4-azobenzylphenylalanine (AzoPhe) introduced at position 55 of HisF resulted in reversible 10-fold regulation of HisH activity. In another study, AzoPhe, 4-azo-(2’,6’-difluorobenzyl)phenylalanine (F2AzoPhe), and 4-azo-(2’,6’-difluorobenzyl)-3,5-difluorophenylalanine (F4AzoPhe) were incorporated into firefly luciferase (FLuc) in both prokaryotic and eukaryotic cells. (454) Unsubstituted AzoPhe demonstrates UV light-induced E to Z photoswitching but spontaneously reverts to the more stable E-isomer, posing challenges for long-term control of protein function. In contrast, F4AzoPhe possesses enhanced thermal stability in the Z state. Furthermore, the fluorine substituents lead to a red-shifted absorption spectrum enabling efficient E to Z and Z to E isomerizations to be initiated by visible light, which facilitates longer-term control of enzyme activity in vivo than would be possible with UV irradiation. Computational modeling, guided by local repacking analysis, identified potential allosteric sites in FLuc to aid rational amino acid substitution. Of the seven sites identified, it was found that reversible on/off switching of luciferase activity could be achieved through azobenzene incorporation at Trp417, enabling up to five cycles of photoswitching.

Figure 21

Figure 21. Photoresponsive ncAAs used in the allosteric light regulation of ImGPS. AzoPhe undergoes light induced reversible E/Z isomerizations enabling on–off switching of HisH activity.

4.4.3. Metal Responsive Regulation

In a recent study, Zubi et al. developed a metal-responsive system for protein (in)activation involving the genetic incorporation of two spatially separated bidentate (2,2’-bipyridin-5-yl)alanine (BpyAla) residues to impart conformational control over proteins that do not otherwise exhibit allostery (Figure 22). (457) The serine protease Pyrococcus furiosus (Pfu) prolyl oligopeptidase (POP) was selected as a model enzyme due to the dynamic domain opening/closing it undergoes to allow substrate entry and positioning of the catalytic triad, as revealed by MD simulations. To minimize screening efforts, pairs of positions for BpyAla incorporation were identified through parametrized MD simulations, which highlighted residues that were appropriately positioned to chelate metal ions in the closed conformational state to occlude the active site, but not in the open state. When incubated in excess divalent metal salts such as those of Ni(II), Cu(II), Co(II), and Zn(II), protease activity of selected POP variants was almost entirely inhibited as intended, with near quantitative recovery of enzyme activity achieved upon addition of EDTA for up to 24 on/off cycles. Spectroscopic, and computational investigations collectively suggested that the Bpy pairs engage in reversible metal binding to generate the targeted M(II)(BpyAla)2 complex, resulting in a conformational change which inhibits catalytic turnover. To illustrate the generality of this approach, the study was extended to develop a metal dependent Photinus pyralis luciferase (Pluc), with the best variant showing a 20-fold decreased Vmax in the presence of Ni(II).

Figure 22

Figure 22. Introduction of a pair of BpyAlas into Pfu POP (PDB: 5T88 (458)) enabled inhibition of protease activity when incubated in divalent metal salts. Metal binding of the noncanonical ligands holds POP in a closed inactive conformation, which can be released through chelation of metal ions with EDTA addition, thereby allowing reversible allosteric control of biocatalyst activity. Created with BioRender.com.

5. Designing New Catalytic Mechanisms and Functions

Click to copy section linkSection link copied!

A major objective of modern biocatalysis is to broaden scope of chemical transformations beyond those known in nature, by developing enzymes that operate through new catalytic manifolds. To maximize the breadth of chemistry accessible with designed biocatalysts, it is important to be able to introduce a wide variety of functional groups within enzyme active sites. One way to achieve this is to embed new catalytic elements as ncAA side chains. A major advantage of this approach for enzyme design is its versatility. Having established suitable aaRS/tRNA pairs to encode a functional amino acid of interest, this residue can be positioned at various sites in diverse protein scaffolds. (14,27) Importantly, it has been shown that directed evolution workflows can be adapted to optimize designed enzymes made from an expanded set of amino acid residues. (28,459,460) As a result of rapid advances in the field, a variety of enzymes have now been developed that use ncAAs as key catalytic motifs.

5.1. Metalloenzymes

2,2′-Bipyridine is a widely used bidentate ligand for divalent cations. (461) Xie et al. expressed T4 lysozyme mutants with BpyAla site-specifically incorporated at several positions on the protein surface. (462) Incubation of the modified lysozyme variants with CuCl2 and subsequent MS and spectroscopic analyses confirmed Cu(II) binding, which was dependent on the BpyAla ligand. Subsequently, BpyAla was used to confer DNA cleavage activity on E. coli catabolite activator protein (CAP). (463) Guided by a crystal structure of dimeric CAP bound to DNA, the surface residue Lys26 was selected for substitution to BpyAla due to its proximity to the DNA–protein interface. When incubated with redox active metal ions such as Fe(II) or Cu(II), a reducing agent and the allosteric activator cAMP, CAP Lys26BpyAla promoted site-selective oxidative cleavage of a bound DNA fragment. Cleavage occurred at either side of a specific nucleobase, suggesting that a freely diffusible oxidizing agent generated at the metal center is responsible for DNA cleavage.
The transcription factor Lactoccocal multidrug resistance Regulator (LmrR) has proven to be a versatile scaffold (464) for developing metalloenzymes using BpyAla. Drienovská et al. generated copper metalloproteins for enantioselective vinylogous Friedel–Crafts alkylations by incorporating BpyAla into LmrR at position 89 (Figure 23A). (465) Copper-loaded LmrR Met89BpyAla promoted alkylations of electron-rich indoles with 1-(1-methyl-1H-imidazol-2-yl)but-2-en-1-one (Figure 23B,C) with modest levels of stereocontrol. Reaction selectivity could be further improved through targeted mutagenesis of residues in proximity to the BpyAla ligand. In a later study, a combination of quantum mechanics, docking and MD calculations facilitated the design an LmrR-based metallohydratase. (466) The designs utilized BpyAla-complexed Cu(II) to catalyze enantioselective hydration of α,β-unsaturated 2-acyl pyridines (Figure 23C), achieving up to 64% ee. The same group also demonstrated the binding and stabilization of 2-semiquinone radicals by artificial metalloproteins containing a variety of first row transition metal ions in the LmrR Met89BpyAla template. (467) This ability to stabilize reactive radical species in the binding pocket of a protein could facilitate the development of artificial enzymes for controlling transformations involving radical intermediates.

Figure 23

Figure 23. Catalytic metal-coordinating ncAAs. (A) Crystal structure of dimeric LmrR, with the positions Val15, Met89, and Trp96 in the binding pocket shown with blue carbons (PDB: 3F8B (474)). (B) BpyAla-coordinated Cu(II) complex which activates 1-(1-methyl-1H-imidazol-2-yl)but-2-en-1-one toward nucleophilic attack. (C) Schemes of vinylogous Friedel–Crafts alkylations (top) and α,β-unsaturated 2-acyl pyridine hydrations (bottom) catalyzed by BpyAla-Cu(II) or 3-HqAla-Cu(II) metalloenzymes.

In addition to LmrR, several other protein scaffolds have been elaborated into artificial copper metalloenzymes by incorporating BpyAla. For example, a Tyr123BpyAla variant of QacR proved to be an effective biocatalyst for stereocontrolled vinylogous Friedel–Crafts alkylations of indoles with 1-(1-methyl-1H-imidazol-2-yl)but-2-en-1-one (Figure 23C). (468) Interestingly, this metalloprotein gave the opposite product enantiomers to those formed with LmrR Met89BpyAla. Another study installed BpyAla in a homohexameric acetyltransferase to create a dinuclear copper oxidase. (469) The innate symmetry in the scaffold allowed the distance between the Cu(II) sites to be varied, inducing electronically coupled dinuclear behavior in variants where the sites were proximal and mononuclear behavior in variants where they were distal. Interestingly, the dinuclear copper variants were found to oxidize ascorbate more rapidly than their mononuclear counterparts.
Beyond BpyAla, other metal-binding ncAAs have been used to generate artificial metalloenzymes. The development of an MjTyrRS pair to incorporate 3-HqAla (150) enabled production of LmrR variants with 3-HqAla ligands at Val15 or Met89 (470) (Figure 23A) which can be used to chelate Cu(II), Zn(II) and Rh(II) ions. Slow amide bond hydrolysis activity was observed for Zn(II)-containing variants, which did not occur in the presence of the Zn(NO3)2 salt alone, suggesting that catalysis occurs within the LmrR binding pocket. Cu(II)-loaded variants were found to be competent catalysts for vinylogous Friedel–Crafts alkylations and hydration of α,β-unsaturated carbonyls, albeit with modest selectivities (Figure 23C). More recently, Stein et al. utilized a regioisomer of 3-HqAla, (8-hydroxyquinolin-5-yl)alanine (5-HqAla) to generate a Ru-based metalloenzyme for allylic deamination. (471) Halotag, a self-labeling protein derived from Rhodococcus dehalogenase, (472) was expressed with 5-HqAla incorporated at positions Phe144, Ala145 or Met175 in the hydrophobic binding cleft. After incubation with [Cp*Ru(MeCN)3]PF6, all three variants displayed increased deamination activity toward an O-allyl carbamate-protected coumarin compared to the WT, with the 5-HqAla175 variant performing up to 9 turnovers. In another study, 3-pyridylalanine (3-PyrAla) was successfully used to transform dimeric bovine pancreatic polypeptide (bPP) into an enantioselective biocatalyst for Diels–Alder cycloadditions and Michael reactions. (473) A Tyr7 to 3-PyrAla mutation was proposed to create a bidentate metal binding site at the dimer interface. 3-PyrAla-modified bPP was produced by SPPS and incubated with Cu(H2O)6(NO3)2. The resulting metalloprotein was found to promote Diels–Alder cycloadditons of α,β-unsaturated carbonyls and cyclopentadiene, and Michael additions of dimethylmalonate and α,β-unsaturated 2-acyl imidazoles, with appreciable levels of stereocontrol (up to 83% and 86% ee, respectively).
Artificial metalloenzymes can also be created by using ncAAs to covalently tether metal complexes to proteins. Biorthogonal SPAACs to 4-AzPhe are well-suited for this purpose. (475) Yang et al. reported the incorporation of 4-AzPhe into tHisF, a thermostable α,β-barrel protein, and subsequent SPAAC conjugation of metal complexes derivatized with bicyclo[6.1.0]nonyne (BCN). (476) Tethering of tetracarboxylate dirhodium complexes and Cu- or Mn-terpyridines was demonstrated, with 50–90% conversion to the conjugated species observed. The artificial metalloenzymes were then evaluated for intermolecular cyclopropanation and Si–H insertion activities. Unfortunately, the catalysts gave reduced activity compared with the free metal complex and only minimal selectivity, suggesting the location of the metal complexes near the mouth of the α,β-barrel did not provide a conducive environment for selective catalysis. Greater success was found using POP as a scaffold, chosen for its thermostability and large internal volume suitable for hosting metal complexes. (477) Initial efforts to conjugate POP variants modified with 4-AzPhe at various positions in the active site with a BCN-derivatized dirhodium paddlewheel complex (Figure 24A) failed, likely due to the enzyme adopting a closed conformation which shields the active site from solvent. Mutation of four residues lining the pore of the β-barrel domain to alanine overcame these issues (Figure 24B), enabling rapid conjugation to form the metalated enzyme. Catalysis of styrene cyclopropanation with a diazo ester (Figure 24C) gave a single diastereomer with 19% conversion and 11% ee, which could be further improved to 38% ee upon optimization of reaction conditions. Additional improvements were achieved with the introduction of an active site His to coordinate the proximal rhodium center and phenylalanine mutations around the distal rhodium, with the best variant achieving 74% conversion and 92% ee. In a subsequent study, a directed evolution platform was established to allow more extensive engineering of artificial metalloenzymes generated using SPAAC. (458) Starting from the aforementioned quadruple alanine POP mutant, three rounds of mutagenesis and screening generated the improved variant 3-VRVH, which contained 12 mutations and achieved 92% ee for 4-methyoxystyrene cyclopropanation, and a significantly higher reaction rate than the optimal variant from the previous study. Directed evolution was also used to optimize the performance of artificial POP metalloenzymes containing dirhodium paddlewheel complexes for efficient cross-coupling of diazoesters. (478) The most highly evolved metalloenzyme, 5-G, was able to perform an impressive 40,000 turnovers and delivered the cross-coupled products with high levels of stereocontrol (14.9:1 E/Z selectivity). 5-G was subsequently integrated into a biocatalytic cascade with an alkene reductase and glucose dehydrogenase for NADPH recycling (Figure 24C), leading to the production of derivatized succinate products with up to 61% conversion and > 99% ee.

Figure 24

Figure 24. 4-AzPhe-anchored metalloenzymes. (A) BCN-Derivatised dirhodium complex. OAc = acetate anion. (B) Crystal structure of POP, with positions of 4-AzPhe incorporation (orange spheres) and pore-opening alanine mutations (blue spheres) shown (PDB: 5T88 (479)). (C) Schemes of styrene cyclopropanations (top) and the diazo cross-coupling cascade (bottom) catalyzed by POP variants containing 4-AzPhe-tethered dirhodium complexes.

5.2. Nucleophilic Catalysis

Nucleophilic catalysis is a versatile strategy used in chemistry and biology to accelerate diverse transformations. For example, natural hydrolases or acyltransferases often feature an activated Ser or Cys nucleophile embedded within catalytic diads or triads. (480−482) Efforts to recapitulate these mechanisms in designed enzymes commonly results in proteins where catalysis stalls due to the formation of stable acyl-enzyme intermediates that are resistant to hydrolysis. (483−485) To overcome these limitations, our lab used GCE to reengineer a computationally designed protein BH32 into an efficient hydrolase. (486) BH32 was originally designed to catalyze the Morita-Baylis-Hillman (MBH) reaction using a His23 nucleophile, (487) but was also found to possess promiscuous ester hydrolase activity. (470) Similar to previously designed hydrolases, BH32 displayed a biphasic reaction profile consistent with rapid acylation of His23 generating a stable acyl-imidazole intermediate that is resistant to hydrolysis. To resolve this bottleneck in catalysis, His23 was replaced by a noncanonical MeHis, which lead to the formation of more reactive acyl-imidazolium intermediates (Figure 25A). Hydrolase activity was subsequently enhanced by directed evolution, affording the enzyme OE1.3 with six mutations (Figure 25B) that is ca. 9,000-fold more active than free MeHis in solution, along with an enantioselective hydrolase OE1.4 that contains an additional three mutations (cf. OE1.3). MeHis has also been shown to be a valuable catalytic nucleophile for more complex chemical conversions. Installation of MeHis into an engineered BH32 template followed by extensive evolutionary optimization afforded a highly efficient and enantioselective enzyme (BHMeHis1.8) for Morita-Baylis-Hillman reactions, which are valuable C–C bond forming processes for which there are no natural enzymes known. (488) BHMeHis1.8 is more than an order of magnitude more active than our earlier BH32.14 enzyme which contains a His23 nucleophile, (489) and can also promote challenging MBH conversions of electron-rich aromatic aldehydes that were inaccessible with BH32.14. Interestingly, introduction of MeHis led to a dramatically altered mechanistic outcome following evolution, where a key catalytic Arg124 found in BH32.14 was abandoned in favor of a Glu26 residue that mediates a rate-limiting proton transfer step (Figure 25C).

Figure 25

Figure 25. Nucleophilic catalysis utilizing MeHis. (A) Scheme of ester hydrolysis, showing the reactive covalent intermediate formed between the substrate and MeHis23 (orange). (B) Structure of OE1.3, with MeHis23 (orange carbons) and sites of mutations installed during evolution (blue spheres) shown (PDB: 6Q7Q (486)). (C) Scheme highlighting the proton transfer role of Glu26 (gray) in the evolved MBHase BHMeHis1.8. Intermediates 2 (left) and 3 (right) are shown, covalently bound to MeHis23 (orange).

Aniline nucleophiles are known to catalyze hydrazone and oxime formations. (490,491) LmrR was engineered to efficiently catalyze hydrazone formations by embedding an aniline nucleophile. (492) A two-step protocol involving initial installation of 4-AzPhe by GCE followed by chemical reduction of the aromatic azide was used. The resulting enzyme, LmrR_Val15(4-NH2Phe), achieved 72% conversion for a hydrazone formation between 4-methoxybenzaldehyde and a benzoxadiazole, and was also found to catalyze the analogous oxime formation (Figure 26A), with a kcat/KM 37-fold higher than WT LmrR. Hydrazone formation activity was subsequently optimized through directed evolution, (460) affording triple and quadruple mutants that displayed 57- and 74-fold improved catalytic efficiency respectively compared to the parent enzyme. Mutation of the 4-NH2Phe nucleophile to an isosteric Tyr led to dramatic activity reductions in both variants, highlighting the critical role played by the ncAA in accelerating hydrazone formation. The evolved hydrazone-forming enzymes have subsequently been used in combination with alcohol oxidases and carboxylic acid reductases for in vivo biocatalytic cascades. (493) The reaction scope of 4-NH2Phe-containing LmrR variants has been extended to include enantioselective Friedel–Crafts alkylations of indoles (Figure 26B) (459,494) and enantioselective Michael additions of enals and 2-acylimidazoles. (495,496) In the latter case, catalysis is dependent upon the synergistic action of the 4-NH2Phe nucleophile and a Cu(II)-phenanthroline complex that is sandwiched between two tryptophan residues within the hydrophobic binding pocket, with up to 99% ee observed.

Figure 26

Figure 26. Nucleophilic catalysis utilizing 4-NH2Phe. (A) Scheme of hydrazone (X = N) and oxime (X = O) formations catalyzed by 4-NH2Phe (orange) incorporated into LmrR, with the covalent adduct formed by the carbonyl substrate and 4-NH2Phe15 shown. (B) Scheme of vinylogous Friedel–Crafts alkylations catalyzed by LmrR_V15_4-NH2Phe_RGN, with the activated imine intermediate formed between 4-NH2Phe15 (orange) and the aldehyde substrate shown. At the end of the reaction time NaBH4 is added to reduce the enzymatic product to the corresponding alcohol (right).

More recent studies have disclosed other catalytic ncAAs. Gran-Scheuch et al. reported the synthesis and incorporation of several ncAAs featuring secondary amine motifs using the MbPylRS system. (497) LmrR containing a pyrrolidine-inspired ncAA at position 15 was found to catalyze Michael addition of nitromethane to cinnamaldehyde with moderate conversions and up to 38% ee. Incubation of the enzyme with cinnamaldehyde and NaBH3CN gave rise to an MS peak shift consistent with formation of a reduced Schiff-base adduct, implicating ncAA-mediated iminium ion activation of the aldehyde. Another study reported the use of 4-boronophenylalanine (4-BoPhe) installed in engineered LmrR variants to catalyze condensation of α-hydroxyketones with hydroxylamine to form enantioenriched oximes in a kinetic resolution process. (498) MS and 11B NMR studies provided evidence for the catalytic role of 4-BoPhe, which was proposed to form transient boronate adducts with vicinal diol reaction intermediates in a stereoselective manner.

5.3. Photocatalysis

Photons provide a convenient and tuneable source of energy to selectively access reactive excited state intermediates under mild reaction conditions. This use of light energy opens up new modes of reactivity that are challenging to access in the ground state, resulting in the development of diverse synthetic methods for constructing carbon-carbon and carbon-heteroatom bonds. (499,500) Light-driven reactions can be broadly divided into photoinduced electron transfer (photoredox) processes and triplet energy transfer processes. There are a handful of natural photoenzymes (501−503) and several engineered photoenzymes (504−506) that operate via photoredox mechanisms. In contrast, there are no natural enzymes known to mediate stereocontrolled transformations through energy transfer mechanisms. To address this limitation, Trimble et al. and Sun et al. independently developed enantioselective enzymes for thermally forbidden [2 + 2] cycloadditions. In the former study, a BpAla triplet sensitizer was incorporated into the active site of a previously designed Diels–Alderase, DA_20_00, (507) giving rise to a first generation photocatalyst EnT1.0 that could mediate intramolecular [2 + 2] cycloadditions of quinolone substrates (Figure 27A) with modest regio- and enantioselectivity upon irradiation with 365 nm light. (28) Introduction of five additional mutations via directed evolution afforded an optimized photoenzyme EnT1.3 with substantially improved activity, regioselectivity and enantioselectivity (>99% ee). This energy transfer photoenzyme can operate efficiently at ambient temperatures and in the presence of oxygen, and can also mediate bimolecular [2 + 2] cycloadditions with high levels of stereocontrol. A product-bound crystal structure of EnT1.3 reveals that the ligand is sandwiched between the BpAla side chain and an active site His (Figure 27B), an arrangement conducive to efficient energy transfer between the photosensitizer and substrate. Sun et al. used a similar approach to develop enantioselective photoenzymes for intramolecular cycloadditions of N-substituted indoles using LmrR as a protein scaffold. (508) In this case, a fluorinated analogue of BpAla (3′-FBpAla) led to improved conversions and selectivities for a number of substrates compared with the parent sensitizer.

Figure 27

Figure 27. [2 + 2] Photocycloadditions catalyzed by BpAla. (A) Schemes of intramolecular [2 + 2] photocycloadditions of derivatized quinolones (top) and indoles (bottom). X = O or C, n = 1 or 2. (B) Crystal structure of EnT1.3 with product (green carbons) bound between BpAla (orange carbons), Trp244, and His287 (blue carbons) (PDB: 7ZP7 (28)).

A large number of chemical transformations can be driven by photoinduced electron transfer (PET) processes. (509,510) Incorporation of ncAAs into fluorescent proteins has delivered photoredox catalysts by tuning the absorption profiles and redox properties of the chromophore. Liu et al. generated a miniature photocatalytic CO2-reducing enzyme by modifying superfolder yellow fluorescent protein (sfYFP), (506) which features a chromophore generated by autocatalytic cyclization and oxidation of residues Ser65, Tyr66 and Gly67 (511) (Figure 28A, top). Substitution of Tyr66 with BpAla (Figure 28A, bottom) in a His148Glu Phe203Asp mutant of sfYFP generated the photosensitizer protein PSP2. (506) Photochemical reduction of PSP2 with sacrificial reductants produced super-reducing radicals (PSP2•) which were sufficiently potent to drive CO2 reduction by a nickel-terpyridine complex ligated to PSP2 through a Cys residue introduced at position 95 (Figure 28B). Introduction of proton-donating Tyr residues around the nickel complex further improved activity, with the resulting enzyme PSP2T2 exhibiting a CO2/CO conversion quantum efficiency of 2.6%. PSP2T2 was also found to catalyze light-driven dehalogenations of simple aryl halides to phenolic products with up to 98% conversion (Figure 28C). (512)

Figure 28

Figure 28. Metal-dependent ncAA-incorporating photoenzymes. (A) Chromophore autocatalytically generated in sfYFP and in PSP2, which incorporates BpAla (orange side chain) at position 66. (B) Structure of PSP2, with a chromophore shown (backbone indicated with gray carbons, BpAla side chain with orange carbons). The Cys95 site of nickel–terpyridine complex ligation is shown in dark gray (PDB: 5YR3 (506)). (C) Scheme of dehalogenation reactions catalyzed by BpAla-incorporating PSP2T2 or by BpyAla-incorporating Mb. X = Cl, Br, or I. (D) Structure of Mb incorporating BpyAla (orange carbons) and with an iridium photocatalyst (green carbons) ligated to Cys45 (gray carbons) (PDB: 7YLK (516)).

Photoenzymes can also be generated via biorthogonal conjugation of organic or metal photocatalysts to 4-AzPhe incorporated into proteins. Gu et al. anchored a BCN-derivatized 9-mesityl-10-methylacridinium cofactor into the active site of POP. (513) Upon irradiation with 450 nm light, the modified enzyme catalyzed the conversion of thioanisoles to the corresponding sulfoxides, albeit with somewhat lower conversions than the free cofactor. More success was found by anchoring a BCN-derivatized Ru(Bpy)32+ complex into POP. (514) The conjugated enzyme catalyzed photoreductive cyclization of a dienone substrate with 72% conversion, compared to 36% with the free ruthenium complex. Similarly, [2 + 2] photocycloaddition of 4-methoxystyrene and cinnamoyl imidazole proceeded with 81% conversion using the conjugated enzyme but only 25% with the free complex. The same reaction could also be catalyzed using polypyridyl iridium complexes anchored within POP. (515) Artificial photoenzymes have also been developed by incorporating multiple metal complexes within a protein scaffold. Lee and Song developed an artificial dehalogenase in Mb by introducing a genetically encoded BpyAla to chelate a nickel cofactor and an iridium photocatalyst ligated to a Cys at position 45 (Figure 28D). (516) The resulting enzyme catalyzed a mixture of light-driven hydrolytic and reductive dehalogenation of 4′-iodoacetophenone (Figure 28C), achieving 86% conversion to the phenolic product and ca. 10:1 selectivity over the reduced side product. Reaction selectivity could be further improved through judicious placement of the BpyAla ligated nickel cofactor. In contrast, using the free metal complexes in solution gave substantially lower conversion and chemoselectivity.

6. Conclusions and Outlook

Click to copy section linkSection link copied!

The ability to introduce new functional elements into proteins as ncAA side chains has opened a wealth of opportunities in the field of biocatalysis. In this review, we have highlighted how an expanded alphabet of amino acids can be used to study enzyme mechanisms, (255,271,392) augment biocatalyst properties, (351,362,370,373,339,517) and create enzymes with new catalytic functions. (450,470) Moving forward, there are opportunities for new innovations to maximize the impact of ncAAs in biocatalysis research.
First, continued expansion of the genetic code to include a greater repertoire of functional amino acids will be key to unlocking new modes of catalysis within protein active sites. For example, the recent encoding of phosphine-containing ncAAs holds great promise for metalloenzyme design and engineering. (518) Similarly, a recent article highlights the potential of genetically encoded boronic acids as catalytic motifs. (498) The introduction of a wider range of ncAAs harboring photoresponsive elements should also allow the development of selective biocatalysts for diverse photochemical reactions. Such photoresponsive ncAAs can also be interfaced with modern structural biology techniques, such as X-ray free electron laser crystallography, to provide new insights into conformational changes that take place during enzyme catalysis. (159) At present, the application of genetic code reprograming in biocatalysis has largely focused on the introduction of one or multiple copies of a single ncAA. However, with advances in synthetic genomics and the discovery of mutually orthogonal aaRS-tRNA pairs it is now possible to selectively introduce multiple ncAAs into proteins. (116,117,519−523) In principle, these advances will greatly expand the range of active site arrangements accessible in proteins, which should allow the development of increasingly sophisticated catalytic mechanisms.
Second, to fully capitalize on our ability to embed new functional motifs into proteins, we require accelerated enzyme engineering pipelines to deliver ncAA-containing biocatalysts with the properties required for target applications. For example, the design and evolution of new enzymes that use ncAAs as key functional elements typically takes several years with existing workflows. One approach to speed up biocatalyst development is to integrate ncAA mutagenesis with ultrahigh throughput screening methods (e.g., fluorescence-activated droplet sorting (524)), which allow greater exploration of protein sequence space to accelerate directed evolution campaigns. To complement experimental engineering techniques, the latest deep-learning methods for protein design (6,7,525−527) and structure prediction (528) also hold enormous promise. These methods have allowed the accurate design of proteins that bind metal ions, small molecules and peptides, and have recently been used to design new and improved enzymes. (8) A crucial next step is to integrate ncAAs into these deep-learning frameworks, which should enable the rapid design of ncAA-containing enzymes with a high degree of accuracy.
Finally, in the coming years it is essential to transition ncAA-containing enzymes from academic laboratories into large-scale biocatalytic applications. Current barriers to translation include the limited efficiency of orthogonal translation components and suboptimal performance of common production strains for genetic code reprogramming applications. These limitations lead to reduced titers of ncAA-containing proteins and the requirement for large excesses of ncAAs supplemented to the culture media, which results in prohibitively high production costs. To overcome these challenges, more efficient orthogonal translation components are required that can operate effectively at low ncAA concentrations. The development of such systems should be achievable through multiple rounds of laboratory evolution to progressively improve performance, or through advanced techniques such as multiplex automated genome engineering, (122) phage-assisted continuous evolution, (120) tRNA display (119) and computational approaches. (529,530) A recent study has also shown that exploration of a wider range of PylRS homologues can dramatically improve the efficiency of encoding catalytically important ncAAs. (531) To further improve protein titers, engineered or synthetic production strains can be used that have been specifically tailored for efficient ncAA incorporation. (532−536) Another attractive option to reduce the costs of producing ncAA-containing proteins is developing engineered hosts that contain the necessary biosynthetic machinery to produce target ncAAs in vivo, (330,537−539) thus avoiding the need to supply expensive ncAAs to the culture media.
For the reasons outlined above, we are optimistic that genetic code reprogramming methodologies will become increasingly important tools in enzymology and biocatalysis in the coming years. By overcoming the constraints of the genetic code, enzyme designers and engineers can now begin to tackle a vast array of chemical transformations that were previously thought inaccessible to biocatalysis.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Zachary Birch-Price - Manchester Institute of Biotechnology, School of Chemistry, University of Manchester, Manchester M1 7DN, U.K.Orcidhttps://orcid.org/0000-0002-2024-3005
    • Florence J. Hardy - Manchester Institute of Biotechnology, School of Chemistry, University of Manchester, Manchester M1 7DN, U.K.Orcidhttps://orcid.org/0000-0003-0671-0209
    • Thomas M. Lister - Manchester Institute of Biotechnology, School of Chemistry, University of Manchester, Manchester M1 7DN, U.K.
    • Anna R. Kohn - Manchester Institute of Biotechnology, School of Chemistry, University of Manchester, Manchester M1 7DN, U.K.
  • Author Contributions

    Z.B.-P., F.J.H., T.M.L., and A.R.K.: writing─original draft, reviewing, and editing. A.P.G.: writing─reviewing and editing, supervision. CRediT: Zachary Birch-Price writing-original draft, writing-review & editing; Florence J. Hardy writing-original draft, writing-review & editing; Thomas M. Lister writing-original draft, writing-review & editing; Anna R. Kohn writing-original draft, writing-review & editing; Anthony P. Green supervision, writing-original draft, writing-review & editing.

    Author Contributions

    Z.B.-P. and F.J.H. contributed equally. CRediT: Zachary Birch-Price writing-original draft, writing-review & editing; Florence J. Hardy writing-original draft, writing-review & editing; Thomas M. Lister writing-original draft, writing-review & editing; Anna R. Kohn writing-original draft, writing-review & editing; Anthony P. Green supervision, writing-original draft, writing-review & editing.

    Author Contributions

    T.M.L. and A.R.K. contributed equally. CRediT: Zachary Birch-Price writing-original draft, writing-review & editing; Florence J. Hardy writing-original draft, writing-review & editing; Thomas M. Lister writing-original draft, writing-review & editing; Anna R. Kohn writing-original draft, writing-review & editing; Anthony P. Green supervision, writing-original draft, writing-review & editing.

  • Notes
    The authors declare no competing financial interest.

Biographies

Click to copy section linkSection link copied!

Zachary Birch-Price received his MSci degree in Natural Sciences (Biochemistry) from the University of Cambridge in 2020, working under the supervision of Prof. Ben Luisi. He was then awarded a BBSRC DTP studentship and moved to the University of Manchester, where he is currently studying for his doctorate under the supervision of Prof. Anthony Green, focusing on enzyme engineering and ncAAs.

Dr. Florence J. Hardy received her MChem degree in 2016 under the supervision of Prof. Chris Schofield at the University of Oxford. She moved to the University of Manchester to work with Prof. Anthony Green where she completed a Ph.D. in 2022, on engineering metalloenzymes with ncAAs. Currently, Florence is a postdoctoral researcher at the University of Manchester working on computational enzyme design.

Thomas M. Lister received his MChem degree from the University of Bath in 2020. During this time, he completed a year-long industrial placement at GSK working in medicinal chemistry and a final year project under the supervision of Dr. Alexander J. Cresswell. He is currently a Ph.D. student on the iCAT CDT at the University of Manchester working under the supervision of Profs. Igor Larrosa and Anthony P. Green.

Anna R. Kohn received her MChem degree from the University of Manchester in 2022, completing her final year project under the supervision of Prof. Anthony Green. She has since continued in the Green group as a Ph.D. student, where she is currently engineering photoenzymes containing ncAAs.

Following his Ph.D. in synthetic organic chemistry under the supervision of E. J. Thomas, Anthony carried out postdoctoral research with N. J. Turner and S. L. Flitsch based in the Manchester Institute of Biotechnology and subsequently with D. Hilvert at ETH Zurich. Anthony started his independent research career in 2016 based in the Manchester Institute of Biotechnology at the University of Manchester, where he is a professor of organic and biological chemistry. His research interests lie in the design, evolution, and characterization of enzymes with a new function.

Acknowledgments

Click to copy section linkSection link copied!

We acknowledge the European Research Council (ERC Starting Grant no. 757991 to A.P.G.), the Biotechnology and Biological Sciences Research Council (David Phillips Fellowship BB/M027023/1 to A.P.G. and grants BB/W014483/1 and BB/X000974/1), and the Human Frontier Science Program research grant (RGP0004/2022). Z.B.-P. was supported by a BBSRC Doctoral Training Partnership (BB/T008725/1). F.J.H. was supported by the EPSRC Doctoral Prize Fellowship (EP/W524347/1). T.M.L. was supported by an integrated catalysis Doctoral Training Program (EP/023755/1). A.R.K. was supported by the Future Biomanufacturing Hub (EP/S01778X/1) on an EPSRC Industrial CASE PhD studentship with GSK.

Abbreviations

Click to copy section linkSection link copied!

aaRS

Aminoacyl tRNA synthetase

AFM

Atomic force microscopy

AKR

Aldehyde ketone reductase

APX2

Ascorbate peroxidase, engineered

AsSec

Aeromonas salmonicida selenocysteine synthase

AzoR

Azoreductase

BCN

Bicyclo[6.1.0]nonyne

BODIPY

Boron-dipyrromethene

bPP

Bovine pancreatic polypeptide

cAA

Canonical amino acid

CalB

Candida antartica lipase B

CAP

Catabolite activator protein

CAT

Chloramphenicol acetyltransferase

CcP

Cytochrome c peroxidase

CDO

Cysteine dioxygenase

CFE

Cell-free expression

ChPylRS

Chimeric pyrrolysyl tRNA synthetase

CuAAC

Copper-catalyzed azide–alkyne coupling

CYP

Cytochrome P450

DADP

Diacetyl deuteroporphyrin

ddNTP

2′,3′-dideoxynucleotide triphosphate

DEER

Double electron–electron resonance

DET

Direct electron transfer

DKR

Diketoreductase

DQF-COSY

Double quantum filtered correlation spectroscopy

dr

Diastereomeric ratio

EcLeuRS

Escherichia coli leucyl tRNA synthetase

EDA

Ethyl diazoacetate

EDTA

Ethylenediaminetetraacetic acid

ee

Enantiomeric excess

ELP

Elastin-like polypeptide

EPL

Expressed protein ligation

EPR

Electron paramagnetic resonance

FDH

Formate dehydrogenase

FLuc

Firefly luciferase

GCE

Genetic code expansion

GFP

Green fluorescent protein

GST

Glutathione S-transferase

HAT

Hydrogen atom transfer

HCO

Heme copper oxidase

HRP

Horseradish peroxidase

HSQC

Heteronuclear single quantum coherence

IEDDA

Inverse-electron-demand Diels–Alder

ImGPS

Imidazole glycerol phosphate synthase

IR

Infrared spectroscopy

KlenTaq

DNA polymerase I from Thermus aquaticus

KSI

Ketosteroid isomerase

LaL

Lysozyme from bacteriophage λ

LmrR

Lactoccocal multidrug resistance regulator

ManA

Mannose-6-phosphate isomerase

Mb

Myoglobin

MBH

Morita–Baylis–Hillman

MbPylCKRS

Methanosarcina barkeri pyrrolysyl photocaged lysine tRNA synthetase

MbPylRS

Methanosarcina barkeri pyrrolysyl tRNA synthetase

MD

Molecular dynamics

MLDH

Malate dehydrogenase

MNDH

Mannitol dehydrogenase

mDHFR

Murine dihydrofolate reductase

MjTyrRS

Methanocaldococcus jannaschii tyrosyl tRNA synthetase

MmPylRS

Methanosarcina mazeii pyrrolysyl tRNA synthetase

MmSepRS

Methanococcus maripaludis phosphoseryl-tRNA synthetase

MS

Mass spectrometry

MTG

Microbial transglutaminase

mtRNAP

Mitochondrial RNA polymerase

ncAA

Noncanonical amino acid

NCL

Native chemical ligation

NiSOD

Nickel-dependent superoxide dismutase

NMR

Nuclear magnetic resonance

PCR

Polymerase chain reaction

pdCpA

5′-phospho-2′-deoxyribocytidylriboadenosine

PEG

Polyethylene glycol

PET

Photoinduced electron transfer

PLA

Phospholipase A2

PLuc

Photinus pyralis luciferase

POP

Prolyl oligopeptidase

POR

Protochlorophyllide oxidoreductase

PRMT1

Protein arginine methyltransferase 1

PTE

Phosphotriesterase

PTM

Post-translational modification

RLuc

Renilla luciferase

RNase

Ribonuclease

RNR

Ribonucleotide reductase

ROS

Reactive oxygen species

SAD

Single-wavelength anomalous diffraction

SAM

S-adenosylmethionine

SCS

Stop codon suppression

sfYFP

Superfolder yellow fluorescent protein

SHC

Squalene-hopene cyclase

SPAAC

Strain-promoted azide–alkyne coupling

SPI

Selective pressure incorporation

SPPS

Solid-phase peptide synthesis

sTCO

Strained trans-cyclooctene

T7RNAP

T7 RNA polymerase

tHisF

Thermotoga maritima synthase subunit of ImGPS

TMS

Trimethylsilyl

TOCSY

Total correlation spectroscopy

TR

Thioredoxin

TrpOx

Tryptophan oxidase

tsCA

Thermostable carbonic anhydrase II

TTL

Thermoanaerobacter thermohydrosulfuricus lipase

TTN

Total turnover number

TvNiR

Thioalkalivibrio nitratireducens cytochrome c nitrite reductase

VHR

Vaccinia H1-related

VSE

Vibrational Stark effect

WT

Wild type

XANES

X-ray absorption near edge structure

XFEL

X-ray free-electron laser

ZFN

Zinc finger nuclease

References

Click to copy section linkSection link copied!

This article references 539 other publications.

  1. 1
    Punekar, N. S. ENZYMES: Catalysis, Kinetics and Mechanisms; Springer Nature Singapore, 2018. DOI: 10.1007/978-981-13-0785-0
  2. 2
    Richter, M. Functional Diversity of Organic Molecule Enzyme Cofactors. Nat. Prod. Rep. 2013, 30 (10), 13241345,  DOI: 10.1039/c3np70045c
  3. 3
    Wagner, C. R.; Benkovic, S. J. Site Directed Mutagenesis: a Tool for Enzyme Mechanism Dissection. Trends Biotechnol. 1990, 8, 263270,  DOI: 10.1016/0167-7799(90)90189-5
  4. 4
    Arnold, F. H. Directed Evolution: Bringing New Chemistry to Life. Angew. Chem. Int. Ed. 2018, 57 (16), 41434148,  DOI: 10.1002/anie.201708408
  5. 5
    Turner, N. J. Directed Evolution Drives the Next Generation of Biocatalysts. Nat. Chem. Biol. 2009, 5, 567573,  DOI: 10.1038/nchembio.203
  6. 6
    Dauparas, J.; Anishchenko, I.; Bennett, N.; Bai, H.; Ragotte, R. J.; Milles, L. F.; Wicky, B. I. M.; Courbet, A.; de Haas, R. J.; Bethel, N. Robust Deep Learning-Based Protein Sequence Design Using ProteinMPNN. Science 2022, 378 (6615), 4956,  DOI: 10.1126/science.add2187
  7. 7
    Watson, J. L.; Juergens, D.; Bennett, N. R.; Trippe, B. L.; Yim, J.; Eisenach, H. E.; Ahern, W.; Borst, A. J.; Ragotte, R. J.; Milles, L. F. De Novo Design of Protein Structure and Function with RFdiffusion. Nature 2023, 620, 10891100,  DOI: 10.1038/s41586-023-06415-8
  8. 8
    Sumida, K. H.; Núñez-Franco, R.; Kalvet, I.; Pellock, S. J.; Wicky, B. I. M.; Milles, L. F.; Dauparas, J.; Wang, J.; Kipnis, Y.; Jameson, N. Improving Protein Expression, Stability, and Function with ProteinMPNN. J. Am. Chem. Soc. 2024, 146 (3), 20542061,  DOI: 10.1021/jacs.3c10941
  9. 9
    Gargiulo, S.; Soumillion, P. Directed Evolution for Enzyme Development in Biocatalysis. Curr. Opin. Chem. Biol. 2021, 61, 107113,  DOI: 10.1016/j.cbpa.2020.11.006
  10. 10
    Planas-Iglesias, J.; Marques, S. M.; Pinto, G. P.; Musil, M.; Stourac, J.; Damborsky, J.; Bednar, D. Computational Design of Enzymes for Biotechnological Applications. Biotechnol. Adv. 2021, 47, 107696,  DOI: 10.1016/j.biotechadv.2021.107696
  11. 11
    Świderek, K.; Tuñón, I.; Moliner, V.; Bertran, J. Computational Strategies for the Design of New Enzymatic Functions. Arch. Biochem. Biophys. 2015, 582, 6879,  DOI: 10.1016/j.abb.2015.03.013
  12. 12
    Renata, H.; Wang, Z. J.; Arnold, F. H. Expanding the Enzyme Universe: Accessing Non-Natural Reactions by Mechanism-Guided Directed Evolution. Angew. Chem. Int. Ed. 2015, 54 (11), 33513367,  DOI: 10.1002/anie.201409470
  13. 13
    Pagar, A. D.; Patil, M. D.; Flood, D. T.; Yoo, T. H.; Dawson, P. E.; Yun, H. Recent Advances in Biocatalysis with Chemical Modification and Expanded Amino Acid Alphabet. Chem. Rev. 2021, 121 (10), 61736245,  DOI: 10.1021/acs.chemrev.0c01201
  14. 14
    Zhao, J.; Burke, A. J.; Green, A. P. Enzymes with Noncanonical Amino Acids. Curr. Opin. Chem. Biol. 2020, 55, 136144,  DOI: 10.1016/j.cbpa.2020.01.006
  15. 15
    Huang, Y.; Liu, T. Therapeutic Applications of Genetic Code Expansion. Synth. Syst. Biotechnol. 2018, 3 (3), 150158,  DOI: 10.1016/j.synbio.2018.09.003
  16. 16
    Icking, L.-S.; Riedlberger, A. M.; Krause, F.; Widder, J.; Frederiksen; Anne, S.; Stockert, F.; Spädt, M.; Edel, N.; Armbruster, D.; Forlani, G. iNClusive: a Database Collecting Useful Information on Non-Canonical Amino Acids and Their Incorporation into Proteins for Easier Genetic Code Expansion Implementation. Nucleic Acids Res. 2024, 52, D476,  DOI: 10.1093/nar/gkad1090
  17. 17
    Chatterjee, A.; Guo, J.; Lee, H. S.; Schultz, P. G. A Genetically Encoded Fluorescent Probe in Mammalian Cells. J. Am. Chem. Soc. 2013, 135 (34), 1254012543,  DOI: 10.1021/ja4059553
  18. 18
    Evans, H. T.; Benetatos, J.; van Roijen, M.; Bodea, L. G.; Götz, J. Decreased Synthesis of Ribosomal Proteins in Tauopathy Revealed by Non-Canonical Amino Acid Labelling. EMBO J. 2019, 38 (13), e101174  DOI: 10.15252/embj.2018101174
  19. 19
    Schultz, K. C.; Supekova, L.; Ryu, Y.; Xie, J.; Perera, R.; Schultz, P. G. A Genetically Encoded Infrared Probe. J. Am. Chem. Soc. 2006, 128 (43), 1398413985,  DOI: 10.1021/ja0636690
  20. 20
    Goettig, P.; Koch, N. G.; Budisa, N. Non-Canonical Amino Acids in Analyses of Protease Structure and Function. Int. J. Mol. Sci. 2023, 24 (18), 14035,  DOI: 10.3390/ijms241814035
  21. 21
    Tinzl, M.; Hilvert, D. Trapping Transient Protein Species by Genetic Code Expansion. ChemBioChem 2021, 22 (1), 9299,  DOI: 10.1002/cbic.202000523
  22. 22
    Birch-Price, Z.; Taylor, C. J.; Ortmayer, M.; Green, A. P. Engineering Enzyme Activity Using an Expanded Amino Acid Alphabet. Protein Eng. Des. Sel. 2023, 36, gzac013,  DOI: 10.1093/protein/gzac013
  23. 23
    Giger, S.; Buller, R. Advances in Noncanonical Amino Acid Incorporation for Enzyme Engineering Applications. CHIMIA 2023, 77 (6), 395402,  DOI: 10.2533/chimia.2023.395
  24. 24
    Hayashi, T.; Hilvert, D.; Green, A. P. Engineered Metalloenzymes with Non-Canonical Coordination Environments. Chem. Eur. J. 2018, 24 (46), 1182111830,  DOI: 10.1002/chem.201800975
  25. 25
    Lugtenburg, T.; Gran-Scheuch, A.; Drienovská, I. Non-Canonical Amino Acids as a Tool for the Thermal Stabilization of Enzymes. Protein Eng. Des. Sel. 2023, 36, gzad003,  DOI: 10.1093/protein/gzad003
  26. 26
    Mirts, E. N.; Bhagi-Damodaran, A.; Lu, Y. Understanding and Modulating Metalloenzymes with Unnatural Amino Acids, Non-Native Metal Ions, and Non-Native Metallocofactors. Acc. Chem. Res. 2019, 52 (4), 935944,  DOI: 10.1021/acs.accounts.9b00011
  27. 27
    Drienovská, I.; Roelfes, G. Expanding the Enzyme Universe with Genetically Encoded Unnatural Amino Acids. Nat. Catal. 2020, 3 (3), 193202,  DOI: 10.1038/s41929-019-0410-8
  28. 28
    Trimble, J. S.; Crawshaw, R.; Hardy, F. J.; Levy, C. W.; Brown, M. J. B.; Fuerst, D. E.; Heyes, D. J.; Obexer, R.; Green, A. P. A Designed Photoenzyme for Enantioselective [2 + 2] Cycloadditions. Nature 2022, 611 (7937), 709714,  DOI: 10.1038/s41586-022-05335-3
  29. 29
    Boutureira, O.; Bernardes, G. J. L. Advances in Chemical Protein Modification. Chem. Rev. 2015, 115 (5), 21742195,  DOI: 10.1021/cr500399p
  30. 30
    Pfleiderer, G. Chemical Modification of Proteins. In Review Articles, including those from an International Conference held in Bielefeld, FR of Germany, June 1–2, 1984 , 2021; Walter de Gruyter GmbH & Co KG, p 207.  DOI: 10.1515/9783112393048-011
  31. 31
    Gunnoo, S. B.; Madder, A. Chemical Protein Modification through Cysteine. ChemBioChem 2016, 17 (7), 529553,  DOI: 10.1002/cbic.201500667
  32. 32
    Bischak, C. G.; Longhi, S.; Snead, D. M.; Costanzo, S.; Terrer, E.; Londergan, C. H. Probing Structural Transitions in the Intrinsically Disordered C-Terminal Domain of the Measles Virus Nucleoprotein by Vibrational Spectroscopy of Cyanylated Cysteines. Biophys. J. 2010, 99 (5), 16761683,  DOI: 10.1016/j.bpj.2010.06.060
  33. 33
    Fafarman, A. T.; Webb, L. J.; Chuang, J. I.; Boxer, S. G. Site-Specific Conversion of Cysteine Thiols into Thiocyanate Creates an IR Probe for Electric Fields in Proteins. J. Am. Chem. Soc. 2006, 128 (41), 1335613357,  DOI: 10.1021/ja0650403
  34. 34
    Weeks, C. L.; Jo, H.; Kier, B.; DeGrado, W. F.; Spiro, T. G. Cysteine-Linked Aromatic Nitriles as UV Resonance Raman Probes of Protein Structure. J. Raman Spectrosc. 2012, 43 (9), 12441249,  DOI: 10.1002/jrs.3167
  35. 35
    Boutureira, O.; Bernardes, G. J.; D’Hooge, F.; Davis, B. G. Direct Radiolabelling of Proteins at Cysteine Using [18 F]-Fluorosugars. Chem. Commun. 2011, 47 (36), 1001010012,  DOI: 10.1039/c1cc13524d
  36. 36
    Kaiser, E. T.; Lawrence, D. S. Chemical Mutation of Enzyme Active Sites. Science 1984, 226 (4674), 505511,  DOI: 10.1126/science.6238407
  37. 37
    Levine, H. L.; Kaiser, E. Oxidation of Dihydronicotinamides by Flavopapain. J. Am. Chem. Soc. 1978, 100 (24), 76707677,  DOI: 10.1021/ja00492a040
  38. 38
    Levine, H. L.; Nakagawa, Y.; Kaiser, E. Flavopapain: Synthesis and Properties of Semi-Synthetic Enzymes. Biochem. Biophys. Res. Commun. 1977, 76 (1), 6470,  DOI: 10.1016/0006-291X(77)91668-0
  39. 39
    Mayer, C.; Gillingham, D. G.; Ward, T. R.; Hilvert, D. An Artificial Metalloenzyme for Olefin Metathesis. Chem. Commun. 2011, 47 (44), 1206812070,  DOI: 10.1039/c1cc15005g
  40. 40
    den Heeten, R.; Muñoz, B. K.; Popa, G.; Laan, W.; Kamer, P. C. J. Synthesis of Hybrid Transition-Metalloproteins via Thiol-Selective Covalent Anchoring of Rh-Phosphine and Ru-Phenanthroline Complexes. Dalton Trans. 2010, 39 (36), 84778483,  DOI: 10.1039/c0dt00239a
  41. 41
    Deuss, P. J.; Popa, G.; Botting, C. H.; Laan, W.; Kamer, P. C. J. Highly Efficient and Site-Selective Phosphane Modification of Proteins through Hydrazone Linkage: Development of Artificial Metalloenzymes. Angew. Chem. Int. Ed. 2010, 49 (31), 53155317,  DOI: 10.1002/anie.201002174
  42. 42
    Jarvis, A. G.; Obrecht, L.; Deuss, P. J.; Laan, W.; Gibson, E. K.; Wells, P. P.; Kamer, P. C. J. Enzyme Activity by Design: An Artificial Rhodium Hydroformylase for Linear Aldehydes. Angew. Chem. Int. Ed. 2017, 56 (44), 1359613600,  DOI: 10.1002/anie.201705753
  43. 43
    Merrifield, R. B. Solid Phase Peptide Synthesis. I. The Synthesis of a Tetrapeptide. J. Am. Chem. Soc. 1963, 85 (14), 21492154,  DOI: 10.1021/ja00897a025
  44. 44
    Eckert, D. M.; Malashkevich, V. N.; Hong, L. H.; Carr, P. A.; Kim, P. S. Inhibiting HIV-1 Entry: Discovery of D-Peptide Inhibitors that Target the gp41 Coiled-Coil Pocket. Cell 1999, 99 (1), 103115,  DOI: 10.1016/S0092-8674(00)80066-5
  45. 45
    Kent, S.; Sohma, Y.; Liu, S.; Bang, D.; Pentelute, B.; Mandal, K. Through The Looking Glass - a New World of Proteins Enabled by Chemical Synthesis. J. Pept. Sci. 2012, 18 (7), 428436,  DOI: 10.1002/psc.2421
  46. 46
    Schumacher, T. N. M.; Mayr, L. M.; Minor, D. L.; Milhollen, M. A.; Burgess, M. W.; Kim, P. S. Identification of D-Peptide Ligands Through Mirror-Image Phage Display. Science 1996, 271 (5257), 18541857,  DOI: 10.1126/science.271.5257.1854
  47. 47
    Pech, A.; Achenbach, J.; Jahnz, M.; Schülzchen, S.; Jarosch, F.; Bordusa, F.; Klussmann, S. A Thermostable D-Polymerase for Mirror-Image PCR. Nucleic Acids Res. 2017, 45 (7), 39974005,  DOI: 10.1093/nar/gkx079
  48. 48
    Jackson, D. Y.; Burnier, J.; Quan, C.; Stanley, M.; Tom, J.; Wells, J. A. A Designed Peptide Ligase for Total Synthesis of Ribonuclease A with Unnatural Catalytic Residues. Science 1994, 266 (5183), 243247,  DOI: 10.1126/science.7939659
  49. 49
    Kaiser, E. Synthetic Approaches to Biologically Active Peptides and Proteins Including Enzymes. Acc. Chem. Res. 1989, 22 (2), 4754,  DOI: 10.1021/ar00158a001
  50. 50
    Schnölzer, M.; Kent, S. B. Constructing Proteins by Dovetailing Unprotected Synthetic Peptides: Backbone-Engineered HIV Protease. Science 1992, 256 (5054), 221225,  DOI: 10.1126/science.1566069
  51. 51
    Bode, J. W. Chemical Protein Synthesis with the α-Ketoacid-Hydroxylamine Ligation. Acc. Chem. Res. 2017, 50 (9), 21042115,  DOI: 10.1021/acs.accounts.7b00277
  52. 52
    Liu, H.; Li, X. Serine/Threonine Ligation: Origin, Mechanistic Aspects, and Applications. Acc. Chem. Res. 2018, 51 (7), 16431655,  DOI: 10.1021/acs.accounts.8b00151
  53. 53
    Muir, T. W. Semisynthesis of Proteins by Expressed Protein Ligation. Annu. Rev. Biochem. 2003, 72 (1), 249289,  DOI: 10.1146/annurev.biochem.72.121801.161900
  54. 54
    Antos, J. M.; Truttmann, M. C.; Ploegh, H. L. Recent Advances in Sortase-Catalyzed Ligation Methodology. Curr. Opin. Struct. Biol. 2016, 38, 111118,  DOI: 10.1016/j.sbi.2016.05.021
  55. 55
    Chen, S.-Y.; Cressman, S.; Mao, F.; Shao, H.; Low, D. W.; Beilan, H. S.; Cagle, E. N.; Carnevali, M.; Gueriguian, V.; Keogh, P. J. Synthetic Erythropoietic Proteins: Tuning Biological Performance by Site-Specific Polymer Attachment. Chem. Biol. 2005, 12 (3), 371383,  DOI: 10.1016/j.chembiol.2005.01.017
  56. 56
    Kochendoerfer, G. G.; Chen, S.-Y.; Mao, F.; Cressman, S.; Traviglia, S.; Shao, H.; Hunter, C. L.; Low, D. W.; Cagle, E. N.; Carnevali, M. Design and Chemical Synthesis of a Homogeneous Polymer-Modified Erythropoiesis Protein. Science 2003, 299 (5608), 884887,  DOI: 10.1126/science.1079085
  57. 57
    Kent, S. B. H. Bringing the Science of Proteins into the Realm of Organic Chemistry: Total Chemical Synthesis of SEP (Synthetic Erythropoiesis Protein). Angew. Chem. Int. Ed. 2013, 52 (46), 1198811996,  DOI: 10.1002/anie.201304116
  58. 58
    Liu, S.; Pentelute, B. L.; Kent, S. B. H. Convergent Chemical Synthesis of [Lysine24, 38, 83] Human Erythropoietin. Angew. Chem. Int. Ed. 2012, 51 (4), 993999,  DOI: 10.1002/anie.201106060
  59. 59
    Baca, M.; Kent, S. B. Catalytic Contribution of Flap-Substrate Hydrogen Bonds in ″HIV-1 Protease″ Explored by Chemical Synthesis. Proc. Natl. Acad. Sci. U.S.A. 1993, 90 (24), 1163811642,  DOI: 10.1073/pnas.90.24.11638
  60. 60
    Beadle, J. D.; Knuhtsen, A.; Hoose, A.; Raubo, P.; Jamieson, A. G.; Shipman, M. Solid-Phase Synthesis of Oxetane Modified Peptides. Org. Lett. 2017, 19 (12), 33033306,  DOI: 10.1021/acs.orglett.7b01466
  61. 61
    Abdildinova, A.; Kurth, M. J.; Gong, Y.-D. Solid-Phase Synthesis of Peptidomimetics with Peptide Backbone Modifications. Asian J. Org. Chem. 2021, 10 (9), 23002317,  DOI: 10.1002/ajoc.202100264
  62. 62
    Flavell, R. R.; Muir, T. W. Expressed Protein Ligation (EPL) in the Study of Signal Transduction, Ion Conduction, And Chromatin Biology. Acc. Chem. Res. 2009, 42 (1), 107116,  DOI: 10.1021/ar800129c
  63. 63
    Valiyaveetil, F. I.; Leonetti, M.; Muir, T. W.; MacKinnon, R. Ion Selectivity in a Semisynthetic K+ Channel Locked in the Conductive Conformation. Science 2006, 314 (5801), 10041007,  DOI: 10.1126/science.1133415
  64. 64
    Vázquez, M. E.; Nitz, M.; Stehn, J.; Yaffe, M. B.; Imperiali, B. Fluorescent Caged Phosphoserine Peptides as Probes to Investigate Phosphorylation-Dependent Protein Associations. J. Am. Chem. Soc. 2003, 125 (34), 1015010151,  DOI: 10.1021/ja0351847
  65. 65
    Müller, M. M.; Kries, H.; Csuhai, E.; Kast, P.; Hilvert, D. Design, Selection, and Characterization of a Split Chorismate Mutase. Protein Sci. 2010, 19 (5), 10001010,  DOI: 10.1002/pro.377
  66. 66
    Choi, Y.; Shin, S. H.; Jung, H.; Kwon, O.; Seo, J. K.; Kee, J.-M. Specific Fluorescent Probe for Protein Histidine Phosphatase Activity. ACS sensors 2019, 4 (4), 10551062,  DOI: 10.1021/acssensors.9b00242
  67. 67
    Sainlos, M.; Imperiali, B. Tools For Investigating Peptide-Protein Interactions: Peptide Incorporation of Environment-Sensitive Fluorophores Through SPPS-Based ’Building Block’ Approach. Nat. Protoc. 2007, 2 (12), 32103218,  DOI: 10.1038/nprot.2007.443
  68. 68
    Wu, Y.; Tam, W.-S.; Chau, H.-F.; Kaur, S.; Thor, W.; Aik, W. S.; Chan, W.-L.; Zweckstetter, M.; Wong, K.-L. Solid-Phase Fluorescent BODIPY-Peptide Synthesis via in situ Dipyrrin Construction. Chem. Sci. 2020, 11 (41), 1126611273,  DOI: 10.1039/D0SC04849F
  69. 69
    Kienhöfer, A.; Kast, P.; Hilvert, D. Selective Stabilization of the Chorismate Mutase Transition State by a Positively Charged Hydrogen Bond Donor. J. Am. Chem. Soc. 2003, 125 (11), 32063207,  DOI: 10.1021/ja0341992
  70. 70
    Roy, R. S.; Imperiali, B. Pyridoxamine-Amino Acid Chimeras in Semisynthetic Aminotransferase Mimics. Protein eng. 1997, 10 (6), 691698,  DOI: 10.1093/protein/10.6.691
  71. 71
    Lopez, G.; Anderson, J. C. Synthetic Auxotrophs with Ligand-Dependent Essential Genes for a BL21 (DE3) Biosafety Strain. ACS Synth. Biol. 2015, 4 (12), 12791286,  DOI: 10.1021/acssynbio.5b00085
  72. 72
    Budisa, N. Amino Acids and Codons - Code Organization and Protein Structure. Engineering the Genetic Code 2005, 6689,  DOI: 10.1002/3527607188.ch4
  73. 73
    Budisa, N.; Steipe, B.; Demange, P.; Eckerskorn, C.; Kellermann, J.; Huber, R. High-Level Biosynthetic Substitution of Methionine in Proteins by Its Analogs 2-Aminohexanoic Acid, Selenomethionine, Telluromethionine and Ethionine in Escherichia coli. Eur. J. Biochem. 1995, 230 (2), 788796,  DOI: 10.1111/j.1432-1033.1995.0788h.x
  74. 74
    Cowie, D. B.; Cohen, G. N. Biosynthesis by Escherichia coli of Active Altered Proteins Containing Selenium Instead of Sulfur. Biochim. Biophys. Acta 1957, 26 (2), 252261,  DOI: 10.1016/0006-3002(57)90003-3
  75. 75
    Wong, J. Membership Mutation of the Genetic Code: Loss of Fitness by Tryptophan. Proc. Natl. Acad. Sci. U.S.A. 1983, 80 (20), 63036306,  DOI: 10.1073/pnas.80.20.6303
  76. 76
    Dóring, V.; Mootz, H. D.; Nangle, L. A.; Hendrickson, T. L.; de Crécy-Lagard, V.; Schimmel, P.; Marliere, P. Enlarging the Amino Acid Set of Escherichia coli by Infiltration of the Valine Coding Pathway. Science 2001, 292 (5516), 501504,  DOI: 10.1126/science.1057718
  77. 77
    Datta, D.; Wang, P.; Carrico, I. S.; Mayo, S. L.; Tirrell, D. A. A Designed Phenylalanyl-tRNA Synthetase Variant Allows Efficient in Vivo Incorporation of Aryl Ketone Functionality into Proteins. J. Am. Chem. Soc. 2002, 124 (20), 56525653,  DOI: 10.1021/ja0177096
  78. 78
    Sharma, N.; Furter, R.; Kast, P.; Tirrell, D. A. Efficient Introduction of Aryl Bromide Functionality into Proteins in vivo. FEBS Lett. 2000, 467 (1), 3740,  DOI: 10.1016/S0014-5793(00)01120-0
  79. 79
    Kiick, K. L.; van Hest, J. C.; Tirrell, D. A. Expanding the Scope of Protein Biosynthesis by Altering the Methionyl-tRNA Synthetase Activity of a Bacterial Expression Host. Angew. Chem. Int. Ed. 2000, 39 (12), 21482152,  DOI: 10.1002/1521-3773(20000616)39:12<2148::AID-ANIE2148>3.0.CO;2-7
  80. 80
    Hendrickson, W. A.; Horton, J. R.; LeMaster, D. M. Selenomethionyl Proteins Produced for Analysis by Multiwavelength Anomalous Diffraction (MAD): a Vehicle for Direct Determination of Three-Dimensional Structure. EMBO J. 1990, 9 (5), 16651672,  DOI: 10.1002/j.1460-2075.1990.tb08287.x
  81. 81
    Boles, J. O.; Lewinski, K.; Kunkle, M.; Odom, J. D.; Dunlap, R. B.; Lebioda, L.; Hatada, M. Bio-Incorporation of Telluromethionine into Buried Residues of Dihydrofolate Reductase. Nat. Struct. Biol. 1994, 1 (5), 283284,  DOI: 10.1038/nsb0594-283
  82. 82
    Strub, M. P.; Hoh, F.; Sanchez, J. F.; Strub, J. M.; Böck, A.; Aumelas, A.; Dumas, C. Selenomethionine and Selenocysteine Double Labeling Strategy for Crystallographic Phasing. Structure 2003, 11 (11), 13591367,  DOI: 10.1016/j.str.2003.09.014
  83. 83
    Budisa, N.; Karnbrock, W.; Steinbacher, S.; Humm, A.; Prade, L.; Neuefeind, T.; Moroder, L.; Huber, R. Bioincorporation of Telluromethionine into Proteins: A Promising New Approach for X-Ray Structure Analysis of Proteins. J. Mol. Biol. 1997, 270 (4), 616623,  DOI: 10.1006/jmbi.1997.1132
  84. 84
    Bae, J. H.; Alefelder, S.; Kaiser, J. T.; Friedrich, R.; Moroder, L.; Huber, R.; Budisa, N. Incorporation of β-Selenolo[3,2-b]Pyrrolyl-Alanine into Proteins for Phase Determination in Protein X-Ray Crystallography. J. Mol. Biol. 2001, 309 (4), 925936,  DOI: 10.1006/jmbi.2001.4699
  85. 85
    Minks, C.; Huber, R.; Moroder, L.; Budisa, N. Atomic Mutations at the Single Tryptophan Residue of Human Recombinant Annexin V: Effects on Structure, Stability, and Activity. Biochemistry 1999, 38 (33), 1064910659,  DOI: 10.1021/bi990580g
  86. 86
    Renner, C.; Alefelder, S.; Bae, J. H.; Budisa, N.; Huber, R.; Moroder, L. Fluoroprolines as Tools for Protein Design and Engineering. Angew. Chem. Int. Ed. 2001, 40 (5), 923925,  DOI: 10.1002/1521-3773(20010302)40:5<923::AID-ANIE923>3.0.CO;2-#
  87. 87
    Tang, Y.; Ghirlanda, G.; Petka, W. A.; Nakajima, T.; DeGrado, W. F.; Tirrell, D. A. Fluorinated Coiled-Coil Proteins Prepared in Vivo Display Enhanced Thermal and Chemical Stability. Angew. Chem. Int. Ed. 2001, 40 (8), 14941496,  DOI: 10.1002/1521-3773(20010417)40:8<1494::AID-ANIE1494>3.0.CO;2-X
  88. 88
    Tang, Y.; Tirrell, D. A. Biosynthesis of a Highly Stable Coiled-Coil Protein Containing Hexafluoroleucine in an Engineered Bacterial Host. J. Am. Chem. Soc. 2001, 123 (44), 1108911090,  DOI: 10.1021/ja016652k
  89. 89
    Wang, P.; Tang, Y.; Tirrell, D. A. Incorporation of Trifluoroisoleucine into Proteins in Vivo. J. Am. Chem. Soc. 2003, 125 (23), 69006906,  DOI: 10.1021/ja0298287
  90. 90
    Montclare, J. K.; Tirrell, D. A. Evolving Proteins of Novel Composition. Angew. Chem. Int. Ed. 2006, 45 (27), 45184521,  DOI: 10.1002/anie.200600088
  91. 91
    Hyun Bae, J.; Rubini, M.; Jung, G.; Wiegand, G.; Seifert, M. H.J.; Azim, M.K.; Kim, J.-S.; Zumbusch, A.; Holak, T. A.; Moroder, L.; Huber, R.; Budisa, N. Expansion of the Genetic Code Enables Design of a Novel “Gold” Class of Green Fluorescent Proteins. J. Mol. Biol. 2003, 328 (5), 10711081,  DOI: 10.1016/S0022-2836(03)00364-4
  92. 92
    Budisa, N.; Rubini, M.; Bae, J. H.; Weyher, E.; Wenger, W.; Golbik, R.; Huber, R.; Moroder, L. Global Replacement of Tryptophan with Aminotryptophans Generates Non-Invasive Protein-Based Optical pH Sensors. Angew. Chem. Int. Ed. 2002, 41 (21), 40664069,  DOI: 10.1002/1521-3773(20021104)41:21<4066::AID-ANIE4066>3.0.CO;2-6
  93. 93
    Duewel, H. S.; Daub, E.; Robinson, V.; Honek, J. F. Elucidation of Solvent Exposure, Side-Chain Reactivity, and Steric Demands of the Trifluoromethionine Residue in a Recombinant Protein. Biochemistry 2001, 40 (44), 1316713176,  DOI: 10.1021/bi011381b
  94. 94
    Seifert, M. H. J.; Ksiazek, D.; Azim, M. K.; Smialowski, P.; Budisa, N.; Holak, T. A. Slow Exchange in the Chromophore of a Green Fluorescent Protein Variant. J. Am. Chem. Soc. 2002, 124 (27), 79327942,  DOI: 10.1021/ja0257725
  95. 95
    Bann, J. G.; Pinkner, J.; Hultgren, S. J.; Frieden, C. Real-Time and Equilibrium 19F-NMR Studies Reveal the Role of Domain-Domain Interactions in the Folding of the Chaperone PapD. Proc. Natl. Acad. Sci. U.S.A. 2002, 99 (2), 709714,  DOI: 10.1073/pnas.022649599
  96. 96
    Deming, T. J.; Fournier, M. J.; Mason, T. L.; Tirrell, D. A. Biosynthetic Incorporation and Chemical Modification of Alkene Functionality in Genetically Engineered Polymers. Journal of Macromolecular Science, Part A 1997, 34 (10), 21432150,  DOI: 10.1080/10601329708010331
  97. 97
    Kiick, K. L.; Saxon, E.; Tirrell, D. A.; Bertozzi, C. R. Incorporation of Azides into Recombinant Proteins for Chemoselective Modification by the Staudinger Ligation. Proc. Natl. Acad. Sci. U.S.A. 2002, 99 (1), 1924,  DOI: 10.1073/pnas.012583299
  98. 98
    Kothakota, S.; Mason, T. L.; Tirrell, D. A.; Fournier, M. J. Biosynthesis of a Periodic Protein Containing 3-Thienylalanine: A Step Toward Genetically Engineered Conducting Polymers. J. Am. Chem. Soc. 1995, 117 (1), 536537,  DOI: 10.1021/ja00106a064
  99. 99
    Link, A. J.; Tirrell, D. A. Cell Surface Labeling of Escherichia coli via Copper(I)-Catalyzed [3 + 2] Cycloaddition. J. Am. Chem. Soc. 2003, 125 (37), 1116411165,  DOI: 10.1021/ja036765z
  100. 100
    van Hest, J. C. M.; Kiick, K. L.; Tirrell, D. A. Efficient Incorporation of Unsaturated Methionine Analogues into Proteins in Vivo. J. Am. Chem. Soc. 2000, 122 (7), 12821288,  DOI: 10.1021/ja992749j
  101. 101
    van Hest, J. C. M.; Tirrell, D. A. Efficient introduction of alkene functionality into proteins in vivo. FEBS Lett. 1998, 428 (1–2), 6870,  DOI: 10.1016/S0014-5793(98)00489-X
  102. 102
    Omari, K. E.; Ren, J.; Bird, L. E.; Bona, M. K.; Klarmann, G.; LeGrice, S. F. J.; Stammers, D. K. Molecular Architecture and Ligand Recognition Determinants for T4 RNA Ligase. J. Biol. Chem. 2006, 281 (3), 15731579,  DOI: 10.1074/jbc.M509658200
  103. 103
    Xiao, H.; Murakami, H.; Suga, H.; Ferré-D’Amaré, A. R. Structural Basis of Specific tRNA Aminoacylation by a Small in vitro Selected Ribozyme. Nature 2008, 454 (7202), 358361,  DOI: 10.1038/nature07033
  104. 104
    Noren, C. J.; Anthony-Cahill, S. J.; Griffith, M. C.; Schultz, P. G. A General Method for Site-specific Incorporation of Unnatural Amino Acids into Proteins. Science 1989, 244 (4901), 182188,  DOI: 10.1126/science.2649980
  105. 105
    Goto, Y.; Katoh, T.; Suga, H. Flexizymes for Genetic Code Reprogramming. Nat. Protoc. 2011, 6 (6), 779790,  DOI: 10.1038/nprot.2011.331
  106. 106
    Shimizu, Y.; Inoue, A.; Tomari, Y.; Suzuki, T.; Yokogawa, T.; Nishikawa, K.; Ueda, T. Cell-Free Translation Reconstituted with Purified Components. Nat. Biotechnol. 2001, 19 (8), 751755,  DOI: 10.1038/90802
  107. 107
    Chapeville, F.; Lipmann, F.; Ehrenstein, G. v.; Weisblum, B.; Ray, W. J.; Benzer, S. On the Role of Soluble Ribonucleic Acid in Coding for Amino Acids. Proc. Natl. Acad. Sci. U.S.A. 1962, 48 (6), 10861092,  DOI: 10.1073/pnas.48.6.1086
  108. 108
    Fahnestock, S.; Rich, A. Ribosome-Catalyzed Polyester Formation. Science 1971, 173 (3994), 340343,  DOI: 10.1126/science.173.3994.340
  109. 109
    Hecht, S. M.; Alford, B. L.; Kuroda, Y.; Kitano, S. “Chemical Aminoacylation” of tRNA’s. J. Biol. Chem. 1978, 253 (13), 45174520,  DOI: 10.1016/S0021-9258(17)30417-9
  110. 110
    Katoh, T.; Goto, Y.; Passioura, T.; Suga, H. Development of Flexizyme Aminoacylation Ribozymes and Their Applications. Ribozymes 2021, 519543,  DOI: 10.1002/9783527814527.ch20
  111. 111
    Morimoto, J.; Hayashi, Y.; Iwasaki, K.; Suga, H. Flexizymes: Their Evolutionary History and the Origin of Catalytic Function. Acc. Chem. Res. 2011, 44 (12), 13591368,  DOI: 10.1021/ar2000953
  112. 112
    Murakami, H.; Ohta, A.; Ashigai, H.; Suga, H. A Highly Flexible tRNA Acylation Method for Non-Natural Polypeptide Synthesis. Nat. Methods 2006, 3 (5), 357359,  DOI: 10.1038/nmeth877
  113. 113
    Katoh, T.; Suga, H. In Vitro Genetic Code Reprogramming for the Expansion of Usable Noncanonical Amino Acids. Annu. Rev. Biochem. 2022, 91 (1), 221243,  DOI: 10.1146/annurev-biochem-040320-103817
  114. 114
    Goto, Y.; Suga, H. The RaPID Platform for the Discovery of Pseudo-Natural Macrocyclic Peptides. Acc. Chem. Res. 2021, 54 (18), 36043617,  DOI: 10.1021/acs.accounts.1c00391
  115. 115
    Schultz, P. Expanding the Genetic Code. Protein Sci. 2023, 32 (1), e4488  DOI: 10.1002/pro.4488
  116. 116
    Neumann, H.; Wang, K.; Davis, L.; Garcia-Alai, M.; Chin, J. W. Encoding Multiple Unnatural Amino Acids via Evolution of a Quadruplet-Decoding Ribosome. Nature 2010, 464 (7287), 441444,  DOI: 10.1038/nature08817
  117. 117
    Robertson, W. E.; Funke, L. F. H.; de la Torre, D.; Fredens, J.; Elliott, T. S.; Spinck, M.; Christova, Y.; Cervettini, D.; Böge, F. L.; Liu, K. C. Sense Codon Reassignment Enables Viral Resistance and Encoded Polymer Synthesis. Science 2021, 372 (6546), 10571062,  DOI: 10.1126/science.abg3029
  118. 118
    Wang, L.; Brock, A.; Herberich, B.; Schultz, P. G. Expanding the Genetic Code of Escherichia coli. Science 2001, 292 (5516), 498500,  DOI: 10.1126/science.1060077
  119. 119
    Dunkelmann, D. L.; Piedrafita, C.; Dickson, A.; Liu, K. C.; Elliott, T. S.; Fiedler, M.; Bellini, D.; Zhou, A.; Cervettini, D.; Chin, J. W. Adding α,α-Disubstituted and β-Linked Monomers to the Genetic Code of an Organism. Nature 2024, 625 (7995), 603610,  DOI: 10.1038/s41586-023-06897-6
  120. 120
    Bryson, D. I. Continuous Directed Evolution of Aminoacyl-tRNA Synthetases. Nat. Chem. Biol. 2017, 13, 1253,  DOI: 10.1038/nchembio.2474
  121. 121
    Krahn, N.; Tharp, J. M.; Crnković, A.; Söll, D. Chapter Twelve - Engineering Aminoacyl-tRNA Synthetases for Use in Synthetic Biology. In The Enzymes; Ribas de Pouplana, L., Kaguni, L. S., Eds.; Academic Press, 2020; Vol. 48, pp 351395. DOI: 10.1016/bs.enz.2020.06.004
  122. 122
    Amiram, M.; Haimovich, A. D.; Fan, C.; Wang, Y.-S.; Aerni, H.-R.; Ntai, I.; Moonan, D. W.; Ma, N. J.; Rovner, A. J.; Hong, S. H. Evolution of Translation Machinery in Recoded Bacteria Enables Multi-Site Incorporation of Nonstandard Amino Acids. Nat. Biotechnol. 2015, 33 (12), 12721279,  DOI: 10.1038/nbt.3372
  123. 123
    Hohl, A.; Karan, R.; Akal, A.; Renn, D.; Liu, X.; Ghorpade, S.; Groll, M.; Rueping, M.; Eppinger, J. Engineering a Polyspecific Pyrrolysyl-tRNA Synthetase by a High Throughput FACS Screen. Sci. Rep. 2019, 9 (1), 11971,  DOI: 10.1038/s41598-019-48357-0
  124. 124
    Wang, L.; Xie, J.; Schultz, P. G. Expanding the Genetic Code. Annu. Rev. Biophys. Biomol. Struct. 2006, 35, 225249,  DOI: 10.1146/annurev.biophys.35.101105.121507
  125. 125
    Santoro, S.; Wang, L.; Herberich, B.; King, D. S.; Schultz, P. G. An Efficient System for the Evolution of Aminoacyl-tRNA Synthetase Specificity. Nat. Biotechnol. 2002, 20, 10441048,  DOI: 10.1038/nbt742
  126. 126
    Dumas, A.; Lercher, L.; Spicer, C. D.; Davis, B. G. Designing Logical Codon Reassignment - Expanding the Chemistry in Biology. Chem. Sci. 2015, 6 (1), 5069,  DOI: 10.1039/C4SC01534G
  127. 127
    Wan, W.; Tharp, J. M.; Liu, W. R. Pyrrolysyl-tRNA Synthetase: An Ordinary Enzyme but an Outstanding Genetic Code Expansion Tool. Biochim. Biophys. Acta Proteins Proteomics 2014, 1844 (6), 10591070,  DOI: 10.1016/j.bbapap.2014.03.002
  128. 128
    Ryu, Y.; Schultz, P. G. Efficient Incorporation of Unnatural Amino Acids into Proteins in Escherichia coli. Nat. Methods 2006, 3 (4), 263265,  DOI: 10.1038/nmeth864
  129. 129
    Srinivasan, G.; James, C. M.; Krzycki, J. A. Pyrrolysine Encoded by UAG in Archaea: Charging of a UAG-Decoding Specialized tRNA. Science 2002, 296 (5572), 14591462,  DOI: 10.1126/science.1069588
  130. 130
    Hao, B.; Gong, W.; Ferguson, T. K.; James, C. M.; Krzycki, J. A.; Chan, M. K. A New UAG-Encoded Residue in the Structure of a Methanogen Methyltransferase. Science 2002, 296 (5572), 14621466,  DOI: 10.1126/science.1069556
  131. 131
    Chin, J. W. Expanding and Reprogramming the Genetic Code of Cells and Animals. Annu. Rev. Biochem. 2014, 83 (1), 379408,  DOI: 10.1146/annurev-biochem-060713-035737
  132. 132
    Ranaghan, K. E.; Hung, J. E.; Bartlett, G. J.; Mooibroek, T. J.; Harvey, J. N.; Woolfson, D. N.; van der Donk, W. A.; Mulholland, A. J. A Catalytic Role for Methionine Revealed by a Combination of Computation and Experiments on Phosphite Dehydrogenase. Chem. Sci. 2014, 5 (6), 21912199,  DOI: 10.1039/C3SC53009D
  133. 133
    Ekanayake, K. B.; Mahawaththa, M. C.; Qianzhu, H.; Abdelkader, E. H.; George, J.; Ullrich, S.; Murphy, R. B.; Fry, S. E.; Johansen-Leete, J.; Payne, R. J. Probing Ligand Binding Sites on Large Proteins by Nuclear Magnetic Resonance Spectroscopy of Genetically Encoded Non-Canonical Amino Acids. J. Med. Chem. 2023, 66 (7), 52895304,  DOI: 10.1021/acs.jmedchem.3c00222
  134. 134
    Ellman, J. A.; Volkman, B. F.; Mendel, D.; Schulz, P. G.; Wemmer, D. E. Site-Specific Isotopic Labeling of Proteins for NMR Studies. J. Am. Chem. Soc. 1992, 114 (20), 79597961,  DOI: 10.1021/ja00046a080
  135. 135
    Schmidt, M. J.; Borbas, J.; Drescher, M.; Summerer, D. A Genetically Encoded Spin Label for Electron Paramagnetic Resonance Distance Measurements. J. Am. Chem. Soc. 2014, 136 (4), 12381241,  DOI: 10.1021/ja411535q
  136. 136
    Fafarman, A. T.; Boxer, S. G. Nitrile Bonds as Infrared Probes of Electrostatics in Ribonuclease S. J. Phys. Chem. B 2010, 114 (42), 1353613544,  DOI: 10.1021/jp106406p
  137. 137
    Xie, J.; Wang, L.; Wu, N.; Brock, A.; Spraggon, G.; Schultz, P. G. The Site-Specific Incorporation of p-Iodo-L-Phenylalanine into Proteins for Structure Determination. Nat. Biotechnol. 2004, 22 (10), 12971301,  DOI: 10.1038/nbt1013
  138. 138
    Summerer, D.; Chen, S.; Wu, N.; Deiters, A.; Chin, J. W.; Schultz, P. G. A Genetically Encoded Fluorescent Amino Acid. Proc. Natl. Acad. Sci. U.S.A. 2006, 103 (26), 97859789,  DOI: 10.1073/pnas.0603965103
  139. 139
    Lang, K.; Davis, L.; Wallace, S.; Mahesh, M.; Cox, D. J.; Blackman, M. L.; Fox, J. M.; Chin, J. W. Genetic Encoding of Bicyclononynes and trans-Cyclooctenes for Site-Specific Protein Labeling in Vitro and in Live Mammalian Cells via Rapid Fluorogenic Diels-Alder Reactions. J. Am. Chem. Soc. 2012, 134 (25), 1031710320,  DOI: 10.1021/ja302832g
  140. 140
    Plass, T.; Milles, S.; Koehler, C.; Schultz, C.; Lemke, E. A. Genetically Encoded Copper-Free Click Chemistry. Angew. Chem. Int. Ed. 2011, 50 (17), 38783881,  DOI: 10.1002/anie.201008178
  141. 141
    Brustad, E. M.; Lemke, E. A.; Schultz, P. G.; Deniz, A. A. A General and Efficient Method for the Site-Specific Dual-Labeling of Proteins for Single Molecule Fluorescence Resonance Energy Transfer. J. Am. Chem. Soc. 2008, 130 (52), 1766417665,  DOI: 10.1021/ja807430h
  142. 142
    Loving, G.; Imperiali, B. A Versatile Amino Acid Analogue of the Solvatochromic Fluorophore 4-N,N-Dimethylamino-1,8-naphthalimide: A Powerful Tool for the Study of Dynamic Protein Interactions. J. Am. Chem. Soc. 2008, 130 (41), 1363013638,  DOI: 10.1021/ja804754y
  143. 143
    Mendes, K. R.; Martinez, J. A.; Kantrowitz, E. R. Asymmetric Allosteric Signaling in Aspartate Transcarbamoylase. ACS Chem. Biol. 2010, 5 (5), 499506,  DOI: 10.1021/cb9003207
  144. 144
    Dean, S. F.; Whalen, K. L.; Spies, M. A. Biosynthesis of a Novel Glutamate Racemase Containing a Site-Specific 7-Hydroxycoumarin Amino Acid: Enzyme-Ligand Promiscuity Revealed at the Atomistic Level. ACS Cent. Sci. 2015, 1 (7), 364373,  DOI: 10.1021/acscentsci.5b00211
  145. 145
    Wang, J.; Xie, J.; Schultz, P. G. A Genetically Encoded Fluorescent Amino Acid. J. Am. Chem. Soc. 2006, 128 (27), 87388739,  DOI: 10.1021/ja062666k
  146. 146
    Li, M.; Peng, T. Genetic Encoding of a Fluorescent Noncanonical Amino Acid as a FRET Donor for the Analysis of Deubiquitinase Activities. In Genetically Incorporated Non-Canonical Amino Acids: Methods and Protocols; Tsai, Y.-H., Elsässer, S. J., Eds.; Springer: US, 2023; pp 55- 67. DOI: 10.1007/978-1-0716-3251-2_4
  147. 147
    Miyake-Stoner, S. J.; Miller, A. M.; Hammill, J. T.; Peeler, J. C.; Hess, K. R.; Mehl, R. A.; Brewer, S. H. Probing Protein Folding Using Site-Specifically Encoded Unnatural Amino Acids as FRET Donors with Tryptophan. Biochemistry 2009, 48 (25), 59535962,  DOI: 10.1021/bi900426d
  148. 148
    Bergfors, T. M. Protein crystallization. Internat’l University Line 2009. ISBN: 978–0-9720774–4-6.
  149. 149
    Sakamoto, K.; Murayama, K.; Oki, K.; Iraha, F.; Kato-Murayama, M.; Takahashi, M.; Ohtake, K.; Kobayashi, T.; Kuramitsu, S.; Shirouzu, M. Genetic Encoding of 3-Iodo-l-Tyrosine in Escherichia coli for Single-Wavelength Anomalous Dispersion Phasing in Protein Crystallography. Structure 2009, 17 (3), 335344,  DOI: 10.1016/j.str.2009.01.008
  150. 150
    Lee, H. S.; Spraggon, G.; Schultz, P. G.; Wang, F. Genetic Incorporation of a Metal-Ion Chelating Amino Acid into Proteins as a Biophysical Probe. J. Am. Chem. Soc. 2009, 131 (7), 24812483,  DOI: 10.1021/ja808340b
  151. 151
    Nogly, P.; Weinert, T.; James, D.; Carbajo, S.; Ozerov, D.; Furrer, A.; Gashi, D.; Borin, V.; Skopintsev, P.; Jaeger, K. Retinal Isomerization in Bacteriorhodopsin Captured by a Femtosecond X-Ray Laser. Science 2018, 361 (6398), eaat0094  DOI: 10.1126/science.aat0094
  152. 152
    Tenboer, J.; Basu, S.; Zatsepin, N.; Pande, K.; Milathianaki, D.; Frank, M.; Hunter, M.; Boutet, S.; Williams, G. J.; Koglin, J. E. Time-Resolved Serial Crystallography Captures High-Resolution Intermediates of Photoactive Yellow Protein. Science 2014, 346 (6214), 12421246,  DOI: 10.1126/science.1259357
  153. 153
    Suga, M.; Akita, F.; Sugahara, M.; Kubo, M.; Nakajima, Y.; Nakane, T.; Yamashita, K.; Umena, Y.; Nakabayashi, M.; Yamane, T. Light-Induced Structural Changes and the Site of O = O Bond Formation in PSII Caught by XFEL. Nature 2017, 543 (7643), 131135,  DOI: 10.1038/nature21400
  154. 154
    Dods, R.; Båth, P.; Morozov, D.; Gagnér, V. A.; Arnlund, D.; Luk, H. L.; Kübel, J.; Maj, M.; Vallejos, A.; Wickstrand, C. Ultrafast Structural Changes within a Photosynthetic Reaction Centre. Nature 2021, 589 (7841), 310314,  DOI: 10.1038/s41586-020-3000-7
  155. 155
    Chapman, H. N. X-Ray Free-Electron Lasers for the Structure and Dynamics of Macromolecules. Annu. Rev. Biochem. 2019, 88 (1), 3558,  DOI: 10.1146/annurev-biochem-013118-110744
  156. 156
    Kern, J.; Chatterjee, R.; Young, I. D.; Fuller, F. D.; Lassalle, L.; Ibrahim, M.; Gul, S.; Fransson, T.; Brewster, A. S.; Alonso-Mori, R. Structures of the Intermediates of Kok’s Photosynthetic Water Oxidation Clock. Nature 2018, 563 (7731), 421425,  DOI: 10.1038/s41586-018-0681-2
  157. 157
    Nango, E.; Royant, A.; Kubo, M.; Nakane, T.; Wickstrand, C.; Kimura, T.; Tanaka, T.; Tono, K.; Song, C.; Tanaka, R. A Three-Dimensional Movie of Structural Changes in Bacteriorhodopsin. Science 2016, 354 (6319), 15521557,  DOI: 10.1126/science.aah3497
  158. 158
    Liu, X.; Liu, P.; Li, H.; Xu, Z.; Jia, L.; Xia, Y.; Yu, M.; Tang, W.; Zhu, X.; Chen, C. Excited-State Intermediates in a Designer Protein Encoding a Phototrigger Caught by an X-Ray Free-Electron Laser. Nat. Chem. 2022, 14 (9), 10541060,  DOI: 10.1038/s41557-022-00992-3
  159. 159
    Hosaka, T.; Katsura, K.; Ishizuka-Katsura, Y.; Hanada, K.; Ito, K.; Tomabechi, Y.; Inoue, M.; Akasaka, R.; Takemoto, C.; Shirouzu, M. Crystal Structure of an Archaeal Tyrosyl-tRNA Synthetase Bound to Photocaged L-Tyrosine and Its Potential Application to Time-Resolved X-ray Crystallography. Int. J. Mol. Sci. 2022, 23 (18), 10399,  DOI: 10.3390/ijms231810399
  160. 160
    Markley, J. L.; Putter, I.; Jardetzky, O. High-Resolution Nuclear Magnetic Resonance Spectra of Selectively Deuterated Staphylococcal Nuclease. Science 1968, 161 (3847), 12491251,  DOI: 10.1126/science.161.3847.1249
  161. 161
    Deiters, A.; Geierstanger, B. H.; Schultz, P. G. Site-Specific in vivo Labeling of Proteins for NMR Studies. ChemBioChem 2005, 6 (1), 5558,  DOI: 10.1002/cbic.200400319
  162. 162
    Jones, D. H.; Cellitti, S. E.; Hao, X.; Zhang, Q.; Jahnz, M.; Summerer, D.; Schultz, P. G.; Uno, T.; Geierstanger, B. H. Site-Specific Labeling of Proteins with NMR-Active Unnatural Amino Acids. J. Biomol. NMR 2010, 46 (1), 89100,  DOI: 10.1007/s10858-009-9365-4
  163. 163
    Abdelkader, E. H.; Qianzhu, H.; Huber, T.; Otting, G. Genetic Encoding of 7-Aza-l-tryptophan: Isoelectronic Substitution of a Single CH-Group in a Protein for a Nitrogen Atom for Site-Selective Isotope Labeling. ACS Sensors 2023, 8 (11), 44024406,  DOI: 10.1021/acssensors.3c01904
  164. 164
    Cellitti, S. E.; Jones, D. H.; Lagpacan, L.; Hao, X.; Zhang, Q.; Hu, H.; Brittain, S. M.; Brinker, A.; Caldwell, J.; Bursulaya, B. In vivo Incorporation of Unnatural Amino Acids to Probe Structure, Dynamics, and Ligand Binding in a Large Protein by Nuclear Magnetic Resonance Spectroscopy. J. Am. Chem. Soc. 2008, 130 (29), 92689281,  DOI: 10.1021/ja801602q
  165. 165
    Lampe, J. N.; Floor, S. N.; Gross, J. D.; Nishida, C. R.; Jiang, Y.; Trnka, M. J.; Ortiz de Montellano, P. R. Ligand-Induced Conformational Heterogeneity of Cytochrome P450 CYP119 Identified by 2D NMR Spectroscopy with the Unnatural Amino Acid (13)C-p-Methoxyphenylalanine. J. Am. Chem. Soc. 2008, 130 (48), 1616816169,  DOI: 10.1021/ja8071463
  166. 166
    Hull, W. E.; Sykes, B. D. Fluorotyrosine Alkaline Phosphatase. Fluorine-19 Nuclear Magnetic Resonance Relaxation Times and Molecular Motion of the Individual Fluorotyrosines. Biochemistry 1974, 13 (17), 34313437,  DOI: 10.1021/bi00714a002
  167. 167
    Gamcsik, M. P.; Gerig, J. T. NMR Studies of Fluorophenylalanine-Containing Carbonic Anhydrase. FEBS Lett. 1986, 196 (1), 7174,  DOI: 10.1016/0014-5793(86)80216-2
  168. 168
    Jackson, J. C.; Hammill, J. T.; Mehl, R. A. Site-Specific Incorporation of a 19F-Amino Acid into Proteins as an NMR Probe for Characterizing Protein Structure and Reactivity. J. Am. Chem. Soc. 2007, 129 (5), 11601166,  DOI: 10.1021/ja064661t
  169. 169
    Hammill, J. T.; Miyake-Stoner, S.; Hazen, J. L.; Jackson, J. C.; Mehl, R. A. Preparation of Site-Specifically Labeled Fluorinated Proteins for 19F-NMR Structural Characterization. Nat. Protoc. 2007, 2 (10), 26012607,  DOI: 10.1038/nprot.2007.379
  170. 170
    Wacks, D. B.; Schachman, H. K. 19F Nuclear Magnetic Resonance Studies of Fluorotyrosine-Labeled Aspartate Transcarbamoylase. Properties of the Enzyme and Its Catalytic and Regulatory Subunits. J. Biol. Chem. 1985, 260 (21), 1165111658,  DOI: 10.1016/S0021-9258(17)39080-4
  171. 171
    Gerig, J. T. Fluorine NMR of Proteins. Prog. Nucl. Magn. Reson. Spectrosc. 1994, 26, 293370,  DOI: 10.1016/0079-6565(94)80009-X
  172. 172
    Furter, R. Expansion of the Genetic Code: Site-Directed p-Fluoro-Phenylalanine Incorporation in Escherichia coli. Protein Sci. 1998, 7 (2), 419426,  DOI: 10.1002/pro.5560070223
  173. 173
    Kim, H.-W.; Perez, J. A.; Ferguson, S. J.; Campbell, I. D. The Specific Incorporation of Labelled Aromatic Amino Acids into Proteins through Growth of Bacteria in the Presence of Glyphosate. FEBS Lett. 1990, 272 (1–2), 3436,  DOI: 10.1016/0014-5793(90)80442-L
  174. 174
    Niu, W.; Shu, Q.; Chen, Z.; Mathews, S.; Di Cera, E.; Frieden, C. The Role of Zn2+ on the Structure and Stability of Murine Adenosine Deaminase. J. Phys. Chem. B 2010, 114 (49), 1615616165,  DOI: 10.1021/jp106041v
  175. 175
    Ruben, E. A.; Gandhi, P. S.; Chen, Z.; Koester, S. K.; DeKoster, G. T.; Frieden, C.; Di Cera, E. 19F NMR Reveals the Conformational Properties of Free Thrombin and Its Zymogen Precursor Prethrombin-2. J. Biol. Chem. 2020, 295 (24), 82278235,  DOI: 10.1074/jbc.RA120.013419
  176. 176
    Duewel, H.; Daub, E.; Robinson, V.; Honek, J. F. Incorporation of Trifluoromethionine into a Phage Lysozyme: Implications and a New Marker for Use in Protein 19F NMR. Biochemistry 1997, 36 (11), 34043416,  DOI: 10.1021/bi9617973
  177. 177
    Cleve, P.; Robinson, V.; Duewel, H. S.; Honek, J. F. Difluoromethionine as a Novel 19F NMR Structural Probe for Internal Amino Acid Packing in Proteins. J. Am. Chem. Soc. 1999, 121 (37), 84758478,  DOI: 10.1021/ja9911418
  178. 178
    Holzberger, B.; Rubini, M.; Möller, H. M.; Marx, A. A Highly Active DNA Polymerase with a Fluorous Core. Angew. Chem. Int. Ed. 2010, 49 (7), 13241327,  DOI: 10.1002/anie.200905978
  179. 179
    Orton, H. W.; Qianzhu, H.; Abdelkader, E. H.; Habel, E. I.; Tan, Y. J.; Frkic, R. L.; Jackson, C. J.; Huber, T.; Otting, G. Through-Space Scalar 19F-19F Couplings between Fluorinated Noncanonical Amino Acids for the Detection of Specific Contacts in Proteins. J. Am. Chem. Soc. 2021, 143 (46), 1958719598,  DOI: 10.1021/jacs.1c10104
  180. 180
    Miao, Q.; Nitsche, C.; Orton, H.; Overhand, M.; Otting, G.; Ubbink, M. Paramagnetic Chemical Probes for Studying Biological Macromolecules. Chem. Rev. 2022, 122 (10), 95719642,  DOI: 10.1021/acs.chemrev.1c00708
  181. 181
    Fanucci, G. E.; Cafiso, D. S. Recent advances and applications of site-directed spin labeling. Curr. Opin. Struct. Biol. 2006, 16 (5), 644653,  DOI: 10.1016/j.sbi.2006.08.008
  182. 182
    Braun, T.; Drescher, M.; Summerer, D. Expanding the Genetic Code for Site-Directed Spin-Labeling. Int. J. Mol. Sci. 2019, 20 (2), 373,  DOI: 10.3390/ijms20020373
  183. 183
    Kálai, T.; Fleissner, M. R.; Jeko, J.; Hubbell, W. L.; Hideg, K. Synthesis of New Spin Labels for Cu-Free Click Conjugation. Tetrahedron Lett. 2011, 52 (21), 27472749,  DOI: 10.1016/j.tetlet.2011.03.077
  184. 184
    Kugele, A.; Braun, T. S.; Widder, P.; Williams, L.; Schmidt, M. J.; Summerer, D.; Drescher, M. Site-Directed Spin Labelling of Proteins by Suzuki-Miyaura Coupling via a Genetically Encoded Aryliodide Amino Acid. Chem. Commun. 2019, 55 (13), 19231926,  DOI: 10.1039/C8CC09325C
  185. 185
    Jana, S.; Evans, E. G. B.; Jang, H. S.; Zhang, S.; Zhang, H.; Rajca, A.; Gordon, S. E.; Zagotta, W. N.; Stoll, S.; Mehl, R. A. Ultrafast Bioorthogonal Spin-Labeling and Distance Measurements in Mammalian Cells Using Small, Genetically Encoded Tetrazine Amino Acids. J. Am. Chem. Soc. 2023, 145 (27), 1460814620,  DOI: 10.1021/jacs.3c00967
  186. 186
    Fleissner, M. R.; Brustad, E. M.; Kálai, T.; Altenbach, C.; Cascio, D.; Peters, F. B.; Hideg, K.; Peuker, S.; Schultz, P. G.; Hubbell, W. L. Site-Directed Spin Labeling of a Genetically Encoded Unnatural Amino Acid. Proc. Natl. Acad. Sci. U.S.A. 2009, 106 (51), 2163721642,  DOI: 10.1073/pnas.0912009106
  187. 187
    Nguyen, D. P.; Lusic, H.; Neumann, H.; Kapadnis, P. B.; Deiters, A.; Chin, J. W. Genetic Encoding and Labeling of Aliphatic Azides and Alkynes in Recombinant Proteins via a Pyrrolysyl-tRNA Synthetase/tRNACUA Pair and Click Chemistry. J. Am. Chem. Soc. 2009, 131 (25), 87208721,  DOI: 10.1021/ja900553w
  188. 188
    Chin, J. W.; Santoro, S. W.; Martin, A. B.; King, D. S.; Wang, L.; Schultz, P. G. Addition of p-Azido-l-phenylalanine to the Genetic Code of Escherichia coli. J. Am. Chem. Soc. 2002, 124 (31), 90269027,  DOI: 10.1021/ja027007w
  189. 189
    Kucher, S.; Korneev, S.; Tyagi, S.; Apfelbaum, R.; Grohmann, D.; Lemke, E. A.; Klare, J. P.; Steinhoff, H.-J.; Klose, D. Orthogonal Spin Labeling Using Click Chemistry for in vitro and in vivo Applications. J. Magn. Reson. 2017, 275, 3845,  DOI: 10.1016/j.jmr.2016.12.001
  190. 190
    Abdelkader, E. H.; Feintuch, A.; Yao, X.; Adams, L. A.; Aurelio, L.; Graham, B.; Goldfarb, D.; Otting, G. Protein Conformation by EPR Spectroscopy Using Gadolinium Tags Clicked to Genetically Encoded p-Azido-L-Phenylalanine. Chem. Commun. 2015, 51 (88), 1589815901,  DOI: 10.1039/C5CC07121F
  191. 191
    Loh, C. T.; Ozawa, K.; Tuck, K. L.; Barlow, N.; Huber, T.; Otting, G.; Graham, B. Lanthanide Tags for Site-Specific Ligation to an Unnatural Amino Acid and Generation of Pseudocontact Shifts in Proteins. Bioconjug. Chem. 2013, 24 (2), 260268,  DOI: 10.1021/bc300631z
  192. 192
    Mahawaththa, M. C.; Lee, M. D.; Giannoulis, A.; Adams, L. A.; Feintuch, A.; Swarbrick, J. D.; Graham, B.; Nitsche, C.; Goldfarb, D.; Otting, G. Small Neutral Gd(iii) Tags for Distance Measurements in Proteins by Double Electron-Electron Resonance Experiments. Phys. Chem. Chem. Phys. 2018, 20 (36), 2353523545,  DOI: 10.1039/C8CP03532F
  193. 193
    Yang, H.; Yang, S.; Kong, J.; Dong, A.; Yu, S. Obtaining Information About Protein Secondary Structures in Aqueous Solution Using Fourier Transform IR Spectroscopy. Nat. Protoc. 2015, 10 (3), 382396,  DOI: 10.1038/nprot.2015.024
  194. 194
    Lorenz-Fonfria, V. A. Infrared Difference Spectroscopy of Proteins: From Bands to Bonds. Chem. Rev. 2020, 120 (7), 34663576,  DOI: 10.1021/acs.chemrev.9b00449
  195. 195
    Chung, J. K.; Thielges, M. C.; Fayer, M. D. Dynamics of the Folded and Unfolded Villin Headpiece (HP35) Measured with Ultrafast 2D IR Vibrational Echo Spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2011, 108 (9), 35783583,  DOI: 10.1073/pnas.1100587108
  196. 196
    Chung, J. K.; Thielges, M. C.; Fayer, M. D. Conformational Dynamics and Stability of HP35 Studied with 2D IR Vibrational Echoes. J. Am. Chem. Soc. 2012, 134 (29), 1211812124,  DOI: 10.1021/ja303017d
  197. 197
    Urbanek, D. C.; Vorobyev, D. Y.; Serrano, A. L.; Gai, F.; Hochstrasser, R. M. The Two-Dimensional Vibrational Echo of a Nitrile Probe of the Villin HP35 Protein. J. Phys. Chem. Lett. 2010, 1 (23), 33113315,  DOI: 10.1021/jz101367d
  198. 198
    Bagchi, S.; Boxer, S. G.; Fayer, M. D. Ribonuclease S Dynamics Measured Using a Nitrile Label with 2D IR Vibrational Echo Spectroscopy. J. Phys. Chem. B 2012, 116 (13), 40344042,  DOI: 10.1021/jp2122856
  199. 199
    Tharp, J. M.; Wang, Y.-S.; Lee, Y.-J.; Yang, Y.; Liu, W. R. Genetic Incorporation of Seven ortho-Substituted Phenylalanine Derivatives. ACS Chem. Biol. 2014, 9 (4), 884890,  DOI: 10.1021/cb400917a
  200. 200
    van Wilderen, L. J. G. W.; Kern-Michler, D.; Müller-Werkmeister, H. M.; Bredenbeck, J. Vibrational Dynamics and Solvatochromism of the Label SCN in Various Solvents and Hemoglobin by Time Dependent IR and 2D-IR Spectroscopy. Phys. Chem. Chem. Phys. 2014, 16 (36), 1964319653,  DOI: 10.1039/C4CP01498G
  201. 201
    Bloem, R.; Koziol, K.; Waldauer, S. A.; Buchli, B.; Walser, R.; Samatanga, B.; Jelesarov, I.; Hamm, P. Ligand Binding Studied by 2D IR Spectroscopy Using the Azidohomoalanine Label. J. Phys. Chem. B 2012, 116 (46), 1370513712,  DOI: 10.1021/jp3095209
  202. 202
    Thielges, M. C.; Axup, J. Y.; Wong, D.; Lee, H. S.; Chung, J. K.; Schultz, P. G.; Fayer, M. D. Two-Dimensional IR Spectroscopy of Protein Dynamics Using Two Vibrational Labels: A Site-Specific Genetically Encoded Unnatural Amino Acid and an Active Site Ligand. J. Phys. Chem. B 2011, 115 (38), 1129411304,  DOI: 10.1021/jp206986v
  203. 203
    Ye, S.; Zaitseva, E.; Caltabiano, G.; Schertler, G. F. X.; Sakmar, T. P.; Deupi, X.; Vogel, R. Tracking G-Protein-Coupled Receptor Activation Using Genetically Encoded Infrared Probes. Nature 2010, 464 (7293), 13861389,  DOI: 10.1038/nature08948
  204. 204
    Le Sueur, A. L.; Horness, R. E.; Thielges, M. C. Applications of Two-Dimensional Infrared Spectroscopy. Analyst 2015, 140 (13), 43364349,  DOI: 10.1039/C5AN00558B
  205. 205
    Smith, E. E.; Linderman, B. Y.; Luskin, A. C.; Brewer, S. H. Probing Local Environments with the Infrared Probe: l-4-Nitrophenylalanine. J. Phys. Chem. B 2011, 115 (10), 23802385,  DOI: 10.1021/jp109288j
  206. 206
    Kooter, I. M.; Moguilevsky, N.; Bollen, A.; van der Veen, L. A.; Otto, C.; Dekker, H. L.; Wever, R. The Sulfonium Ion Linkage in Myeloperoxidase: direct spectroscopic detection by isotopic labeling and effect of mutation. J. Biol. Chem. 1999, 274 (38), 2679426802,  DOI: 10.1074/jbc.274.38.26794
  207. 207
    Thielges, M. C.; Case, D. A.; Romesberg, F. E. Carbon-Deuterium Bonds as Probes of Dihydrofolate Reductase. J. Am. Chem. Soc. 2008, 130 (20), 65976603,  DOI: 10.1021/ja0779607
  208. 208
    Le Sueur, A. L.; Schaugaard, R. N.; Baik, M.-H.; Thielges, M. C. Methionine Ligand Interaction in a Blue Copper Protein Characterized by Site-Selective Infrared Spectroscopy. J. Am. Chem. Soc. 2016, 138 (22), 71877193,  DOI: 10.1021/jacs.6b03916
  209. 209
    Horness, R. E.; Basom, E. J.; Thielges, M. C. Site-Selective Characterization of Src Homology 3 Domain Molecular Recognition with Cyanophenylalanine Infrared Probes. Anal. Methods 2015, 7 (17), 72347241,  DOI: 10.1039/C5AY00523J
  210. 210
    Basom, E. J.; Maj, M.; Cho, M.; Thielges, M. C. Site-Specific Characterization of Cytochrome P450cam Conformations by Infrared Spectroscopy. Anal. Chem. 2016, 88 (12), 65986606,  DOI: 10.1021/acs.analchem.6b01520
  211. 211
    Bagchi, S.; Fried, S. D.; Boxer, S. G. A Solvatochromic Model Calibrates Nitriles’ Vibrational Frequencies to Electrostatic Fields. J. Am. Chem. Soc. 2012, 134 (25), 1037310376,  DOI: 10.1021/ja303895k
  212. 212
    Wang, L.; Zhang, J.; Han, M.-J.; Zhang, L.; Chen, C.; Huang, A.; Xie, R.; Wang, G.; Zhu, J.; Wang, Y. A Genetically Encoded Two-Dimensional Infrared Probe for Enzyme Active-Site Dynamics. Angew. Chem. Int. Ed. 2021, 60 (20), 1114311147,  DOI: 10.1002/anie.202016880
  213. 213
    Huguenin-Dezot, N.; Alonzo, D. A.; Heberlig, G. W.; Mahesh, M.; Nguyen, D. P.; Dornan, M. H.; Boddy, C. N.; Schmeing, T. M.; Chin, J. W. Trapping Biosynthetic Acyl-Enzyme Intermediates with Encoded 2,3-Diaminopropionic Acid. Nature 2019, 565 (7737), 112117,  DOI: 10.1038/s41586-018-0781-z
  214. 214
    Tang, S.; Beattie, A. T.; Kafkova, L.; Petris, G.; Huguenin-Dezot, N.; Fiedler, M.; Freeman, M.; Chin, J. W. Mechanism-Based Traps Enable Protease and Hydrolase Substrate Discovery. Nature 2022, 602 (7898), 701707,  DOI: 10.1038/s41586-022-04414-9
  215. 215
    Dorman, G.; Prestwich, G. D. Benzophenone Photophores in Biochemistry. Biochemistry 1994, 33 (19), 56615673,  DOI: 10.1021/bi00185a001
  216. 216
    Chin, J. W.; Martin, A. B.; King, D. S.; Wang, L.; Schultz, P. G. Addition of a Photocrosslinking Amino Acid to the Genetic Code of Escherichia coli. Proc. Natl. Acad. Sci. U.S.A. 2002, 99 (17), 1102011024,  DOI: 10.1073/pnas.172226299
  217. 217
    Farrell, I. S.; Toroney, R.; Hazen, J. L.; Mehl, R. A.; Chin, J. W. Photo-Cross-Linking Interacting Proteins with a Genetically Encoded Benzophenone. Nat. Methods 2005, 2 (5), 377384,  DOI: 10.1038/nmeth0505-377
  218. 218
    Chin, J. W.; Schultz, P. G. In Vivo Photocrosslinking with Unnatural Amino Acid Mutagenesis. ChemBioChem 2002, 3 (11), 11351137,  DOI: 10.1002/1439-7633(20021104)3:11<1135::AID-CBIC1135>3.0.CO;2-M
  219. 219
    Mori, H.; Ito, K. Different Modes of SecY-SecA Interactions Revealed by Site-Directed in vivo Photo-Cross-Linking. Proc. Natl. Acad. Sci. U.S.A. 2006, 103 (44), 1615916164,  DOI: 10.1073/pnas.0606390103
  220. 220
    Majmudar, C. Y.; Lee, L. W.; Lancia, J. K.; Nwokoye, A.; Wang, Q.; Wands, A. M.; Wang, L.; Mapp, A. K. Impact of Nonnatural Amino Acid Mutagenesis on the in Vivo Function and Binding Modes of a Transcriptional Activator. J. Am. Chem. Soc. 2009, 131 (40), 1424014242,  DOI: 10.1021/ja904378z
  221. 221
    Liu, C.; Young, A. L.; Starling-Windhof, A.; Bracher, A.; Saschenbrecker, S.; Rao, B. V.; Rao, K. V.; Berninghausen, O.; Mielke, T.; Hartl, F. U. Coupled Chaperone Action in Folding and Assembly of Hexadecameric Rubisco. Nature 2010, 463 (7278), 197202,  DOI: 10.1038/nature08651
  222. 222
    Ye, Z.; Bair, M.; Desai, H.; Williams, G. J. A Photocrosslinking Assay for Reporting Protein Interactions in Polyketide and Fatty Acid Synthases. Mol. BioSyst. 2011, 7 (11), 31523156,  DOI: 10.1039/c1mb05270e
  223. 223
    Ye, Z.; Williams, G. J. Mapping a Ketosynthase:Acyl Carrier Protein Binding Interface via Unnatural Amino Acid-Mediated Photo-Cross-Linking. Biochemistry 2014, 53 (48), 74947502,  DOI: 10.1021/bi500936u
  224. 224
    Pavic, K.; Rios, P.; Dzeyk, K.; Koehler, C.; Lemke, E. A.; Köhn, M. Unnatural Amino Acid Mutagenesis Reveals Dimerization As a Negative Regulatory Mechanism of VHR’s Phosphatase Activity. ACS Chem. Biol. 2014, 9 (7), 14511459,  DOI: 10.1021/cb500240n
  225. 225
    Morozov, Y. I.; Agaronyan, K.; Cheung, A. C. M.; Anikin, M.; Cramer, P.; Temiakov, D. A Novel Intermediate in Transcription Initiation by Human Mitochondrial RNA Polymerase. Nucleic Acids Res. 2014, 42 (6), 38843893,  DOI: 10.1093/nar/gkt1356
  226. 226
    Wong, H. E.; Kwon, I. Effects of Non-Natural Amino Acid Incorporation into the Enzyme Core Region on Enzyme Structure and Function. Int. J. Mol. Sci. 2015, 16 (9), 2273522753,  DOI: 10.3390/ijms160922735
  227. 227
    Rappaport, F.; Boussac, A.; Force, D. A.; Peloquin, J.; Brynda, M.; Sugiura, M.; Un, S.; Britt, R. D.; Diner, B. A. Probing the Coupling between Proton and Electron Transfer in Photosystem II Core Complexes Containing a 3-Fluorotyrosine. J. Am. Chem. Soc. 2009, 131 (12), 44254433,  DOI: 10.1021/ja808604h
  228. 228
    Schlesinger, S.; Schlesinger, M. J. The Effect of Amino Acid Analogues on Alkaline Phosphatase Formation in Escherichia coli K-12: I. Substitution of Triazolealanine for Histidine. J. Biol. Chem. 1967, 242 (14), 33693372,  DOI: 10.1016/S0021-9258(18)95919-3
  229. 229
    Fried, S. D.; Bagchi, S.; Boxer, S. G. Extreme Electric Fields Power Catalysis in the Active Site of Ketosteroid Isomerase. Science 2014, 346 (6216), 15101514,  DOI: 10.1126/science.1259802
  230. 230
    Ortmayer, M.; Hardy, F. J.; Quesne, M. G.; Fisher, K.; Levy, C.; Heyes, D. J.; Catlow, C. R. A.; de Visser, S. P.; Rigby, S. E. J.; Hay, S.; Green, A. P. A Noncanonical Tryptophan Analogue Reveals an Active Site Hydrogen Bond Controlling Ferryl Reactivity in a Heme Peroxidase. JACS Au 2021, 1 (7), 913918,  DOI: 10.1021/jacsau.1c00145
  231. 231
    Kang, G.; Taguchi, A. T.; Stubbe, J.; Drennan, C. L. Structure of a Trapped Radical Transfer Pathway within a Ribonucleotide Reductase Holocomplex. Science 2020, 368 (6489), 424427,  DOI: 10.1126/science.aba6794
  232. 232
    Beiboer, S. H. W.; Berg, B. v. d.; Dekker, N.; Cox, R. C.; Verheij, H. M. Incorporation of an Unnatural Amino Acid in the Active Site of Porcine Pancreatic Phospholipase A2. Substitution of Histidine by l,2,4-Triazole-3-Alanine Yields an Enzyme with High Activity at Acidic pH. Protein Eng. Des. Sel. 1996, 9 (4), 345352,  DOI: 10.1093/protein/9.4.345
  233. 233
    Soumillion, P.; Fastrez, J. Incorporation of 1,2,4-Triazole-3-Alanine into a Mutant of Phage Lambda Lysozyme Containing a Single Histidine. Protein Eng. Des. Sel. 1998, 11 (3), 213217,  DOI: 10.1093/protein/11.3.213
  234. 234
    Blatter, N.; Prokup, A.; Deiters, A.; Marx, A. Modulating the pKa of a Tyrosine in KlenTaq DNA Polymerase that Is Crucial for Abasic Site Bypass by in Vivo Incorporation of a Non-canonical Amino Acid. Chem Bio Chem 2014, 15, 17351737,  DOI: 10.1002/cbic.201400051
  235. 235
    Obeid, S.; Blatter, N.; Kranaster, R.; Schnur, A.; Diederichs, K.; Welte, W.; Marx, A. Replication Through an Abasic DNA Lesion: Structural Basis for Adenine Selectivity. EMBO J. 2010, 29 (10), 1738,  DOI: 10.1038/emboj.2010.64
  236. 236
    Greene, B. L.; Kang, G.; Cui, C.; Bennati, M.; Nocera, D. G.; Drennan, C. L.; Stubbe, J. Ribonucleotide Reductases: Structure, Chemistry, and Metabolism Suggest New Therapeutic Targets. Annu. Rev. Biochem. 2020, 89 (1), 4575,  DOI: 10.1146/annurev-biochem-013118-111843
  237. 237
    Uhlin, U.; Eklund, H. Structure of Ribonucleotide Reductase Protein R1. Nature 1994, 370 (6490), 533539,  DOI: 10.1038/370533a0
  238. 238
    Reece, S. Y.; Seyedsayamdost, M. R.; Stubbe, J.; Nocera, D. G. Electron Transfer Reactions of Fluorotyrosyl Radicals. J. Am. Chem. Soc. 2006, 128 (42), 1365413655,  DOI: 10.1021/ja0636688
  239. 239
    Seyedsayamdost, M. R.; Yee, C. S.; Reece, S. Y.; Nocera, D. G.; Stubbe, J. pH Rate Profiles of FnY356-R2s (n = 2, 3, 4) in Escherichia coli Ribonucleotide Reductase: Evidence that Y356 Is a Redox-Active Amino Acid along the Radical Propagation Pathway. J. Am. Chem. Soc. 2006, 128 (5), 15621568,  DOI: 10.1021/ja055927j
  240. 240
    Seyedsayamdost, M. R.; Xie, J.; Chan, C. T. Y.; Schultz, P. G.; Stubbe, J. Site-Specific Insertion of 3-Aminotyrosine into Subunit α2 of E. coli Ribonucleotide Reductase: Direct Evidence for Involvement of Y730 and Y731 in Radical Propagation. J. Am. Chem. Soc. 2007, 129 (48), 1506015071,  DOI: 10.1021/ja076043y
  241. 241
    Seyedsayamdost, M. R.; Chan, C. T. Y.; Mugnaini, V.; Stubbe, J.; Bennati, M. PELDOR Spectroscopy with DOPA-β2 and NH2Y-α2s: Distance Measurements between Residues Involved in the Radical Propagation Pathway of E. coli Ribonucleotide Reductase. J. Am. Chem. Soc. 2007, 129 (51), 1574815749,  DOI: 10.1021/ja076459b
  242. 242
    Seyedsayamdost, M. R.; Stubbe, J. Site-Specific Replacement of Y356 with 3,4-Dihydroxyphenylalanine in the β2 Subunit of E. coli Ribonucleotide Reductase. J. Am. Chem. Soc. 2006, 128 (8), 25222523,  DOI: 10.1021/ja057776q
  243. 243
    Seyedsayamdost, M. R.; Stubbe, J. Forward and Reverse Electron Transfer with the Y356DOPA-β2 Heterodimer of E. coli Ribonucleotide Reductase. J. Am. Chem. Soc. 2007, 129 (8), 22262227,  DOI: 10.1021/ja0685607
  244. 244
    Yokoyama, K.; Smith, A. A.; Corzilius, B.; Griffin, R. G.; Stubbe, J. Equilibration of Tyrosyl Radicals (Y356•, Y731•, Y730•) in the Radical Propagation Pathway of the Escherichia coli Class Ia Ribonucleotide Reductase. J. Am. Chem. Soc. 2011, 133 (45), 1842018432,  DOI: 10.1021/ja207455k
  245. 245
    Lin, C.-Y.; Muñoz Hernández, A. L.; Laremore, T. N.; Silakov, A.; Krebs, C.; Boal, A. K.; Bollinger, J. M., Jr Use of Noncanonical Tyrosine Analogues to Probe Control of Radical Intermediates during Endoperoxide Installation by Verruculogen Synthase (FtmOx1). ACS Catal. 2022, 12 (12), 69686979,  DOI: 10.1021/acscatal.2c01037
  246. 246
    Taylor, A.; Heyes, D. J.; Scrutton, N. S. Catalysis by Nature’s photoenzymes. Curr. Opin. Struct. Biol. 2022, 77, 102491,  DOI: 10.1016/j.sbi.2022.102491
  247. 247
    Taylor, A.; Zhang, S.; Johannissen, L. O.; Sakuma, M.; Phillips, R. S.; Green, A. P.; Hay, S.; Heyes, D. J.; Scrutton, N. S. Mechanistic Implications of the Ternary Complex Structural Models for the Photoenzyme Protochlorophyllide Oxidoreductase. FEBS J. 2024, 291, 1404,  DOI: 10.1111/febs.17025
  248. 248
    Horowitz, S.; Adhikari, U.; Dirk, L. M. A.; Del Rizzo, P. A.; Mehl, R. A.; Houtz, R. L.; Al-Hashimi, H. M.; Scheiner, S.; Trievel, R. C. Manipulating Unconventional CH-Based Hydrogen Bonding in a Methyltransferase via Noncanonical Amino Acid Mutagenesis. ACS Chem. Biol. 2014, 9 (8), 16921697,  DOI: 10.1021/cb5001185
  249. 249
    Kirsh, J. M.; Weaver, J. B.; Boxer, S. G.; Kozuch, J. Comprehensive Analysis of Nitrile Probe IR Shifts and Intensities in Proteins: Experiment and Critical Evaluation of Simulations. ChemRxiv 2024,  DOI: 10.26434/chemrxiv-2023-j935v-v2
  250. 250
    Weaver, J. B.; Kozuch, J.; Kirsh, J. M.; Boxer, S. G. Nitrile Infrared Intensities Characterize Electric Fields and Hydrogen Bonding in Protic, Aprotic, and Protein Environments. J. Am. Chem. Soc. 2022, 144 (17), 75627567,  DOI: 10.1021/jacs.2c00675
  251. 251
    Zheng, C.; Mao, Y.; Kozuch, J.; Atsango, A. O.; Ji, Z.; Markland, T. E.; Boxer, S. G. A Two-Directional Vibrational Probe Reveals Different Electric Field Orientations in Solution and an Enzyme Active Site. Nat. Chem. 2022, 14 (8), 891897,  DOI: 10.1038/s41557-022-00937-w
  252. 252
    Kedzierski, P.; Zaczkowska, M.; Sokalski, W. A. Extreme Catalytic Power of Ketosteroid Isomerase Related to the Reversal of Proton Dislocations in Hydrogen-Bond Network. J. Phys. Chem. B 2020, 124 (18), 36613666,  DOI: 10.1021/acs.jpcb.0c01489
  253. 253
    Pollack, R. M. Enzymatic Mechanisms for Catalysis of Enolization: Ketosteroid Isomerase. Bioorg. Chem. 2004, 32 (5), 341353,  DOI: 10.1016/j.bioorg.2004.06.005
  254. 254
    Wu, Y.; Boxer, S. G. A Critical Test of the Electrostatic Contribution to Catalysis with Noncanonical Amino Acids in Ketosteroid Isomerase. J. Am. Chem. Soc. 2016, 138 (36), 1189011895,  DOI: 10.1021/jacs.6b06843
  255. 255
    Faraldos, J. A.; Antonczak, A. K.; González, V.; Fullerton, R.; Tippmann, E. M.; Allemann, R. K. Probing Eudesmane Cation-π Interactions in Catalysis by Aristolochene Synthase with Non-canonical Amino Acids. J. Am. Chem. Soc. 2011, 133 (35), 1390613909,  DOI: 10.1021/ja205927u
  256. 256
    Morikubo, N.; Fukuda, Y.; Ohtake, K.; Shinya, N.; Kiga, D.; Sakamoto, K.; Asanuma, M.; Hirota, H.; Yokoyama, S.; Hoshino, T. Cation-π Interaction in the Polyolefin Cyclization Cascade Uncovered by Incorporating Unnatural Amino Acids into the Catalytic Sites of Squalene Cyclase. J. Am. Chem. Soc. 2006, 128 (40), 1318413194,  DOI: 10.1021/ja063358p
  257. 257
    Herbst, R. W.; Guce, A.; Bryngelson, P. A.; Higgins, K. A.; Ryan, K. C.; Cabelli, D. E.; Garman, S. C.; Maroney, M. J. Role of Conserved Tyrosine Residues in NiSOD Catalysis: A Case of Convergent Evolution. Biochemistry 2009, 48 (15), 33543369,  DOI: 10.1021/bi802029t
  258. 258
    Campeciño, J. O.; Dudycz, L. W.; Tumelty, D.; Berg, V.; Cabelli, D. E.; Maroney, M. J. A Semisynthetic Strategy Leads to Alteration of the Backbone Amidate Ligand in the NiSOD Active Site. J. Am. Chem. Soc. 2015, 137 (28), 90449052,  DOI: 10.1021/jacs.5b03629
  259. 259
    Evans, R. M.; Krahn, N.; Murphy, B. J.; Lee, H.; Armstrong, F. A.; Söll, D. Selective Cysteine-To-Selenocysteine Changes in a [NiFe]-Hydrogenase Confirm a Special Position for Catalysis and Oxygen Tolerance. Proc. Natl. Acad. Sci. U.S.A. 2021, 118 (13), e2100921118  DOI: 10.1073/pnas.2100921118
  260. 260
    Mukai, T.; Sevostyanova, A.; Suzuki, T.; Fu, X.; Söll, D. A Facile Method for Producing Selenocysteine-Containing Proteins. Angew. Chem. Int. Ed. 2018, 57 (24), 72157219,  DOI: 10.1002/anie.201713215
  261. 261
    Snider, G. W.; Ruggles, E.; Khan, N.; Hondal, R. J. Selenocysteine Confers Resistance to Inactivation by Oxidation in Thioredoxin Reductase: Comparison of Selenium and Sulfur Enzymes. Biochemistry 2013, 52 (32), 54725481,  DOI: 10.1021/bi400462j
  262. 262
    Wu, Z. P.; Hilvert, D. Selenosubtilisin as a Glutathione Peroxidase Mimic. J. Am. Chem. Soc. 1990, 112 (14), 56475648,  DOI: 10.1021/ja00170a043
  263. 263
    Hardy, F. J.; Ortmayer, M.; Green, A. P.; Noble, C. E. M.; Anderson, J. L. R. Recent Advances in Understanding, Enhancing and Creating Heme Peroxidases. In Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier, 2020. DOI: 10.1016/B978-0-08-102688-5.00021-0
  264. 264
    Poulos, T. L. Heme Enzyme Structure and Function. Chem. Rev. 2014, 114, 39193962,  DOI: 10.1021/cr400415k
  265. 265
    Behan, R. K.; Green, M. T. On the Status of Ferryl Protonation. J. Inorg. Biochem. 2006, 100 (4), 448459,  DOI: 10.1016/j.jinorgbio.2005.12.019
  266. 266
    Sivaramakrishnan, S.; Ouellet, H.; Du, J.; McLean, K. J.; Medzihradszky, K. F.; Dawson, J. H.; Munro, A. W.; Ortiz de Montellano, P. R. A Novel Intermediate in the Reaction of Seleno CYP119 with m-Chloroperbenzoic Acid. Biochemistry 2011, 50 (14), 30143024,  DOI: 10.1021/bi101728y
  267. 267
    Aldag, C.; Gromov, I. A.; García-Rubio, I.; Von Koenig, K.; Schlichting, I.; Jaun, B.; Hilvert, D. Probing the Role of the Proximal Heme Ligand in Cytochrome P450cam by Recombinant Incorporation of Selenocysteine. Proc. Natl. Acad. Sci. U.S.A. 2009, 106 (14), 54815486,  DOI: 10.1073/pnas.0810503106
  268. 268
    Jiang, Y.; Sivaramakrishnan, S.; Hayashi, T.; Cohen, S.; Moënne-Loccoz, P.; Shaik, S.; Ortiz de Montellano, P. R. Calculated and Experimental Spin State of Seleno Cytochrome P450. Angew. Chem. Int. Ed. 2009, 48 (39), 71937195,  DOI: 10.1002/anie.200901485
  269. 269
    Onderko, E. L.; Silakov, A.; Yosca, T. H.; Green, M. T. Characterization of a Selenocysteine-Ligated P450 Compound I Reveals Direct Link Between Electron Donation and Reactivity. Nat. Chem. 2017, 9 (7), 623628,  DOI: 10.1038/nchem.2781
  270. 270
    Sivaramakrishnan, S.; Ouellet, H.; Matsumura, H.; Guan, S.; Moënne-Loccoz, P.; Burlingame, A. L.; Ortiz De Montellano, P. R. Proximal Ligand Electron Donation and Reactivity of the Cytochrome P450 Ferric-Peroxo Anion. J. Am. Chem. Soc. 2012, 134 (15), 66736684,  DOI: 10.1021/ja211499q
  271. 271
    Ortmayer, M.; Fisher, K.; Basran, J.; Wolde-Michael, E. M.; Heyes, D. J.; Levy, C.; Lovelock, S. L.; Anderson, J. L. R.; Raven, E. L.; Hay, S. Rewiring the “ Push-Pull ” Catalytic Machinery of a Heme Enzyme Using an Expanded Genetic Code. ACS Catal. 2020, 10, 27352746,  DOI: 10.1021/acscatal.9b05129
  272. 272
    Xiao, H.; Peters, F. B.; Yang, P.-Y.; Reed, S.; Chittuluru, J. R.; Schultz, P. G. Genetic Incorporation of Histidine Derivatives Using an Engineered Pyrrolysyl-tRNA Synthetase. ACS Chem. Biol. 2014, 9 (5), 10921096,  DOI: 10.1021/cb500032c
  273. 273
    Martin, C.; Zhang, Y. The Diverse Functions of Histone Lysine Methylation. Nat. Rev. Mol. Cell Biol. 2005, 6 (11), 838849,  DOI: 10.1038/nrm1761
  274. 274
    Niu, W.; Guo, J. Co-translational Installation of Posttranslational Modifications by Non-canonical Amino Acid Mutagenesis. ChemBioChem 2023, 24 (9), e202300039  DOI: 10.1002/cbic.202300039
  275. 275
    Rust, H. L.; Subramanian, V.; West, G. M.; Young, D. D.; Schultz, P. G.; Thompson, P. R. Using Unnatural Amino Acid Mutagenesis To Probe the Regulation of PRMT1. ACS Chem. Biol. 2014, 9 (3), 649655,  DOI: 10.1021/cb400859z
  276. 276
    Neumann, H.; Neumann-Staubitz, P.; Witte, A.; Summerer, D. Epigenetic Chromatin Modification by Amber Suppression Technology. Curr. Opin. Chem. Biol. 2018, 45, 19,  DOI: 10.1016/j.cbpa.2018.01.017
  277. 277
    Wang, Z. A.; Cole, P. A. The Chemical Biology of Reversible Lysine Post-translational Modifications. Cell Chem. Biol. 2020, 27 (8), 953969,  DOI: 10.1016/j.chembiol.2020.07.002
  278. 278
    Wang, T.; Zhou, Q.; Li, F.; Yu, Y.; Yin, X.; Wang, J. Genetic Incorporation of Nε-Formyllysine, a New Histone Post-translational Modification. ChemBioChem 2015, 16 (10), 14401442,  DOI: 10.1002/cbic.201500170
  279. 279
    Cao, L.; Liu, J.; Ghelichkhani, F.; Rozovsky, S.; Wang, L. Genetic Incorporation of ε-N-Benzoyllysine by Engineering Methanomethylophilus alvus Pyrrolysyl-tRNA Synthetase. ChemBioChem 2021, 22 (15), 25302534,  DOI: 10.1002/cbic.202100218
  280. 280
    Ren, C.; Wu, Q.; Xiao, R.; Ji, Y.; Yang, X.; Zhang, Z.; Qin, H.; Ma, J.-A.; Xuan, W. Expanding the Scope of Genetically Encoded Lysine Post-Translational Modifications with Lactylation, β-Hydroxybutyrylation and Lipoylation. ChemBioChem 2022, 23 (18), e202200302  DOI: 10.1002/cbic.202200302
  281. 281
    Nguyen, D. P.; Garcia Alai, M. M.; Kapadnis, P. B.; Neumann, H.; Chin, J. W. Genetically Encoding Nϵ-Methyl-l-lysine in Recombinant Histones. J. Am. Chem. Soc. 2009, 131 (40), 1419414195,  DOI: 10.1021/ja906603s
  282. 282
    Wang, Z. A.; Liu, W. R. Proteins with Site-Specific Lysine Methylation. Chem. Eur. J. 2017, 23 (49), 1173211737,  DOI: 10.1002/chem.201701655
  283. 283
    Wang, Y.-S.; Wu, B.; Wang, Z.; Huang, Y.; Wan, W.; Russell, W. K.; Pai, P.-J.; Moe, Y. N.; Russell, D. H.; Liu, W. R. A Genetically Encoded Photocaged Nε-Methyl-L-Lysine. Mol. BioSyst. 2010, 6 (9), 15571560,  DOI: 10.1039/c002155e
  284. 284
    Neumann, H.; Peak-Chew, S. Y.; Chin, J. W. Genetically Encoding Nε-Acetyllysine in Recombinant Proteins. Nat. Chem. Biol. 2008, 4 (4), 232234,  DOI: 10.1038/nchembio.73
  285. 285
    Gattner, M. J.; Vrabel, M.; Carell, T. Synthesis of ε-N-Propionyl-, ε-N-Butyryl-, and ε-N-Crotonyl-Lysine Containing Histone H3 Using the Pyrrolysine System. Chem. Commun. 2013, 49 (4), 379381,  DOI: 10.1039/C2CC37836A
  286. 286
    Wilkins, B. J.; Hahn, L. E.; Heitmüller, S.; Frauendorf, H.; Valerius, O.; Braus, G. H.; Neumann, H. Genetically Encoding Lysine Modifications on Histone H4. ACS Chem. Biol. 2015, 10 (4), 939944,  DOI: 10.1021/cb501011v
  287. 287
    Xiao, H.; Xuan, W.; Shao, S.; Liu, T.; Schultz, P. G. Genetic Incorporation of ε-N-2-Hydroxyisobutyryl-Lysine into Recombinant Histones. ACS Chem. Biol. 2015, 10 (7), 15991603,  DOI: 10.1021/cb501055h
  288. 288
    Kim, C. H.; Kang, M.; Kim, H. J.; Chatterjee, A.; Schultz, P. G. Site-Specific Incorporation of ε-N-Crotonyllysine into Histones. Angew. Chem. Int. Ed. 2012, 51 (29), 72467249,  DOI: 10.1002/anie.201203349
  289. 289
    Tian, H.; Yang, J.; Guo, A.-D.; Ran, Y.; Yang, Y.-Z.; Yang, B.; Huang, R.; Liu, H.; Chen, X.-H. Genetically Encoded Benzoyllysines Serve as Versatile Probes for Interrogating Histone Benzoylation and Interactions in Living Cells. ACS Chem. Biol. 2021, 16 (11), 25602569,  DOI: 10.1021/acschembio.1c00614
  290. 290
    Fatema, N.; Fan, C. Studying Lysine Acetylation of Citric Acid Cycle Enzymes by Genetic Code Expansion. Mol. Microbiol. 2023, 119 (5), 551559,  DOI: 10.1111/mmi.15052
  291. 291
    Araujo, J.; Ottinger, S.; Venkat, S.; Gan, Q.; Fan, C. Studying Acetylation of Aconitase Isozymes by Genetic Code Expansion. Front. Chem. 2022, 10, 1,  DOI: 10.3389/fchem.2022.862483
  292. 292
    Venkat, S.; Chen, H.; Stahman, A.; Hudson, D.; McGuire, P.; Gan, Q.; Fan, C. Characterizing Lysine Acetylation of Isocitrate Dehydrogenase in Escherichia coli. J. Mol. Biol. 2018, 430 (13), 19011911,  DOI: 10.1016/j.jmb.2018.04.031
  293. 293
    Venkat, S.; Gregory, C.; Sturges, J.; Gan, Q.; Fan, C. Studying the Lysine Acetylation of Malate Dehydrogenase. J. Mol. Biol. 2017, 429 (9), 13961405,  DOI: 10.1016/j.jmb.2017.03.027
  294. 294
    Venkat, S.; Chen, H.; McGuire, P.; Stahman, A.; Gan, Q.; Fan, C. Characterizing Lysine Acetylation of Escherichia coli Type II Citrate Synthase. FEBS J. 2019, 286 (14), 27992808,  DOI: 10.1111/febs.14845
  295. 295
    Wright, D. E.; Altaany, Z.; Bi, Y.; Alperstein, Z.; O’Donoghue, P. Acetylation Regulates Thioredoxin Reductase Oligomerization and Activity. Antioxid. Redox Signal. 2018, 29 (4), 377388,  DOI: 10.1089/ars.2017.7082
  296. 296
    Rogerson, D. T.; Sachdeva, A.; Wang, K.; Haq, T.; Kazlauskaite, A.; Hancock, S. M.; Huguenin-Dezot, N.; Muqit, M. M. K.; Fry, A. M.; Bayliss, R. Efficient Genetic Encoding of Phosphoserine and its Nonhydrolyzable Analog. Nat. Chem. Biol. 2015, 11 (7), 496503,  DOI: 10.1038/nchembio.1823
  297. 297
    Venkat, S.; Sturges, J.; Stahman, A.; Gregory, C.; Gan, Q.; Fan, C. Genetically Incorporating Two Distinct Post-translational Modifications into One Protein Simultaneously. ACS Synth. Biol. 2018, 7 (2), 689695,  DOI: 10.1021/acssynbio.7b00408
  298. 298
    Zang, J.; Chen, Y.; Liu, C.; Lin, S. Probing the Role of Aurora Kinase A Threonylation with Site-Specific Lysine Threonylation. ACS Chem. Biol. 2023, 18 (4), 674678,  DOI: 10.1021/acschembio.1c00682
  299. 299
    Wan, N.; Wang, N.; Yu, S.; Zhang, H.; Tang, S.; Wang, D.; Lu, W.; Li, H.; Delafield, D. G.; Kong, Y. Cyclic Immonium Ion of Lactyllysine Reveals Widespread Lactylation in the Human Proteome. Nat. Methods 2022, 19 (7), 854864,  DOI: 10.1038/s41592-022-01523-1
  300. 300
    Whittaker, J. W. Free Radical Catalysis by Galactose Oxidase. Chem. Rev. 2003, 103 (6), 23472364,  DOI: 10.1021/cr020425z
  301. 301
    Polyakov, K. M.; Boyko, K. M.; Tikhonova, T. V.; Slutsky, A.; Antipov, A. N.; Zvyagilskaya, R. A.; Popov, A. N.; Bourenkov, G. P.; Lamzin, V. S.; Popov, V. O. High-Resolution Structural Analysis of a Novel Octaheme Cytochrome c Nitrite Reductase from the Haloalkaliphilic Bacterium Thioalkalivibrio nitratireducens. J. Mol. Biol. 2009, 389 (5), 846862,  DOI: 10.1016/j.jmb.2009.04.037
  302. 302
    Ye, S.; Wu, X. a.; Wei, L.; Tang, D.; Sun, P.; Bartlam, M.; Rao, Z. An Insight into the Mechanism of Human Cysteine Dioxygenase: key roles of the thioether-bonded tyrosine-cysteine cofactor. J. Biol. Chem. 2007, 282 (5), 33913402,  DOI: 10.1074/jbc.M609337200
  303. 303
    Zhou, Q.; Hu, M.; Zhang, W.; Jiang, L.; Perrett, S.; Zhou, J.; Wang, J. Probing the Function of the Tyr-Cys Cross-Link in Metalloenzymes by the Genetic Incorporation of 3-Methylthiotyrosine. Angew. Chem. Int. Ed. 2013, 52 (4), 12031207,  DOI: 10.1002/anie.201207229
  304. 304
    Dominy, J. E.; Hwang, J.; Guo, S.; Hirschberger, L. L.; Zhang, S.; Stipanuk, M. H. Synthesis of Amino Acid Cofactor in Cysteine Dioxygenase Is Regulated by Substrate and Represents a Novel Post-translational Regulation of Activity. J. Biol. Chem. 2008, 283 (18), 1218812201,  DOI: 10.1074/jbc.M800044200
  305. 305
    Li, J.; Griffith, W. P.; Davis, I.; Shin, I.; Wang, J.; Li, F.; Wang, Y.; Wherritt, D. J.; Liu, A. Cleavage of a Carbon-Fluorine Bond by an Engineered Cysteine Dioxygenase. Nat. Chem. Biol. 2018, 14 (9), 853860,  DOI: 10.1038/s41589-018-0085-5
  306. 306
    Li, J.; Koto, T.; Davis, I.; Liu, A. Probing the Cys-Tyr Cofactor Biogenesis in Cysteine Dioxygenase by the Genetic Incorporation of Fluorotyrosine. Biochemistry 2019, 58 (17), 22182227,  DOI: 10.1021/acs.biochem.9b00006
  307. 307
    Chen, L.; Naowarojna, N.; Song, H.; Wang, S.; Wang, J.; Deng, Z.; Zhao, C.; Liu, P. Use of a Tyrosine Analogue To Modulate the Two Activities of a Nonheme Iron Enzyme OvoA in Ovothiol Biosynthesis, Cysteine Oxidation versus Oxidative C-S Bond Formation. J. Am. Chem. Soc. 2018, 140 (13), 46044612,  DOI: 10.1021/jacs.7b13628
  308. 308
    Trumpower, B. L.; Gennis, R. B. Energy Transduction By Cytochrome Complexes In Mitochondrial And Bacterial Respiration: The Enzymology of Coupling Electron Transfer Reactions to Transmembrane Proton Translocation. Annu. Rev. Biochem. 1994, 63 (1), 675716,  DOI: 10.1146/annurev.bi.63.070194.003331
  309. 309
    Kim, E.; Chufán, E. E.; Kamaraj, K.; Karlin, K. D. Synthetic Models for Heme-Copper Oxidases. Chem. Rev. 2004, 104 (2), 10771134,  DOI: 10.1021/cr0206162
  310. 310
    Miner, K. D.; Mukherjee, A.; Gao, Y.-G.; Null, E. L.; Petrik, I. D.; Zhao, X.; Yeung, N.; Robinson, H.; Lu, Y. A Designed Functional Metalloenzyme that Reduces O2 to H2O with Over One Thousand Turnovers. Angew. Chem. Int. Ed. 2012, 51 (23), 55895592,  DOI: 10.1002/anie.201201981
  311. 311
    Liu, X.; Yu, Y.; Hu, C.; Zhang, W.; Lu, Y.; Wang, J. Significant Increase of Oxidase Activity through the Genetic Incorporation of a Tyrosine-Histidine Cross-Link in a Myoglobin Model of Heme-Copper Oxidase. Angew. Chem. Int. Ed. 2012, 51 (18), 43124316,  DOI: 10.1002/anie.201108756
  312. 312
    Yu, Y.; Lv, X.; Li, J.; Zhou, Q.; Cui, C.; Hosseinzadeh, P.; Mukherjee, A.; Nilges, M. J.; Wang, J.; Lu, Y. Defining the Role of Tyrosine and Rational Tuning of Oxidase Activity by Genetic Incorporation of Unnatural Tyrosine Analogs. J. Am. Chem. Soc. 2015, 137 (14), 45944597,  DOI: 10.1021/ja5109936
  313. 313
    Yu, Y.; Zhou, Q.; Wang, L.; Liu, X.; Zhang, W.; Hu, M.; Dong, J.; Li, J.; Lv, X.; Ouyang, H. Significant Improvement of Oxidase Activity Through the Genetic Incorporation of a Redox-Active Unnatural Amino Acid. Chem. Sci. 2015, 6 (7), 38813885,  DOI: 10.1039/C5SC01126D
  314. 314
    Rigoldi, F.; Donini, S.; Redaelli, A.; Parisini, E.; Gautieri, A. Review: Engineering of Thermostable Enzymes for Industrial Applications. APL Bioengineering 2018, 2 (1), 011501,  DOI: 10.1063/1.4997367
  315. 315
    Sheldon, R. A.; Basso, A.; Brady, D. New Frontiers in Enzyme Immobilisation: Robust Biocatalysts for a Circular Bio-Based Economy. Chem. Soc. Rev. 2021, 50 (10), 58505862,  DOI: 10.1039/D1CS00015B
  316. 316
    Baker, P. J.; Montclare, J. K. Enhanced Refoldability and Thermoactivity of Fluorinated Phosphotriesterase. ChemBioChem 2011, 12 (12), 18451848,  DOI: 10.1002/cbic.201100221
  317. 317
    Deepankumar, K.; Shon, M.; Nadarajan, S. P.; Shin, G.; Mathew, S.; Ayyadurai, N.; Kim, B.-G.; Choi, S.-H.; Lee, S.-H.; Yun, H. Enhancing Thermostability and Organic Solvent Tolerance of ω-Transaminase through Global Incorporation of Fluorotyrosine. Adv. Synth. Catal. 2014, 356 (5), 993998,  DOI: 10.1002/adsc.201300706
  318. 318
    Ohtake, K.; Mukai, T.; Iraha, F.; Takahashi, M.; Haruna, K.-i.; Date, M.; Yokoyama, K.; Sakamoto, K. Engineering an Automaturing Transglutaminase with Enhanced Thermostability by Genetic Code Expansion with Two Codon Reassignments. ACS Synth. Biol. 2018, 7 (9), 21702176,  DOI: 10.1021/acssynbio.8b00157
  319. 319
    Politzer, P.; Murray, J. S. Halogen Bonding: An Interim Discussion. ChemPhysChem 2013, 14 (2), 278294,  DOI: 10.1002/cphc.201200799
  320. 320
    Scholfield, M. R.; Ford, M. C.; Carlsson, A.-C. C.; Butta, H.; Mehl, R. A.; Ho, P. S. Structure-Energy Relationships of Halogen Bonds in Proteins. Biochemistry 2017, 56 (22), 27942802,  DOI: 10.1021/acs.biochem.7b00022
  321. 321
    Acevedo-Rocha, C. G.; Hoesl, M. G.; Nehring, S.; Royter, M.; Wolschner, C.; Wiltschi, B.; Antranikian, G.; Budisa, N. Non-Canonical Amino Acids as a Useful Synthetic Biological Tool for Lipase-Catalysed Reactions in Hostile Environments. Catal. Sci. Technol. 2013, 3 (5), 11981201,  DOI: 10.1039/c3cy20712a
  322. 322
    Mendel, D.; Ellman, J. A.; Chang, Z.; Veenstra, D. L.; Kollman, P. A.; Schultz, P. G. Probing Protein Stability with Unnatural Amino Acids. Science 1992, 256 (5065), 17981802,  DOI: 10.1126/science.1615324
  323. 323
    Ismail, A. R.; Kashtoh, H.; Baek, K.-H. Temperature-Resistant and Solvent-Tolerant Lipases as Industrial Biocatalysts: Biotechnological Approaches and Applications. Int. J. Biol. Macromol. 2021, 187, 127142,  DOI: 10.1016/j.ijbiomac.2021.07.101
  324. 324
    Vivek, K.; Sandhia, G. S.; Subramaniyan, S. Extremophilic Lipases for Industrial Applications: A General Review. Biotechnol. Adv. 2022, 60, 108002,  DOI: 10.1016/j.biotechadv.2022.108002
  325. 325
    Budisa, N.; Wenger, W.; Wiltschi, B. Residue-Specific Global Fluorination of Candida antarctica Lipase B in Pichia pastoris. Mol. BioSyst. 2010, 6 (9), 16301639,  DOI: 10.1039/c002256j
  326. 326
    Merkel, L.; Schauer, M.; Antranikian, G.; Budisa, N. Parallel Incorporation of Different Fluorinated Amino Acids: On the Way to “Teflon” Proteins. ChemBioChem 2010, 11 (11), 15051507,  DOI: 10.1002/cbic.201000295
  327. 327
    Hoesl, M. G.; Acevedo-Rocha, C. G.; Nehring, S.; Royter, M.; Wolschner, C.; Wiltschi, B.; Budisa, N.; Antranikian, G. Lipase Congeners Designed by Genetic Code Engineering. ChemCatChem 2011, 3 (1), 213221,  DOI: 10.1002/cctc.201000253
  328. 328
    Kelly, S. A.; Pohle, S.; Wharry, S.; Mix, S.; Allen, C. C. R.; Moody, T. S.; Gilmore, B. F. Application of ω-Transaminases in the Pharmaceutical Industry. Chem. Rev. 2018, 118 (1), 349367,  DOI: 10.1021/acs.chemrev.7b00437
  329. 329
    Mathew, S.; Yun, H. ω-Transaminases for the Production of Optically Pure Amines and Unnatural Amino Acids. ACS Catal. 2012, 2 (6), 9931001,  DOI: 10.1021/cs300116n
  330. 330
    Won, Y.; Jeon, H.; Pagar, A. D.; Patil, M. D.; Nadarajan, S. P.; Flood, D. T.; Dawson, P. E.; Yun, H. In vivo Biosynthesis of Tyrosine Analogs and their Concurrent Incorporation into a Residue-Specific Manner for Enzyme Engineering. Chem. Commun. 2019, 55 (100), 1513315136,  DOI: 10.1039/C9CC08503C
  331. 331
    Votchitseva, Y. A.; Efremenko, E. N.; Varfolomeyev, S. D. Insertion of an Unnatural Amino Acid into the Protein Structure: Preparation and Properties of 3-Fluorotyrosine-Containing Organophosphate Hydrolase. Russ. Chem. Bull. 2006, 55 (2), 369374,  DOI: 10.1007/s11172-006-0262-7
  332. 332
    Castro, A. A. d.; Prandi, I. G.; Kuca, K.; Ramalho, T. C. Enzimas Degradantes de Organofosforados: Base Molecular e Perspectivas para Biorremediação Enzimática de Agroquímicos. Ciênc. Agrotec. 2017, 41 (5), 471,  DOI: 10.1590/1413-70542017415000417
  333. 333
    Makkar, R. S.; DiNovo, A. A.; Westwater, C.; Schofield, D. A. Enzyme-Mediated Bioremediation of Organophosphates using Stable Yeast Biocatalysts. J. Bioremed. Biodeg. 2013, 4 (182), 2,  DOI: 10.4172/2155-6199.1000182
  334. 334
    Holzberger, B.; Marx, A. Replacing 32 Proline Residues by a Noncanonical Amino Acid Results in a Highly Active DNA Polymerase. J. Am. Chem. Soc. 2010, 132 (44), 1570815713,  DOI: 10.1021/ja106525y
  335. 335
    Holzberger, B.; Obeid, S.; Welte, W.; Diederichs, K.; Marx, A. Structural Insights into the Potential of 4-Fluoroproline to Modulate Biophysical Properties of Proteins. Chem. Sci. 2012, 3 (10), 29242931,  DOI: 10.1039/c2sc20545a
  336. 336
    Panchenko, T.; Zhu, W. W.; Montclare, J. K. Influence of Global Fluorination on Chloramphenicol Acetyltransferase Activity and Stability. Biotechnol. Bioeng. 2006, 94 (5), 921930,  DOI: 10.1002/bit.20910
  337. 337
    Yang, C.-Y.; Renfrew, P. D.; Olsen, A. J.; Zhang, M.; Yuvienco, C.; Bonneau, R.; Montclare, J. K. Improved Stability and Half-Life of Fluorinated Phosphotriesterase Using Rosetta. ChemBioChem 2014, 15 (12), 17611764,  DOI: 10.1002/cbic.201402062
  338. 338
    Ohtake, K.; Yamaguchi, A.; Mukai, T.; Kashimura, H.; Hirano, N.; Haruki, M.; Kohashi, S.; Yamagishi, K.; Murayama, K.; Tomabechi, Y. Protein Stabilization Utilizing a Redefined Codon. Sci. Rep. 2015, 5 (1), 9762,  DOI: 10.1038/srep09762
  339. 339
    Carlsson, A. C.; Scholfield, M. R.; Rowe, R. K.; Ford, M. C.; Alexander, A. T.; Mehl, R. A.; Ho, P. S. Increasing Enzyme Stability and Activity through Hydrogen Bond-Enhanced Halogen Bonds. Biochemistry 2018, 57 (28), 41354147,  DOI: 10.1021/acs.biochem.8b00603
  340. 340
    Nicholson, H.; Anderson, D. E.; Dao Pin, S.; Matthews, B. W. Analysis of the Interaction Between Charged Side Chains and the Alpha-Helix Dipole Using Designed Thermostable Mutants of Phage T4 Lysozyme. Biochemistry 1991, 30 (41), 98169828,  DOI: 10.1021/bi00105a002
  341. 341
    Klink, T. A.; Woycechowsky, K. J.; Taylor, K. M.; Raines, R. T. Contribution of Disulfide Bonds to the Conformational Stability and Catalytic Activity Of Ribonuclease A. Eur. J. Biochem. 2000, 267 (2), 566572,  DOI: 10.1046/j.1432-1327.2000.01037.x
  342. 342
    Yin, X.; Hu, D.; Li, J.-F.; He, Y.; Zhu, T.-D.; Wu, M.-C. Contribution of Disulfide Bridges to the Thermostability of a Type A Feruloyl Esterase from Aspergillus usamii. PLOS ONE 2015, 10 (5), e0126864  DOI: 10.1371/journal.pone.0126864
  343. 343
    Gihaz, S.; Bash, Y.; Rush, I.; Shahar, A.; Pazy, Y.; Fishman, A. Bridges to Stability: Engineering Disulfide Bonds Towards Enhanced Lipase Biodiesel Synthesis. ChemCatChem 2020, 12 (1), 181192,  DOI: 10.1002/cctc.201901369
  344. 344
    Zhou, X.; Xu, Z.; Li, Y.; He, J.; Zhu, H. Improvement of the Stability and Activity of an LPMO Through Rational Disulfide Bonds Design. Front. bioeng. biotechnol. 2022,  DOI: 10.3389/fbioe.2021.815990
  345. 345
    Dombkowski, A. A.; Sultana, K. Z.; Craig, D. B. Protein Disulfide Engineering. FEBS Lett. 2014, 588 (2), 206212,  DOI: 10.1016/j.febslet.2013.11.024
  346. 346
    Liu, T.; Wang, Y.; Luo, X.; Li, J.; Reed, S. A.; Xiao, H.; Young, T. S.; Schultz, P. G. Enhancing Protein Stability with Extended Disulfide Bonds. Proc. Natl. Acad. Sci. U.S.A. 2016, 113 (21), 59105915,  DOI: 10.1073/pnas.1605363113
  347. 347
    Hecky, J.; Müller, K. M. Structural Perturbation and Compensation by Directed Evolution at Physiological Temperature Leads to Thermostabilization of β-Lactamase. Biochemistry 2005, 44 (38), 1264012654,  DOI: 10.1021/bi0501885
  348. 348
    Brown, N. G.; Pennington, J. M.; Huang, W.; Ayvaz, T.; Palzkill, T. Multiple Global Suppressors of Protein Stability Defects Facilitate the Evolution of Extended-Spectrum TEM β-Lactamases. J. Mol. Biol. 2010, 404 (5), 832846,  DOI: 10.1016/j.jmb.2010.10.008
  349. 349
    Moore, E. J.; Zorine, D.; Hansen, W. A.; Khare, S. D.; Fasan, R. Enzyme Stabilization via Computationally Guided Protein Stapling. Proc. Natl. Acad. Sci. U.S.A. 2017, 114 (47), 1247212477,  DOI: 10.1073/pnas.1708907114
  350. 350
    Bi, J.; Jing, X.; Wu, L.; Zhou, X.; Gu, J.; Nie, Y.; Xu, Y. Computational Design of Noncanonical Amino Acid-Based Thioether Staples at N/C-Terminal Domains of Multi-Modular Pullulanase for Thermostabilization in Enzyme Catalysis. Comput. Struct. Biotechnol. J. 2021, 19, 577585,  DOI: 10.1016/j.csbj.2020.12.043
  351. 351
    Iannuzzelli, J. A.; Bacik, J.-P.; Moore, E. J.; Shen, Z.; Irving, E. M.; Vargas, D. A.; Khare, S. D.; Ando, N.; Fasan, R. Tuning Enzyme Thermostability via Computationally Guided Covalent Stapling and Structural Basis of Enhanced Stabilization. Biochemistry 2022, 61 (11), 10411054,  DOI: 10.1021/acs.biochem.2c00033
  352. 352
    Gur, E.; Biran, D.; Gazit, E.; Ron, E. Z. In vivo Aggregation of a Single Enzyme Limits Growth of Escherichia coli at Elevated Temperatures. Mol. Microbiol. 2002, 46 (5), 13911397,  DOI: 10.1046/j.1365-2958.2002.03257.x
  353. 353
    Li, J. C.; Liu, T.; Wang, Y.; Mehta, A. P.; Schultz, P. G. Enhancing Protein Stability with Genetically Encoded Noncanonical Amino Acids. J. Am. Chem. Soc. 2018, 140 (47), 1599716000,  DOI: 10.1021/jacs.8b07157
  354. 354
    Xuan, W.; Li, J.; Luo, X.; Schultz, P. G. Genetic Incorporation of a Reactive Isothiocyanate Group into Proteins. Angew. Chem. Int. Ed. 2016, 55 (34), 1006510068,  DOI: 10.1002/anie.201604891
  355. 355
    Li, J. C.; Nastertorabi, F.; Xuan, W.; Han, G. W.; Stevens, R. C.; Schultz, P. G. A Single Reactive Noncanonical Amino Acid Is Able to Dramatically Stabilize Protein Structure. ACS Chem. Biol. 2019, 14 (6), 11501153,  DOI: 10.1021/acschembio.9b00002
  356. 356
    Deiters, A.; Cropp, T. A.; Summerer, D.; Mukherji, M.; Schultz, P. G. Site-Specific PEGylation of Proteins Containing Unnatural Amino Acids. Bioorg. Med. Chem. Lett. 2004, 14 (23), 57435745,  DOI: 10.1016/j.bmcl.2004.09.059
  357. 357
    Schoffelen, S.; Lambermon, M. H. L.; Eldijk, M. B. v.; Hest, J. C. M. v. Site-Specific Modification of Candida antarctica Lipase B via Residue-Specific Incorporation of a Non-Canonical Amino Acid. Bioconjug. Chem. 2008, 19 (6), 11271131,  DOI: 10.1021/bc800019v
  358. 358
    van Dongen, S. F. M.; Nallani, M.; Schoffelen, S.; Cornelissen, J. J. L. M.; Nolte, R. J. M.; van Hest, J. C. M. A Block Copolymer for Functionalisation of Polymersome Surfaces. Macromol. Rapid Commun. 2008, 29 (4), 321325,  DOI: 10.1002/marc.200700765
  359. 359
    Teeuwen, R. L. M.; van Berkel, S. S.; van Dulmen, T. H. H.; Schoffelen, S.; Meeuwissen, S. A.; Zuilhof, H.; de Wolf, F. A.; van Hest, J. C. M. “Clickable” Elastins: Elastin-Like Polypeptides Functionalized with Azide or Alkyne Groups. Chem. Commun. 2009, (27), 40224024,  DOI: 10.1039/b903903a
  360. 360
    Debets, M. F.; van Berkel, S. S.; Schoffelen, S.; Rutjes, F. P. J. T.; van Hest, J. C. M.; van Delft, F. L. Aza-Dibenzocyclooctynes for Fast and Efficient Enzyme PEGylation via Copper-Free (3 + 2) Cycloaddition. Chem. Commun. 2010, 46 (1), 9799,  DOI: 10.1039/B917797C
  361. 361
    Wilding, K. M.; Smith, A. K.; Wilkerson, J. W.; Bush, D. B.; Knotts, T. A. I. V.; Bundy, B. C. The Locational Impact of Site-Specific PEGylation: Streamlined Screening with Cell-Free Protein Expression and Coarse-Grain Simulation. ACS Synth. Biol. 2018, 7 (2), 510521,  DOI: 10.1021/acssynbio.7b00316
  362. 362
    Wu, J. C. Y.; Hutchings, C. H.; Lindsay, M. J.; Werner, C. J.; Bundy, B. C. Enhanced Enzyme Stability Through Site-Directed Covalent Immobilization. J. Biotechnol. 2015, 193, 8390,  DOI: 10.1016/j.jbiotec.2014.10.039
  363. 363
    Basso, A.; Serban, S. Industrial Applications of Immobilized Enzymes─A Review. Mol. Catal. 2019, 479, 110607,  DOI: 10.1016/j.mcat.2019.110607
  364. 364
    Hernandez, K.; Fernandez-Lafuente, R. Control of Protein Immobilization: Coupling Immobilization and Site-Directed Mutagenesis to Improve Biocatalyst or Biosensor Performance. Enzyme Microb. Technol. 2011, 48 (2), 107122,  DOI: 10.1016/j.enzmictec.2010.10.003
  365. 365
    Guan, D.; Kurra, Y.; Liu, W.; Chen, Z. A Click Chemistry Approach to Site-Specific Immobilization of a Small Laccase Enables Efficient Direct Electron Transfer in a Biocathode. Chem. Commun. 2015, 51 (13), 25222525,  DOI: 10.1039/C4CC09179E
  366. 366
    Lim, S. I.; Mizuta, Y.; Takasu, A.; Kim, Y. H.; Kwon, I. Site-Specific Bioconjugation of a Murine Dihydrofolate Reductase Enzyme by Copper(I)-Catalyzed Azide-Alkyne Cycloaddition with Retained Activity. PLOS ONE 2014, 9 (6), e98403  DOI: 10.1371/journal.pone.0098403
  367. 367
    Wang, A.; Du, F.; Pei, X.; Chen, C.; Wu, S. G.; Zheng, Y. Rational Immobilization of Lipase by Combining the Structure Analysis and Unnatural Amino Acid Insertion. J. Mol. Catal. B Enzym. 2016, 132, 5460,  DOI: 10.1016/j.molcatb.2016.06.015
  368. 368
    Li, H.; Yin, Y.; Wang, A.; Li, N.; Wang, R.; Zhang, J.; Chen, X.; Pei, X.; Xie, T. Stable Immobilization of Aldehyde Ketone Reductase Mutants Containing Nonstandard Amino Acids on an Epoxy Resin via Strain-Promoted Alkyne-Azide Cycloaddition. RSC Adv. 2020, 10 (5), 26242633,  DOI: 10.1039/C9RA09067C
  369. 369
    Umeda, A.; Thibodeaux, G. N.; Zhu, J.; Lee, Y.; Zhang, Z. J. Site-specific Protein Cross-Linking with Genetically Incorporated 3,4-Dihydroxy-L-Phenylalanine. ChemBioChem 2009, 10 (8), 13021304,  DOI: 10.1002/cbic.200900127
  370. 370
    Deepankumar, K.; Nadarajan, S. P.; Mathew, S.; Lee, S.-G.; Yoo, T. H.; Hong, E. Y.; Kim, B.-G.; Yun, H. Engineering Transaminase for Stability Enhancement and Site-Specific Immobilization through Multiple Noncanonical Amino Acids Incorporation. ChemCatChem 2015, 7 (3), 417421,  DOI: 10.1002/cctc.201402882
  371. 371
    Bednar, R. M.; Golbek, T. W.; Kean, K. M.; Brown, W. J.; Jana, S.; Baio, J. E.; Karplus, P. A.; Mehl, R. A. Immobilization of Proteins with Controlled Load and Orientation. ACS Appl. Mater. Interfaces 2019, 11 (40), 3639136398,  DOI: 10.1021/acsami.9b12746
  372. 372
    Blizzard, R. J.; Backus, D. R.; Brown, W.; Bazewicz, C. G.; Li, Y.; Mehl, R. A. Ideal Bioorthogonal Reactions Using A Site-Specifically Encoded Tetrazine Amino Acid. J. Am. Chem. Soc. 2015, 137 (32), 1004410047,  DOI: 10.1021/jacs.5b03275
  373. 373
    Switzer, H. J.; Howard, C. A.; Halonski, J. F.; Peairs, E. M.; Smith, N.; Zamecnik, M. P.; Verma, S.; Young, D. D. Employing Non-Canonical Amino Acids Towards the Immobilization of a Hyperthermophilic Enzyme to Increase Protein Stability. RSC Adv. 2023, 13 (13), 84968501,  DOI: 10.1039/D3RA00392B
  374. 374
    Ray, S.; Chand, S.; Zhang, Y.; Nussbaum, S.; Rajeshwar, K.; Perera, R. Implications of Active Site Orientation in Myoglobin for Direct Electron Transfer and Electrocatalysis Based on Monolayer and Multilayer Covalent Immobilization on Gold Electrodes. Electrochim. Acta 2013, 99, 8593,  DOI: 10.1016/j.electacta.2013.03.080
  375. 375
    Xia, L.; Han, M.-J.; Zhou, L.; Huang, A.; Yang, Z.; Wang, T.; Li, F.; Yu, L.; Tian, C.; Zang, Z. S-Click Reaction for Isotropic Orientation of Oxidases on Electrodes to Promote Electron Transfer at Low Potentials. Angew. Chem. Int. Ed. 2019, 58 (46), 1648016484,  DOI: 10.1002/anie.201909203
  376. 376
    Pan, Y.; Li, G.; Liu, R.; Guo, J.; Liu, Y.; Liu, M.; Zhang, X.; Chi, L.; Xu, K.; Wu, R. Unnatural Activities and Mechanistic Insights of Cytochrome P450 PikC Gained from Site-Specific Mutagenesis by Non-Canonical Amino Acids. Nature Commun. 2023, 14 (1), 1669,  DOI: 10.1038/s41467-023-37288-0
  377. 377
    Kolev, J. N.; Zaengle, J. M.; Ravikumar, R.; Fasan, R. Enhancing the Efficiency and Regioselectivity of P450 Oxidation Catalysts by Unnatural Amino Acid Mutagenesis. ChemBioChem 2014, 15 (7), 10011010,  DOI: 10.1002/cbic.201400060
  378. 378
    Ma, H.; Yang, X.; Lu, Z.; Liu, N.; Chen, Y. The “Gate Keeper” Role of Trp222 Determines the Enantiopreference of Diketoreductase toward 2-Chloro-1-Phenylethanone. PLOS ONE 2014, 9 (7), e103792  DOI: 10.1371/journal.pone.0103792
  379. 379
    Yu, Z.; Yu, H.; Tang, H.; Wang, Z.; Wu, J.; Yang, L.; Xu, G. Site-specifically Incorporated Non-Canonical Amino Acids into Pseudomonas alcaligenes Lipase to Hydrolyze L-menthol Propionate among the Eight Isomers. ChemCatChem 2021, 13 (11), 26912701,  DOI: 10.1002/cctc.202100358
  380. 380
    Drienovská, I.; Gajdoš, M.; Kindler, A.; Takhtehchian, M.; Darnhofer, B.; Birner-Gruenberger, R.; Dörr, M.; Bornscheuer, U. T.; Kourist, R. Folding Assessment of Incorporation of Noncanonical Amino Acids Facilitates Expansion of Functional-Group Diversity for Enzyme Engineering. Chem. Eur. J. 2020, 26 (54), 1233812342,  DOI: 10.1002/chem.202002077
  381. 381
    Green, A. P.; Hayashi, T.; Mittl, P. R. E.; Hilvert, D. A Chemically Programmed Proximal Ligand Enhances the Catalytic Properties of a Heme Enzyme. J. Am. Chem. Soc. 2016, 138 (35), 1134411352,  DOI: 10.1021/jacs.6b07029
  382. 382
    Sharp, K. H.; Mewies, M.; Moody, P. C. E.; Raven, E. L. Crystal Structure of the Ascorbate Peroxidase-Ascorbate Complex. Nat. Struct. Biol. 2003, 10 (4), 303307,  DOI: 10.1038/nsb913
  383. 383
    Vojtechovský, J.; Chu, K.; Berendzen, J.; Sweet, R. M.; Schlichting, I. Crystal Structures of Myoglobin-Ligand Complexes at Near-Atomic Resolution. Biophys. J. 1999, 77 (4), 21532174,  DOI: 10.1016/S0006-3495(99)77056-6
  384. 384
    Pott, M.; Hayashi, T.; Mori, T.; Mittl, P. R. E.; Green, A. P.; Hilvert, D. A Noncanonical Proximal Heme Ligand Affords an Efficient Peroxidase in a Globin Fold. J. Am. Chem. Soc. 2018, 140 (4), 15351543,  DOI: 10.1021/jacs.7b12621
  385. 385
    Matsuo, T.; Fukumoto, K.; Watanabe, T.; Hayashi, T. Precise Design of Artificial Cofactors for Enhancing Peroxidase Activity of Myoglobin: Myoglobin Mutant H64D Reconstituted with a ″Single-Winged Cofactor″ Is Equivalent to Native Horseradish Peroxidase in Oxidation Activity. Chem. Asian J. 2011, 6 (9), 24912499,  DOI: 10.1002/asia.201100107
  386. 386
    Hayashi, T.; Tinzl, M.; Mori, T.; Krengel, U.; Proppe, J.; Soetbeer, J.; Klose, D.; Jeschke, G.; Reiher, M.; Hilvert, D. Capture and Characterization of a Reactive Haem-Carbenoid Complex in an Artificial Metalloenzyme. Nat. Catal. 2018, 1 (8), 578584,  DOI: 10.1038/s41929-018-0105-6
  387. 387
    Pott, M.; Tinzl, M.; Hayashi, T.; Ota, Y.; Dunkelmann, D. L.; Mittl, P. R. E.; Hilvert, D. Noncanonical Heme Ligands Steer Carbene Transfer Reactivity in an Artificial Metalloenzyme. Angew. Chem. Int. Ed 2021, 60, 1506315068,  DOI: 10.1002/anie.202103437
  388. 388
    Carminati, D. M.; Fasan, R. Stereoselective Cyclopropanation of Electron-Deficient Olefins with a Cofactor Redesigned Carbene Transferase Featuring Radical Reactivity. ACS Catal. 2019, 9 (10), 96839697,  DOI: 10.1021/acscatal.9b02272
  389. 389
    Moore, E. J.; Fasan, R. Effect of Proximal Ligand Substitutions on the Carbene and Nitrene Transferase Activity of Myoglobin. Tetrahedron 2019, 75 (16), 23572363,  DOI: 10.1016/j.tet.2019.03.009
  390. 390
    Gan, F.; Liu, R.; Wang, F.; Schultz, P. G. Functional Replacement of Histidine in Proteins To Generate Noncanonical Amino Acid Dependent Organisms. J. Am. Chem. Soc. 2018, 140 (11), 38293832,  DOI: 10.1021/jacs.7b13452
  391. 391
    Chand, S.; Ray, S.; Yadav, P.; Samanta, S.; Pierce, B. S.; Perera, R. Abiological Catalysis by Myoglobin Mutant with a Genetically Incorporated Unnatural Amino Acid. Biochem. J. 2021, 478 (9), 17951808,  DOI: 10.1042/BCJ20210091
  392. 392
    Jackson, J. C.; Duffy, S. P.; Hess, K. R.; Mehl, R. A. Improving Nature’s Enzyme Active Site with Genetically Encoded Unnatural Amino Acids. J. Am. Chem. Soc. 2006, 128 (34), 1112411127,  DOI: 10.1021/ja061099y
  393. 393
    Grove, J. I.; Lovering, A. L.; Guise, C.; Race, P. R.; Wrighton, C. J.; White, S. A.; Hyde, E. I.; Searle, P. F. Generation of Escherichia coli Nitroreductase Mutants Conferring Improved Cell Sensitization to the Prodrug CB19541. Cancer Res. 2003, 63 (17), 55325537
  394. 394
    Ugwumba, I. N.; Ozawa, K.; Xu, Z.-Q.; Ely, F.; Foo, J.-L.; Herlt, A. J.; Coppin, C.; Brown, S.; Taylor, M. C.; Ollis, D. L. Improving a Natural Enzyme Activity through Incorporation of Unnatural Amino Acids. J. Am. Chem. Soc. 2011, 133 (2), 326333,  DOI: 10.1021/ja106416g
  395. 395
    Pagar, A. D.; Jeon, H.; Khobragade, T. P.; Sarak, S.; Giri, P.; Lim, S.; Yoo, T. H.; Ko, B. J.; Yun, H. Non-Canonical Amino Acid-Based Engineering of (R)-Amine Transaminase. Front. Chem. 2022,  DOI: 10.3389/fchem.2022.839636
  396. 396
    Savile, C. K.; Janey, J. M.; Mundorff, E. C.; Moore, J. C.; Tam, S.; Jarvis, W. R.; Colbeck, J. C.; Krebber, A.; Fleitz, F. J.; Brands, J. Biocatalytic Asymmetric Synthesis of Chiral Amines from Ketones Applied to Sitagliptin Manufacture. Science 2010, 329 (5989), 305309,  DOI: 10.1126/science.1188934
  397. 397
    Wilkinson, H. C.; Dalby, P. A. Fine-Tuning the Activity and Stability of an Evolved Enzyme Active-Site Through Noncanonical Amino-Acids. FEBS J. 2021, 288 (6), 19351955,  DOI: 10.1111/febs.15560
  398. 398
    Parsons, J. F.; Xiao, G.; Gilliland, G. L.; Armstrong, R. N. Enzymes Harboring Unnatural Amino Acids: Mechanistic and Structural Analysis of the Enhanced Catalytic Activity of a Glutathione Transferase Containing 5-Fluorotryptophan. Biochemistry 1998, 37 (18), 62866294,  DOI: 10.1021/bi980219e
  399. 399
    Dominguez, M. A., Jr; Thornton, K. C.; Melendez, M. G.; Dupureur, C. M. Differential Effects of Isomeric Incorporation of Fluorophenylalanines into PvuII Endonuclease. Proteins 2001, 45 (1), 5561,  DOI: 10.1002/prot.1123
  400. 400
    Hoesl, M. G.; Budisa, N. Expanding and Engineering the Genetic Code in a Single Expression Experiment. ChemBioChem 2011, 12, 552555,  DOI: 10.1002/cbic.201000586
  401. 401
    Cirino, P. C.; Tang, Y.; Takahashi, K.; Tirrell, D. A.; Arnold, F. H. Global Incorporation of Norleucine in Place of Methionine in Cytochrome P450 BM-3 Heme Domain Increases Peroxygenase Activity. Biotechnol. Bioeng. 2003, 83 (6), 729734,  DOI: 10.1002/bit.10718
  402. 402
    Xiao, H.; Nasertorabi, F.; Choi, S.-h.; Han, G. W.; Reed, S. A.; Stevens, R. C.; Schultz, P. G. Exploring the Potential Impact of an Expanded Genetic Code on Protein Function. Proc. Natl. Acad. Sci. U.S.A. 2015, 112 (22), 69616966,  DOI: 10.1073/pnas.1507741112
  403. 403
    Young, T. S.; Ahmad, I.; Yin, J. A.; Schultz, P. G. An Enhanced System for Unnatural Amino Acid Mutagenesis in E. coli. J. Mol. Biol. 2010, 395, 361374,  DOI: 10.1016/j.jmb.2009.10.030
  404. 404
    Schoffelen, S.; Beekwilder, J.; Debets, M. F.; Bosch, D.; Hest, J. C. M. v. Construction of a Multifunctional Enzyme Complex via the Strain-Promoted Azide-Alkyne Cycloaddition. Bioconjug. Chem. 2013, 24 (6), 987996,  DOI: 10.1021/bc400021j
  405. 405
    Lim, S. I.; Cho, J.; Kwon, I. Double Clicking for Site-Specific Coupling of Multiple Enzymes. Chem. Commun. 2015, 51 (71), 1360713610,  DOI: 10.1039/C5CC04611D
  406. 406
    Lim, S. I.; Yang, B.; Jung, Y.; Cha, J.; Cho, J.; Choi, E.-S.; Kim, Y. H.; Kwon, I. Controlled Orientation of Active Sites in a Nanostructured Multienzyme Complex. Sci. Rep. 2016, 6 (1), 39587,  DOI: 10.1038/srep39587
  407. 407
    Ha, J. M.; Jeon, S. T.; Yoon, H. J.; Lee, H. H. Formate Dehydrogenase. PDB 2014,  DOI: 10.2210/pdb3WR5/pdb
  408. 408
    Kavanagh, K. L.; Klimacek, M.; Nidetzky, B.; Wilson, D. K. Crystal Structure of Pseudomonas fluorescens Mannitol 2-Dehydrogenase Binary and Ternary Complexes: Specificity And Catalytic Mechanism. J. Biol. Chem. 2002, 277 (45), 4343343442,  DOI: 10.1074/jbc.M206914200
  409. 409
    Li, J.; Jia, S.; Chen, P. R. Diels-Alder Reaction-Triggered Bioorthogonal Protein Decaging in Living Cells. Nat. Chem. Biol. 2014, 10 (12), 10031005,  DOI: 10.1038/nchembio.1656
  410. 410
    Li, J.; Yu, J.; Zhao, J.; Wang, J.; Zheng, S.; Lin, S.; Chen, L.; Yang, M.; Jia, S.; Zhang, X. Palladium-Triggered Deprotection Chemistry for Protein Activation in Living Cells. Nat. Chem. 2014, 6 (4), 352361,  DOI: 10.1038/nchem.1887
  411. 411
    Wang, J.; Zheng, S.; Liu, Y.; Zhang, Z.; Lin, Z.; Li, J.; Zhang, G.; Wang, X.; Li, J.; Chen, P. R. Palladium-Triggered Chemical Rescue of Intracellular Proteins via Genetically Encoded Allene-Caged Tyrosine. J. Am. Chem. Soc. 2016, 138 (46), 1511815121,  DOI: 10.1021/jacs.6b08933
  412. 412
    Georgianna, W. E.; Lusic, H.; McIver, A. L.; Deiters, A. Photocleavable Polyethylene Glycol for the Light-Regulation of Protein Function. Bioconjug. Chem. 2010, 21 (8), 14041407,  DOI: 10.1021/bc100084n
  413. 413
    Wu, N.; Deiters, A.; Cropp, T. A.; King, D.; Schultz, P. G. A Genetically Encoded Photocaged Amino Acid. J. Am. Chem. Soc. 2004, 126 (44), 1430614307,  DOI: 10.1021/ja040175z
  414. 414
    Deiters, A.; Groff, D.; Ryu, Y.; Xie, J.; Schultz, P. G. A Genetically Encoded Photocaged Tyrosine. Angew. Chem. Int. Ed. 2006, 45 (17), 27282731,  DOI: 10.1002/anie.200600264
  415. 415
    Chou, C.; Young, D. D.; Deiters, A. A Light-Activated DNA Polymerase. Angew. Chem. Int. Ed. 2009, 48 (32), 59505953,  DOI: 10.1002/anie.200901115
  416. 416
    Chou, C.; Young, D. D.; Deiters, A. Photocaged T7 RNA Polymerase for the Light Activation of Transcription and Gene Function in Pro- and Eukaryotic Cells. ChemBioChem 2010, 11 (7), 972977,  DOI: 10.1002/cbic.201000041
  417. 417
    Chou, C.; Deiters, A. Light-Activated Gene Editing with a Photocaged Zinc-Finger Nuclease. Angew. Chem. Int. Ed. 2011, 50 (30), 6839,  DOI: 10.1002/anie.201101157
  418. 418
    Gautier, A.; Nguyen, D. P.; Lusic, H.; An, W.; Deiters, A.; Chin, J. W. Genetically Encoded Photocontrol of Protein Localization in Mammalian Cells. J. Am. Chem. Soc. 2010, 132 (12), 40864088,  DOI: 10.1021/ja910688s
  419. 419
    Gautier, A.; Deiters, A.; Chin, J. W. Light-Activated Kinases Enable Temporal Dissection of Signaling Networks in Living Cells. J. Am. Chem. Soc. 2011, 133 (7), 21242127,  DOI: 10.1021/ja1109979
  420. 420
    Hemphill, J.; Chou, C.; Chin, J. W.; Deiters, A. Genetically Encoded Light-Activated Transcription for Spatiotemporal Control of Gene Expression and Gene Silencing in Mammalian Cells. J. Am. Chem. Soc. 2013, 135 (36), 1343313439,  DOI: 10.1021/ja4051026
  421. 421
    Hemphill, J.; Borchardt, E. K.; Brown, K.; Asokan, A.; Deiters, A. Optical Control of CRISPR/Cas9 Gene Editing. J. Am. Chem. Soc. 2015, 137 (17), 56425645,  DOI: 10.1021/ja512664v
  422. 422
    Nguyen, D. P.; Mahesh, M.; Elsässer, S. J.; Hancock, S. M.; Uttamapinant, C.; Chin, J. W. Genetic Encoding of Photocaged Cysteine Allows Photoactivation of TEV Protease in Live Mammalian Cells. J. Am. Chem. Soc. 2014, 136 (6), 22402243,  DOI: 10.1021/ja412191m
  423. 423
    Uprety, R.; Luo, J.; Liu, J.; Naro, Y.; Samanta, S.; Deiters, A. Genetic Encoding of Caged Cysteine and Caged Homocysteine in Bacterial and Mammalian Cells. ChemBioChem 2014, 15 (12), 17931799,  DOI: 10.1002/cbic.201400073
  424. 424
    Yang, X.; Zhao, L.; Wang, Y.; Ji, Y.; Su, X. C.; Ma, J. A.; Xuan, W. Constructing Photoactivatable Protein with Genetically Encoded Photocaged Glutamic Acid. Angew. Chem. Int. Ed. 2023, 135 (40), e202308472  DOI: 10.1002/ange.202308472
  425. 425
    Givens, R. S.; Weber, J. F.; Jung, A. H.; Park, C.-H. New Photoprotecting Groups: Desyl and p-Hydroxyphenacyl Phosphate and Carboxylate Esters. In Methods in enzymology; Elsevier, 1998; Vol. 291, pp 129. DOI: 10.1016/s0076-6879(98)91004-7 .
  426. 426
    Mangubat-Medina, A. E.; Ball, Z. T. Triggering Biological Processes: Methods and Applications of Photocaged Peptides and Proteins. Chem. Soc. Rev. 2021, 50 (18), 1040310421,  DOI: 10.1039/D0CS01434F
  427. 427
    Grasso, K. T.; Singha Roy, S. J.; Osgood, A. O.; Yeo, M. J. R.; Soni, C.; Hillenbrand, C. M.; Ficaretta, E. D.; Chatterjee, A. A Facile Platform to Engineer Escherichia coli Tyrosyl-tRNA Synthetase Adds New Chemistries to the Eukaryotic Genetic Code, Including a Phosphotyrosine Mimic. ACS Cent. Sci. 2022, 8 (4), 483492,  DOI: 10.1021/acscentsci.1c01465
  428. 428
    Edwards, W. F.; Young, D. D.; Deiters, A. Light-Activated Cre Recombinase as a Tool for the Spatial and Temporal Control of Gene Function in Mammalian Cells. ACS Chem. Biol. 2009, 4 (6), 441445,  DOI: 10.1021/cb900041s
  429. 429
    Le Provost, F.; Lillico, S.; Passet, B.; Young, R.; Whitelaw, B.; Vilotte, J.-L. Zinc Finger Nuclease Technology Heralds a New Era in Mammalian Transgenesis. Trends Biotechnol. 2010, 28 (3), 134141,  DOI: 10.1016/j.tibtech.2009.11.007
  430. 430
    Porteus, M. H. Mammalian Gene Targeting with Designed Zinc Finger Nucleases. Molecular Therapy 2006, 13 (2), 438446,  DOI: 10.1016/j.ymthe.2005.08.003
  431. 431
    Porteus, M. H.; Carroll, D. Gene Targeting Using Zinc Finger Nucleases. Nat. Biotechnol. 2005, 23 (8), 967973,  DOI: 10.1038/nbt1125
  432. 432
    Urnov, F. D.; Rebar, E. J.; Holmes, M. C.; Zhang, H. S.; Gregory, P. D. Genome Editing with Engineered Zinc Finger Nucleases. Nat. Rev. Genet. 2010, 11 (9), 636646,  DOI: 10.1038/nrg2842
  433. 433
    Rémy, S.; Tesson, L.; Ménoret, S.; Usal, C.; Scharenberg, A. M.; Anegon, I. Zinc-Finger Nucleases: A Powerful Tool for Genetic Engineering of Animals. Transgenic Res. 2010, 19, 363371,  DOI: 10.1007/s11248-009-9323-7
  434. 434
    Li, Y.; Korolev, S.; Waksman, G. Crystal Structures of Open and Closed Forms of Binary and Ternary Complexes of the Large Fragment of Thermus aquaticus DNA Polymerase I: Structural Basis for Nucleotide Incorporation. EMBO J. 1998, 17 (24), 75147525,  DOI: 10.1093/emboj/17.24.7514
  435. 435
    Suzuki, M.; Baskin, D.; Hood, L.; Loeb, L. A. Random Mutagenesis of Thermus aquaticus DNA Polymerase I: Concordance of Immutable Sites in vivo with the Crystal Structure. Proc. Natl. Acad. Sci. U.S.A. 1996, 93 (18), 96709675,  DOI: 10.1073/pnas.93.18.9670
  436. 436
    Paul, N.; Shum, J.; Le, T. Hot Start PCR. RT-PCR Protocols, Second ed.; 2010; pp 301318. DOI: 10.1007/978-1-60761-629-0
  437. 437
    Cramer, P. Common Structural Features of Nucleic Acid Polymerases. Bioessays 2002, 24 (8), 724729,  DOI: 10.1002/bies.10127
  438. 438
    Banghart, M. R.; Mourot, A.; Fortin, D. L.; Yao, J. Z.; Kramer, R. H.; Trauner, D. Photochromic Blockers of Voltage-Gated Potassium Channels. Angew. Chem. Int. Ed. 2009, 48 (48), 90979101,  DOI: 10.1002/anie.200904504
  439. 439
    Beharry, A. A.; Wong, L.; Tropepe, V.; Woolley, G. A. Fluorescence Imaging of Azobenzene Photoswitching in vivo. Angew. Chem. Int. Ed. 2011, 50 (6), 13251327,  DOI: 10.1002/anie.201006506
  440. 440
    Bléger, D.; Schwarz, J.; Brouwer, A. M.; Hecht, S. o-Fluoroazobenzenes as Readily Synthesized Photoswitches Offering Nearly Quantitative Two-Way Isomerization with Visible Light. J. Am. Chem. Soc. 2012, 134 (51), 2059720600,  DOI: 10.1021/ja310323y
  441. 441
    Bonardi, F.; London, G.; Nouwen, N.; Feringa, B. L.; Driessen, A. J. Light-Induced Control of Protein Translocation by the SecYEG Complex. Angew. Chem. Int. Ed. 2010, 49 (40), 72347238,  DOI: 10.1002/anie.201002243
  442. 442
    Gorostiza, P.; Isacoff, E. Y. Nanoengineering Ion Channels for Optical Control. Physiology 2008, 23 (5), 238247,  DOI: 10.1152/physiol.00018.2008
  443. 443
    Knie, C.; Utecht, M.; Zhao, F.; Kulla, H.; Kovalenko, S.; Brouwer, A. M.; Saalfrank, P.; Hecht, S.; Bléger, D. ortho-Fluoroazobenzenes: Visible Light Switches with Very Long-Lived Z Isomers. Chem. Eur. J. 2014, 20 (50), 1649216501,  DOI: 10.1002/chem.201404649
  444. 444
    Liang, X.; Mochizuki, T.; Asanuma, H. A Supra-Photoswitch Involving Sandwiched DNA Base Pairs and Azobenzenes for Light-Driven Nanostructures and Nanodevices. Small 2009, 5 (15), 17611768,  DOI: 10.1002/smll.200900223
  445. 445
    Samanta, S.; Beharry, A. A.; Sadovski, O.; McCormick, T. M.; Babalhavaeji, A.; Tropepe, V.; Woolley, G. A. Photoswitching Azo Compounds in vivo with Red Light. J. Am. Chem. Soc. 2013, 135 (26), 97779784,  DOI: 10.1021/ja402220t
  446. 446
    Schierling, B.; Noel, A.-J.; Wende, W.; Hien, L. T.; Volkov, E.; Kubareva, E.; Oretskaya, T.; Kokkinidis, M.; Rompp, A.; Spengler, B.; Pingoud, A. Controlling the Enzymatic Activity of a Restriction Enzyme by Light. Proc. Natl. Acad. Sci. U.S.A. 2010, 107 (4), 13611366,  DOI: 10.1073/pnas.0909444107
  447. 447
    Zhang, F.; Zarrine-Afsar, A.; Al-Abdul-Wahid, M. S.; Prosser, R. S.; Davidson, A. R.; Woolley, G. A. Structure-Based Approach to the Photocontrol of Protein Folding. J. Am. Chem. Soc. 2009, 131 (6), 22832289,  DOI: 10.1021/ja807938v
  448. 448
    Bose, M.; Groff, D.; Xie, J.; Brustad, E.; Schultz, P. G. The Incorporation of a Photoisomerizable Amino Acid into Proteins in E. c oli. J. Am. Chem. Soc. 2006, 128 (2), 388389,  DOI: 10.1021/ja055467u
  449. 449
    Hoppmann, C.; Lacey, V. K.; Louie, G. V.; Wei, J.; Noel, J. P.; Wang, L. Genetically Encoding Photoswitchable Click Amino Acids in Escherichia coli and Mammalian Cells. Angew. Chem. Int. Ed. 2014, 53 (15), 39323936,  DOI: 10.1002/anie.201400001
  450. 450
    Hoppmann, C.; Maslennikov, I.; Choe, S.; Wang, L. In situ Formation of an Azo Bridge on Proteins Controllable by Visible Light. J. Am. Chem. Soc. 2015, 137 (35), 1121811221,  DOI: 10.1021/jacs.5b06234
  451. 451
    John, A. A.; Ramil, C. P.; Tian, Y.; Cheng, G.; Lin, Q. Synthesis and Site-Specific Incorporation of Red-Shifted Azobenzene Amino Acids into Proteins. Org. Lett. 2015, 17 (24), 62586261,  DOI: 10.1021/acs.orglett.5b03268
  452. 452
    Klippenstein, V.; Hoppmann, C.; Ye, S.; Wang, L.; Paoletti, P. Optocontrol of Glutamate Receptor Activity by Single Side-Chain Photoisomerization. Elife 2017, 6, e25808  DOI: 10.7554/eLife.25808
  453. 453
    Kneuttinger, A. C.; Straub, K.; Bittner, P.; Simeth, N. A.; Bruckmann, A.; Busch, F.; Rajendran, C.; Hupfeld, E.; Wysocki, V. H.; Horinek, D. Light Regulation of Enzyme Allostery through Photo-responsive Unnatural Amino Acids. Cell Chem. Biol. 2019, 26 (11), 15011514,  DOI: 10.1016/j.chembiol.2019.08.006
  454. 454
    Luo, J.; Samanta, S.; Convertino, M.; Dokholyan, N. V.; Deiters, A. Reversible and Tunable Photoswitching of Protein Function through Genetic Encoding of Azobenzene Amino Acids in Mammalian Cells. ChemBioChem 2018, 19 (20), 21782185,  DOI: 10.1002/cbic.201800226
  455. 455
    Kneuttinger, A. C.; Winter, M.; Simeth, N. A.; Heyn, K.; Merkl, R.; König, B.; Sterner, R. Artificial Light Regulation of an Allosteric Bienzyme Complex by a Photosensitive Ligand. ChemBioChem 2018, 19 (16), 17501757,  DOI: 10.1002/cbic.201800219
  456. 456
    Kneuttinger, A. C.; Rajendran, C.; Simeth, N. A.; Bruckmann, A.; König, B.; Sterner, R. Significance of the Protein Interface Configuration for Allostery in Imidazole Glycerol Phosphate Synthase. Biochemistry 2020, 59 (29), 27292742,  DOI: 10.1021/acs.biochem.0c00332
  457. 457
    Zubi, Y. S.; Seki, K.; Li, Y.; Hunt, A. C.; Liu, B.; Roux, B.; Jewett, M. C.; Lewis, J. C. Metal-Responsive Regulation of Enzyme Catalysis Using Genetically Encoded Chemical Switches. Nature Commun. 2022, 13 (1), 1864,  DOI: 10.1038/s41467-022-29239-y
  458. 458
    Yang, H.; Swartz, A. M.; Park, H. J.; Srivastava, P.; Ellis-Guardiola, K.; Upp, D. M.; Lee, G.; Belsare, K.; Gu, Y.; Zhang, C. Evolving Artificial Metalloenzymes via Random Mutagenesis. Nat. Chem. 2018, 10 (3), 318324,  DOI: 10.1038/nchem.2927
  459. 459
    Leveson-Gower, R. B.; Zhou, Z.; Drienovská, I.; Roelfes, G. Unlocking Iminium Catalysis in Artificial Enzymes to Create a Friedel-Crafts Alkylase. ACS Catal. 2021, 11 (12), 67636770,  DOI: 10.1021/acscatal.1c00996
  460. 460
    Mayer, C.; Dulson, C.; Reddem, E.; Thunnissen, A.-M. W. H.; Roelfes, G. Directed Evolution of a Designer Enzyme Featuring an Unnatural Catalytic Amino Acid. Angew. Chem. Int. Ed. 2019, 58 (7), 20832087,  DOI: 10.1002/anie.201813499
  461. 461
    Kaes, C.; Katz, A.; Hosseini, M. W. Bipyridine: The Most Widely Used Ligand. A Review of Molecules Comprising at Least Two 2,2‘-Bipyridine Units. Chem. Rev. 2000, 100 (10), 35533590,  DOI: 10.1021/cr990376z
  462. 462
    Xie, J.; Liu, W.; Schultz, P. G. A Genetically Encoded Bidentate, Metal-Binding Amino Acid. Angew. Chem. Int. Ed. 2007, 46 (48), 92399242,  DOI: 10.1002/anie.200703397
  463. 463
    Lee, H. S.; Schultz, P. G. Biosynthesis of a Site-Specific DNA Cleaving Protein. J. Am. Chem. Soc. 2008, 130 (40), 1319413195,  DOI: 10.1021/ja804653f
  464. 464
    Roelfes, G. LmrR: A Privileged Scaffold for Artificial Metalloenzymes. Acc. Chem. Res. 2019, 52 (3), 545556,  DOI: 10.1021/acs.accounts.9b00004
  465. 465
    Drienovská, I.; Rioz-Martínez, A.; Draksharapu, A.; Roelfes, G. Novel Artificial Metalloenzymes by in vivo Incorporation of Metal-Binding Unnatural Amino Acids. Chem. Sci. 2015, 6 (1), 770776,  DOI: 10.1039/C4SC01525H
  466. 466
    Drienovská, I.; Alonso-Cotchico, L.; Vidossich, P.; Lledós, A.; Maréchal, J.-D.; Roelfes, G. Design of an Enantioselective Artificial Metallo-Hydratase Enzyme Containing an Unnatural Metal-Binding Amino Acid. Chem. Sci. 2017, 8 (10), 72287235,  DOI: 10.1039/C7SC03477F
  467. 467
    Ségaud, N.; Drienovská, I.; Chen, J.; Browne, W. R.; Roelfes, G. Artificial Metalloproteins for Binding and Stabilization of a Semiquinone Radical. Inorg. Chem. 2017, 56 (21), 1329313299,  DOI: 10.1021/acs.inorgchem.7b02073
  468. 468
    Bersellini, M.; Roelfes, G. Multidrug Resistance Regulators (MDRs) as Scaffolds for the Design of Artificial Metalloenzymes. Org. Biomol. Chem. 2017, 15 (14), 30693073,  DOI: 10.1039/C7OB00390K
  469. 469
    Jung, S.-M.; Yang, M.; Song, W. J. Symmetry-Adapted Synthesis of Dicopper Oxidases with Divergent Dioxygen Reactivity. Inorg. Chem. 2022, 61 (31), 1243312441,  DOI: 10.1021/acs.inorgchem.2c01898
  470. 470
    Drienovská, I.; Scheele, R. A.; Gutiérrez de Souza, C.; Roelfes, G. A Hydroxyquinoline-Based Unnatural Amino Acid for the Design of Novel Artificial Metalloenzymes. ChemBioChem 2020, 21 (21), 30773081,  DOI: 10.1002/cbic.202000306
  471. 471
    Stein, A.; Liang, A. D.; Sahin, R.; Ward, T. R. Incorporation of Metal-Chelating Unnatural Amino Acids into Halotag for Allylic Deamination. J. Organomet. Chem. 2022, 962, 122272,  DOI: 10.1016/j.jorganchem.2022.122272
  472. 472
    Los, G. V.; Encell, L. P.; McDougall, M. G.; Hartzell, D. D.; Karassina, N.; Zimprich, C.; Wood, M. G.; Learish, R.; Ohana, R. F.; Urh, M. HaloTag: A Novel Protein Labeling Technology for Cell Imaging and Protein Analysis. ACS Chem. Biol. 2008, 3 (6), 373382,  DOI: 10.1021/cb800025k
  473. 473
    Coquière, D.; Bos, J.; Beld, J.; Roelfes, G. Enantioselective Artificial Metalloenzymes Based on a Bovine Pancreatic Polypeptide Scaffold. Angew. Chem. Int. Ed. 2009, 48 (28), 51595162,  DOI: 10.1002/anie.200901134
  474. 474
    Madoori, P. K.; Agustiandari, H.; Driessen, A. J. M.; Thunnissen, A. M. W. H. Structure of the Transcriptional Regulator LmrR and Its Mechanism of Multidrug Recognition. EMBO J. 2009, 28 (2), 156166,  DOI: 10.1038/emboj.2008.263
  475. 475
    Yu, Y.; Hu, C.; Xia, L.; Wang, J. Artificial Metalloenzyme Design with Unnatural Amino Acids and Non-Native Cofactors. ACS Catal. 2018, 8 (3), 18511863,  DOI: 10.1021/acscatal.7b03754
  476. 476
    Yang, H.; Srivastava, P.; Zhang, C.; Lewis, J. C. A General Method for Artificial Metalloenzyme Formation through Strain-Promoted Azide-Alkyne Cycloaddition. ChemBioChem 2014, 15 (2), 223227,  DOI: 10.1002/cbic.201300661
  477. 477
    Srivastava, P.; Yang, H.; Ellis-Guardiola, K.; Lewis, J. C. Engineering a Dirhodium Artificial Metalloenzyme for Selective Olefin Cyclopropanation. Nature Commun. 2015, 6 (1), 7789,  DOI: 10.1038/ncomms8789
  478. 478
    Upp, D. M.; Huang, R.; Li, Y.; Bultman, M. J.; Roux, B.; Lewis, J. C. Engineering Dirhodium Artificial Metalloenzymes for Diazo Coupling Cascade Reactions. Angew. Chem. Int. Ed. 2021, 60 (44), 2367223677,  DOI: 10.1002/anie.202107982
  479. 479
    Ellis-Guardiola, K.; Rui, H.; Beckner, R. L.; Srivastava, P.; Sukumar, N.; Roux, B.; Lewis, J. C. Crystal Structure and Conformational Dynamics of Pyrococcus furiosus Prolyl Oligopeptidase. Biochemistry 2019, 58 (12), 16161626,  DOI: 10.1021/acs.biochem.9b00031
  480. 480
    Brady, L.; Brzozowski, A. M.; Derewenda, Z. S.; Dodson, E.; Dodson, G.; Tolley, S.; Turkenburg, J. P.; Christiansen, L.; Huge-Jensen, B.; Norskov, L. A Serine Protease Triad Forms the Catalytic Centre of a Triacylglycerol Lipase. Nature 1990, 343 (6260), 767770,  DOI: 10.1038/343767a0
  481. 481
    Buller, A. R.; Townsend, C. A. Intrinsic Evolutionary Constraints on Protease Structure, Enzyme Acylation, and the Identity of the Catalytic Triad. Proc. Natl. Acad. Sci. U.S.A. 2013, 110 (8), E653-E661  DOI: 10.1073/pnas.1221050110
  482. 482
    Smith, A. J. T.; Müller, R.; Toscano, M. D.; Kast, P.; Hellinga, H. W.; Hilvert, D.; Houk, K. N. Structural Reorganization and Preorganization in Enzyme Active Sites: Comparisons of Experimental and Theoretically Ideal Active Site Geometries in the Multistep Serine Esterase Reaction Cycle. J. Am. Chem. Soc. 2008, 130 (46), 1536115373,  DOI: 10.1021/ja803213p
  483. 483
    Burton, A. J.; Thomson, A. R.; Dawson, W. M.; Brady, R. L.; Woolfson, D. N. Installing Hydrolytic Activity into a Completely De Novo Protein Framework. Nat. Chem. 2016, 8 (9), 837844,  DOI: 10.1038/nchem.2555
  484. 484
    Rajagopalan, S.; Wang, C.; Yu, K.; Kuzin, A. P.; Richter, F.; Lew, S.; Miklos, A. E.; Matthews, M. L.; Seetharaman, J.; Su, M. Design of Activated Serine-Containing Catalytic Triads with Atomic-Level Accuracy. Nat. Chem. Biol. 2014, 10 (5), 386391,  DOI: 10.1038/nchembio.1498
  485. 485
    Richter, F.; Blomberg, R.; Khare, S. D.; Kiss, G.; Kuzin, A. P.; Smith, A. J. T.; Gallaher, J.; Pianowski, Z.; Helgeson, R. C.; Grjasnow, A. Computational Design of Catalytic Dyads and Oxyanion Holes for Ester Hydrolysis. J. Am. Chem. Soc. 2012, 134 (39), 1619716206,  DOI: 10.1021/ja3037367
  486. 486
    Burke, A. J.; Lovelock, S. L.; Frese, A.; Crawshaw, R.; Ortmayer, M.; Dunstan, M.; Levy, C.; Green, A. P. Design and Evolution of an Enzyme with a Non-Canonical Organocatalytic Mechanism. Nature 2019, 570 (7760), 219223,  DOI: 10.1038/s41586-019-1262-8
  487. 487
    Bjelic, S.; Nivón, L. G.; Çelebi-Ölçüm, N.; Kiss, G.; Rosewall, C. F.; Lovick, H. M.; Ingalls, E. L.; Gallaher, J. L.; Seetharaman, J.; Lew, S. Computational Design of Enone-Binding Proteins with Catalytic Activity for the Morita-Baylis-Hillman Reaction. ACS Chem. Biol. 2013, 8 (4), 749757,  DOI: 10.1021/cb3006227
  488. 488
    Hutton, A. E.; Foster, J.; Crawshaw, R.; Hardy, F. J.; Johannissen, L. O.; Lister, T. M.; Gerard, E. F.; Birch-Price, Z.; Obexer, R.; Hay, S.; Green, A. P. A Non-Canonical Nucleophile Unlocks a New Mechanistic Pathway in a Designed Enzyme. Nat Commun 2024,  DOI: 10.1038/s41467-024-46123-z
  489. 489
    Crawshaw, R.; Crossley, A.; Johannissen, L.; Burke, A.; Hay, S.; Levy, C.; Baker, D.; Lovelock, S.; Green, A. Engineering an Efficient and Enantioselective Enzyme for the Morita-Baylis-Hillman Reaction. Nat. Chem. 2022, 14, 313,  DOI: 10.1038/s41557-021-00833-9
  490. 490
    Agten, S. M.; Dawson, P. E.; Hackeng, T. M. Oxime Conjugation in Protein Chemistry: From Carbonyl Incorporation to Nucleophilic Catalysis. J. Pept. Sci. 2016, 22 (5), 271279,  DOI: 10.1002/psc.2874
  491. 491
    Kölmel, D. K.; Kool, E. T. Oximes and Hydrazones in Bioconjugation: Mechanism and Catalysis. Chem. Rev. 2017, 117 (15), 1035810376,  DOI: 10.1021/acs.chemrev.7b00090
  492. 492
    Drienovská, I.; Mayer, C.; Dulson, C.; Roelfes, G. A Designer Enzyme for Hydrazone and Oxime Formation Featuring an Unnatural Catalytic Aniline Residue. Nat. Chem. 2018, 10 (9), 946952,  DOI: 10.1038/s41557-018-0082-z
  493. 493
    Ofori Atta, L.; Zhou, Z.; Roelfes, G. In vivo Biocatalytic Cascades Featuring an Artificial-Enzyme-Catalysed New-to-Nature Reaction. Angew. Chem. Int. Ed. 2023, 62 (1), e202214191  DOI: 10.1002/anie.202214191
  494. 494
    Leveson-Gower, R. B.; de Boer, R. M.; Roelfes, G. Tandem Friedel-Crafts-Alkylation-Enantioselective-Protonation by Artificial Enzyme Iminium Catalysis. ChemCatChem 2022, 14 (8), e202101875  DOI: 10.1002/cctc.202101875
  495. 495
    Zhou, Z.; Roelfes, G. Synergistic Catalysis in an Artificial Enzyme by Simultaneous Action of Two Abiological Catalytic Sites. Nat. Catal. 2020, 3 (3), 289294,  DOI: 10.1038/s41929-019-0420-6
  496. 496
    Zhou, Z.; Roelfes, G. Synergistic Catalysis of Tandem Michael Addition/Enantioselective Protonation Reactions by an Artificial Enzyme. ACS Catal. 2021, 11 (15), 93669369,  DOI: 10.1021/acscatal.1c02298
  497. 497
    Gran-Scheuch, A.; Bonandi, E.; Drienovská, I. Expanding the Genetic Code: Incorporation of Functional Secondary Amines via Stop Codon Suppression. ChemCatChem 2024, 16, e202301004  DOI: 10.1002/cctc.202301004
  498. 498
    Longwitz, L.; Leveson-Gower, R. B.; Rozeboom, H. J.; Thunnissen, A.-M. W. H.; Roelfes, G. Boron Catalysis in a Designer Enzyme. Nature 2024, 629, 824,  DOI: 10.1038/s41586-024-07391-3
  499. 499
    Garrido-Castro, A. F.; Maestro, M. C.; Alemán, J. Asymmetric Induction in Photocatalysis - Discovering a New Side to Light-Driven Chemistry. Tetrahedron Lett. 2018, 59 (14), 12861294,  DOI: 10.1016/j.tetlet.2018.02.040
  500. 500
    Yao, W.; Bazan-Bergamino, E. A.; Ngai, M.-Y. Asymmetric Photocatalysis Enabled by Chiral Organocatalysts. ChemCatChem 2022, 14 (1), e202101292  DOI: 10.1002/cctc.202101292
  501. 501
    Heyes, D. J.; Lakavath, B.; Hardman, S. J. O.; Sakuma, M.; Hedison, T. M.; Scrutton, N. S. Photochemical Mechanism of Light-Driven Fatty Acid Photodecarboxylase. ACS Catal. 2020, 10 (12), 66916696,  DOI: 10.1021/acscatal.0c01684
  502. 502
    Heyes, D. J.; Zhang, S.; Taylor, A.; Johannissen, L. O.; Hardman, S. J. O.; Hay, S.; Scrutton, N. S. Photocatalysis as the ‘Master Switch’ of Photomorphogenesis in Early Plant Development. Nature Plants 2021, 7 (3), 268276,  DOI: 10.1038/s41477-021-00866-5
  503. 503
    Tan, C.; Liu, Z.; Li, J.; Guo, X.; Wang, L.; Sancar, A.; Zhong, D. The Molecular Origin of High DNA-Repair Efficiency by Photolyase. Nature Commun. 2015, 6 (1), 7302,  DOI: 10.1038/ncomms8302
  504. 504
    Black, M. J.; Biegasiewicz, K. F.; Meichan, A. J.; Oblinsky, D. G.; Kudisch, B.; Scholes, G. D.; Hyster, T. K. Asymmetric Redox-Neutral Radical Cyclization Catalysed by Flavin-Dependent ‘Ene’-Reductases. Nat. Chem. 2020, 12 (1), 7175,  DOI: 10.1038/s41557-019-0370-2
  505. 505
    Emmanuel, M. A.; Greenberg, N. R.; Oblinsky, D. G.; Hyster, T. K. Accessing Non-Natural Reactivity by Irradiating Nicotinamide-Dependent Enzymes with Light. Nature 2016, 540 (7633), 414417,  DOI: 10.1038/nature20569
  506. 506
    Liu, X.; Kang, F.; Hu, C.; Wang, L.; Xu, Z.; Zheng, D.; Gong, W.; Lu, Y.; Ma, Y.; Wang, J. A Genetically Encoded Photosensitizer Protein Facilitates the Rational Design of a Miniature Photocatalytic CO2-Reducing Enzyme. Nat. Chem. 2018, 10 (12), 12011206,  DOI: 10.1038/s41557-018-0150-4
  507. 507
    Siegel, J. B.; Zanghellini, A.; Lovick, H. M.; Kiss, G.; Lambert, A. R.; St. Clair, J. L.; Gallaher, J. L.; Hilvert, D.; Gelb, M. H.; Stoddard, B. L.; Houk, K. N.; Michael, F. E.; Baker, D. Computational Design of an Enzyme Catalyst for a Stereoselective Bimolecular Diels-Alder Reaction. Science 2010, 329, 309313,  DOI: 10.1126/science.1190239
  508. 508
    Sun, N.; Huang, J.; Qian, J.; Zhou, T.-P.; Guo, J.; Tang, L.; Zhang, W.; Deng, Y.; Zhao, W.; Wu, G. Enantioselective [2 + 2]-Cycloadditions with Triplet Photoenzymes. Nature 2022, 611 (7937), 715720,  DOI: 10.1038/s41586-022-05342-4
  509. 509
    Allen, A. R.; Noten, E. A.; Stephenson, C. R. J. Aryl Transfer Strategies Mediated by Photoinduced Electron Transfer. Chem. Rev. 2022, 122 (2), 26952751,  DOI: 10.1021/acs.chemrev.1c00388
  510. 510
    Dadashi-Silab, S.; Doran, S.; Yagci, Y. Photoinduced Electron Transfer Reactions for Macromolecular Syntheses. Chem. Rev. 2016, 116 (17), 1021210275,  DOI: 10.1021/acs.chemrev.5b00586
  511. 511
    Reid, B. G.; Flynn, G. C. Chromophore Formation in Green Fluorescent Protein. Biochemistry 1997, 36 (22), 67866791,  DOI: 10.1021/bi970281w
  512. 512
    Fu, Y.; Huang, J.; Wu, Y.; Liu, X.; Zhong, F.; Wang, J. Biocatalytic Cross-Coupling of Aryl Halides with a Genetically Engineered Photosensitizer Artificial Dehalogenase. J. Am. Chem. Soc. 2021, 143 (2), 617622,  DOI: 10.1021/jacs.0c10882
  513. 513
    Gu, Y.; Ellis-Guardiola, K.; Srivastava, P.; Lewis, J. C. Preparation, Characterization, and Oxygenase Activity of a Photocatalytic Artificial Enzyme. ChemBioChem 2015, 16 (13), 18801883,  DOI: 10.1002/cbic.201500165
  514. 514
    Zubi, Y. S.; Liu, B.; Gu, Y.; Sahoo, D.; Lewis, J. C. Controlling the Optical and Catalytic Properties of Artificial Metalloenzyme Photocatalysts Using Chemogenetic Engineering. Chem. Sci. 2022, 13 (5), 14591468,  DOI: 10.1039/D1SC05792H
  515. 515
    Liu, B.; Zubi, Y. S.; Lewis, J. C. Iridium(iii) Polypyridine Artificial Metalloenzymes with Tunable Photophysical Properties: A New Platform for Visible Light Photocatalysis in Aqueous Solution. Dalton Trans. 2023, 52 (16), 50345038,  DOI: 10.1039/D3DT00932G
  516. 516
    Lee, J.; Song, W. J. Photocatalytic C-O Coupling Enzymes That Operate via Intramolecular Electron Transfer. J. Am. Chem. Soc. 2023, 145 (9), 52115221,  DOI: 10.1021/jacs.2c12226
  517. 517
    Mills, J. H.; Sheffler, W.; Ener, M. E.; Almhjell, P. J.; Oberdorfer, G.; Pereira, J. H.; Parmeggiani, F.; Sankaran, B.; Zwart, P. H.; Baker, D. Computational Design of a Homotrimeric Metalloprotein with a Trisbipyridyl Core. Proc. Natl. Acad. Sci. U.S.A. 2016, 113 (52), 1501215017,  DOI: 10.1073/pnas.1600188113
  518. 518
    Duan, H.-Z.; Hu, C.; Li, Y.-L.; Wang, S.-H.; Xia, Y.; Liu, X.; Wang, J.; Chen, Y.-X. Genetically Encoded Phosphine Ligand for Metalloprotein Design. J. Am. Chem. Soc. 2022, 144 (50), 2283122837,  DOI: 10.1021/jacs.2c09683
  519. 519
    Beattie, A. T.; Dunkelmann, D. L.; Chin, J. W. Quintuply Orthogonal Pyrrolysyl-tRNA Synthetase/tRNAPyl Pairs. Nat. Chem. 2023, 15 (7), 948959,  DOI: 10.1038/s41557-023-01232-y
  520. 520
    Dunkelmann, D. L.; Willis, J. C. W.; Beattie, A. T.; Chin, J. W. Engineered Triply Orthogonal Pyrrolysyl-tRNA Synthetase/tRNA Pairs Enable the Genetic Encoding of Three Distinct Non-Canonical Amino Acids. Nat. Chem. 2020, 12 (6), 535544,  DOI: 10.1038/s41557-020-0472-x
  521. 521
    Italia, J. S.; Addy, P. S.; Erickson, S. B.; Peeler, J. C.; Weerapana, E.; Chatterjee, A. Mutually Orthogonal Nonsense-Suppression Systems and Conjugation Chemistries for Precise Protein Labeling at up to Three Distinct Sites. J. Am. Chem. Soc. 2019, 141 (15), 62046212,  DOI: 10.1021/jacs.8b12954
  522. 522
    Hashimoto, K.; Fischer, E. C.; Romesberg, F. E. Efforts toward Further Integration of an Unnatural Base Pair into the Biology of a Semisynthetic Organism. J. Am. Chem. Soc. 2021, 143 (23), 86038607,  DOI: 10.1021/jacs.1c03860
  523. 523
    Ledbetter, M. P.; Karadeema, R. J.; Romesberg, F. E. Reprograming the Replisome of a Semisynthetic Organism for the Expansion of the Genetic Alphabet. J. Am. Chem. Soc. 2018, 140 (2), 758765,  DOI: 10.1021/jacs.7b11488
  524. 524
    Stucki, A.; Vallapurackal, J.; Ward, T. R.; Dittrich, P. S. Droplet Microfluidics and Directed Evolution of Enzymes: An Intertwined Journey. Angew. Chem. Int. Ed. 2021, 60 (46), 2436824387,  DOI: 10.1002/anie.202016154
  525. 525
    Wicky, B. I. M.; Milles, L. F.; Courbet, A.; Ragotte, R. J.; Dauparas, J.; Kinfu, E.; Tipps, S.; Kibler, R. D.; Baek, M.; DiMaio, F. Hallucinating Symmetric Protein Assemblies. Science 2022, 378 (6615), 5661,  DOI: 10.1126/science.add1964
  526. 526
    Wang, J.; Lisanza, S.; Juergens, D.; Tischer, D.; Watson, J. L.; Castro, K. M.; Ragotte, R.; Saragovi, A.; Milles, L. F.; Baek, M. Scaffolding Protein Functional Sites Using Deep Learning. Science 2022, 377 (6604), 387394,  DOI: 10.1126/science.abn2100
  527. 527
    Krishna, R.; Wang, J.; Ahern, W.; Sturmfels, P.; Venkatesh, P.; Kalvet, I.; Lee, G. R.; Morey-Burros, F. S.; Anishchenko, I.; Humphreys, I. R. Generalized Biomolecular Modeling and Design with RoseTTAFold All-Atom. Science 2024,  DOI: 10.1126/science.adl2528
  528. 528
    Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A. Highly Accurate Protein Structure Prediction with AlphaFold. Nature 2021, 596 (7873), 583589,  DOI: 10.1038/s41586-021-03819-2
  529. 529
    Baumann, T.; Hauf, M.; Richter, F.; Albers, S.; Möglich, A.; Ignatova, Z.; Budisa, N. Computational Aminoacyl-tRNA Synthetase Library Design for Photocaged Tyrosine. Int. J. Mol. Sci. 2019, 20 (9), 2343,  DOI: 10.3390/ijms20092343
  530. 530
    Cervettini, D.; Tang, S.; Fried, S. D.; Willis, J. C. W.; Funke, L. F. H.; Colwell, L. J.; Chin, J. W. Rapid Discovery and Evolution of Orthogonal Aminoacyl-tRNA Synthetase-tRNA Pairs. Nat. Biotechnol. 2020, 38 (8), 989999,  DOI: 10.1038/s41587-020-0479-2
  531. 531
    Taylor, C. J.; Hardy, F. J.; Burke, A. J.; Bednar, R. M.; Mehl, R. A.; Green, A. P.; Lovelock, S. L. Engineering Mutually Orthogonal PylRS/tRNA Pairs for Dual Encoding of Functional Histidine Analogues. Protein Sci. 2023, 32 (5), e4640  DOI: 10.1002/pro.4640
  532. 532
    Fredens, J.; Wang, K.; de la Torre, D.; Funke, L. F. H.; Robertson, W. E.; Christova, Y.; Chia, T.; Schmied, W. H.; Dunkelmann, D. L.; Beránek, V. Total Synthesis of Escherichia coli with a Recoded Genome. Nature 2019, 569 (7757), 514518,  DOI: 10.1038/s41586-019-1192-5
  533. 533
    Lajoie, M. J.; Rovner, A. J.; Goodman, D. B.; Aerni, H.-R.; Haimovich, A. D.; Kuznetsov, G.; Mercer, J. A.; Wang, H. H.; Carr, P. A.; Mosberg, J. A. Genomically Recoded Organisms Expand Biological Functions. Science 2013, 342 (6156), 357360,  DOI: 10.1126/science.1241459
  534. 534
    Mukai, T.; Hoshi, H.; Ohtake, K.; Takahashi, M.; Yamaguchi, A.; Hayashi, A.; Yokoyama, S.; Sakamoto, K. Highly Reproductive Escherichia coli Cells with No Specific Assignment to the UAG Codon. Sci. Rep. 2015, 5 (1), 9699,  DOI: 10.1038/srep09699
  535. 535
    Wang, T.; Liang, C.; An, Y.; Xiao, S.; Xu, H.; Zheng, M.; Liu, L.; Wang, G.; Nie, L. Engineering the Translational Machinery for Biotechnology Applications. Mol. Biotechnol. 2020, 62 (4), 219227,  DOI: 10.1007/s12033-020-00246-y
  536. 536
    Zackin, M. T.; Stieglitz, J. T.; Van Deventer, J. A. Genome-Wide Screen for Enhanced Noncanonical Amino Acid Incorporation in Yeast. ACS Synth. Biol. 2022, 11 (11), 36693680,  DOI: 10.1021/acssynbio.2c00267
  537. 537
    Exner, M. P.; Kuenzl, T.; To, T. M. T.; Ouyang, Z.; Schwagerus, S.; Hoesl, M. G.; Hackenberger, C. P. R.; Lensen, M. C.; Panke, S.; Budisa, N. Design of S-Allylcysteine in Situ Production and Incorporation Based on a Novel Pyrrolysyl-tRNA Synthetase Variant. ChemBioChem 2017, 18 (1), 8590,  DOI: 10.1002/cbic.201600537
  538. 538
    Kim, S.; Sung, B. H.; Kim, S. C.; Lee, H. S. Genetic Incorporation of L-Dihydroxyphenylalanine (DOPA) Biosynthesized by a Tyrosine Phenol-Lyase. Chem. Commun. 2018, 54 (24), 30023005,  DOI: 10.1039/C8CC00281A
  539. 539
    Nojoumi, S.; Ma, Y.; Schwagerus, S.; Hackenberger, C. P. R.; Budisa, N. In-Cell Synthesis of Bioorthogonal Alkene Tag S-Allyl-Homocysteine and Its Coupling with Reprogrammed Translation. Int. J. Mol. Sci. 2019, 20 (9), 2299,  DOI: 10.3390/ijms20092299

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 29 publications.

  1. Min Li, Yan Zhang, Kai Fu, Zhiwei Deng, Zhenbo Yuan, Zhengshan Luo, Yijian Rao. Light-Driven Deracemization by a Designed Photoenzyme. Journal of the American Chemical Society 2025, Article ASAP.
  2. Jack Toor, Wiktoria R. Grabowska, Adam L. Johnson, Jane Jones, William G. Stetler-Stevenson, Hanieh Khalili, David Peeney. Histidine Tag-Specific PEGylation Improves the Circulating Half-Life of TIMP2. ACS Applied Bio Materials 2025, 8 (3) , 1946-1955. https://doi.org/10.1021/acsabm.4c01385
  3. Alexander Zlobin, Valentina Maslova, Julia Beliaeva, Jens Meiler, Andrey Golovin. Long-Range Electrostatics in Serine Proteases: Machine Learning-Driven Reaction Sampling Yields Insights for Enzyme Design. Journal of Chemical Information and Modeling 2025, 65 (4) , 2003-2013. https://doi.org/10.1021/acs.jcim.4c01827
  4. Lei Cheng, Zhiyu Bo, Benjamin Krohn-Hansen, Yang Yang. Directed Evolution and Unusual Protonation Mechanism of Pyridoxal Radical C–C Coupling Enzymes for the Enantiodivergent Photobiocatalytic Synthesis of Noncanonical Amino Acids. Journal of the American Chemical Society 2025, 147 (5) , 4602-4612. https://doi.org/10.1021/jacs.4c16716
  5. Lars Longwitz, Marijn D. Kamer, Bart Brouwer, Andy-Mark W. H. Thunnissen, Gerard Roelfes. Boron Designer Enzyme with a Hybrid Catalytic Dyad. ACS Catalysis 2024, 14 (24) , 18469-18476. https://doi.org/10.1021/acscatal.4c06052
  6. Emilie Wheatley, Heorhii Melnychenko, Mattia Silvi. Iterative One-Carbon Homologation of Unmodified Carboxylic Acids. Journal of the American Chemical Society 2024, 146 (50) , 34285-34291. https://doi.org/10.1021/jacs.4c13630
  7. Michael L. Pigula, Yahui Ban, Hengyao You, Peter G. Schultz. A Genetically Encoded Redox-Active Nicotinamide Amino Acid. Biochemistry 2024, 63 (24) , 3184-3188. https://doi.org/10.1021/acs.biochem.4c00530
  8. Briana R. Lino, Sean J. Williams, Michelle E. Castor, James A. Van Deventer. Reaching New Heights in Genetic Code Manipulation with High Throughput Screening. Chemical Reviews 2024, 124 (21) , 12145-12175. https://doi.org/10.1021/acs.chemrev.4c00329
  9. Shu-Bin Mou, Kai-Yue Chen, Thittaya Kunthic, Zheng Xiang. Design and Evolution of an Artificial Friedel–Crafts Alkylation Enzyme Featuring an Organoboronic Acid Residue. Journal of the American Chemical Society 2024, 146 (39) , 26676-26686. https://doi.org/10.1021/jacs.4c03795
  10. Manman Li, Guohua Xu, Zhou Gong, Qiong Wu, Ling Jiang, Conggang Li. Simultaneous measurement of multiple fluorine labelling effect on GB1 stability by 19F NMR. Talanta 2025, 292 , 127959. https://doi.org/10.1016/j.talanta.2025.127959
  11. Xiaowei Gao, Yingjie Sun, Yanhong Yang, Xiu Yang, Qiuyu Liu, Xiurong Guo, Lijuan Wu, Qin Wang. Directed evolution of hydroxylase XcP4H for enhanced 5-HTP production in engineered probiotics to treat depression. International Journal of Biological Macromolecules 2025, 307 , 142250. https://doi.org/10.1016/j.ijbiomac.2025.142250
  12. Hanbing Yu, Siyi Li, Jingyang Liu, Ning Zhao, Zuodi Liang, Lele Huang, Tianyi Sun, Hongyang Zhang. Optimization of D-amino acid oxidase and epoxide hydrolase through high-performance characterization for targeted cancer therapy in sustainable biocatalysis. International Journal of Biological Macromolecules 2025, 309 , 142784. https://doi.org/10.1016/j.ijbiomac.2025.142784
  13. Reuben B. Leveson‐Gower. Designing Enzymatic Reactivity with an Expanded Palette. ChemBioChem 2025, 51 https://doi.org/10.1002/cbic.202500076
  14. Zibo Lin, Yuanyuan Liao, Donglei Tian, Junyu Liao, Qiong Chen, Jun Yin. Small‐Molecule Fluorescent Probes for Butyrylcholinesterase. ChemMedChem 2025, 20 (7) https://doi.org/10.1002/cmdc.202400875
  15. Hui Liu, Jiahui Zou, Xiaotong Li, Yizhi Ge, Wei He. Drug delivery for platinum therapeutics. Journal of Controlled Release 2025, 380 , 503-523. https://doi.org/10.1016/j.jconrel.2025.02.006
  16. Dongheon Lee, Suk Min Yun, Jong-il Choi. Expanding the genetic code: In vivo approaches for incorporating non-proteinogenic monomers. Journal of Microbiology 2025, 63 (3) , e2501005. https://doi.org/10.71150/jm.2501005
  17. Qifan Wu, Xiaochen Yang, Ying Wang, Hongqiang Qin, Xun‐Cheng Su, Weimin Xuan. Site‐Selectively Accelerating the Generation of β‐Linked Residue Isoaspartate in Proteins. Angewandte Chemie 2025, https://doi.org/10.1002/ange.202500983
  18. Qifan Wu, Xiaochen Yang, Ying Wang, Hongqiang Qin, Xun‐Cheng Su, Weimin Xuan. Site‐Selectively Accelerating the Generation of β‐Linked Residue Isoaspartate in Proteins. Angewandte Chemie International Edition 2025, https://doi.org/10.1002/anie.202500983
  19. Shelby R. Anderson, Madan R. Gopal, Abigail P. Spangler, Michaela A. Jones, D'Jana R. Wyllis, Aditya M. Kunjapur. A One‐Pot Biocatalytic Cascade to Access Diverse l ‐Phenylalanine Derivatives from Aldehydes or Carboxylic Acids. ChemCatChem 2025, 17 (5) https://doi.org/10.1002/cctc.202401979
  20. Zhen Hou, Junkai Tuo, Xiaoyan Ma, Yi-Xin Huo. Recent advances in biosynthesis of non-canonical amino acids and their potentials in strain engineering. Results in Engineering 2025, 25 , 103641. https://doi.org/10.1016/j.rineng.2024.103641
  21. Ru Jiang, Fabrizio Casilli, Andy‐Mark W. H. Thunnissen, Gerard Roelfes. An Artificial Copper‐Michaelase Featuring a Genetically Encoded Bipyridine Ligand for Asymmetric Additions to Nitroalkenes. Angewandte Chemie 2025, https://doi.org/10.1002/ange.202423182
  22. Ru Jiang, Fabrizio Casilli, Andy‐Mark W. H. Thunnissen, Gerard Roelfes. An Artificial Copper‐Michaelase Featuring a Genetically Encoded Bipyridine Ligand for Asymmetric Additions to Nitroalkenes. Angewandte Chemie International Edition 2025, https://doi.org/10.1002/anie.202423182
  23. Caitlin L. Gare, Andrew M. White, Lara R. Malins. From lead to market: chemical approaches to transform peptides into therapeutics. Trends in Biochemical Sciences 2025, 20 https://doi.org/10.1016/j.tibs.2025.01.009
  24. Manman Li, Guohua Xu, Qiong Wu, Ling Jiang, Conggang Li. 19F NMR study of proteins with parallel incorporation of multiple fluorinated aromatic amino acids. Magnetic Resonance Letters 2025, 253 , 200193. https://doi.org/10.1016/j.mrl.2025.200193
  25. Kai‐Yue Chen, Hui Ming, He‐Xiang Wang, Hua‐Qi Wang, Zheng Xiang. Genetic Incorporation of a Thioxanthone‐Containing Amino Acid for the Design of Artificial Photoenzymes. Angewandte Chemie International Edition 2025, 64 (5) https://doi.org/10.1002/anie.202419022
  26. Kai‐Yue Chen, Hui Ming, He‐Xiang Wang, Hua‐Qi Wang, Zheng Xiang. Genetic Incorporation of a Thioxanthone‐Containing Amino Acid for the Design of Artificial Photoenzymes. Angewandte Chemie 2025, 137 (5) https://doi.org/10.1002/ange.202419022
  27. Dyuti Bhandary, Sam P. de Visser, Gourab Mukherjee. Implications of non-native metal substitution in carbonic anhydrase – engineered enzymes and models. Chemical Communications 2025, 61 (4) , 612-626. https://doi.org/10.1039/D4CC05003G
  28. Angelica Graciano, Aimin Liu. Protein-derived cofactors: chemical innovations expanding enzyme catalysis. Chemical Society Reviews 2025, 98 https://doi.org/10.1039/D4CS00981A
  29. Hafiz Muhammad Zohaib, Madiha Saqlain, Maroof Ahmad Khan, Sara Masood, Ijaz Gul, Muhammad Irfan, Hui Li. Exploring enantioselective recognition of dTMP-Co-bpe coordination polymer for natural amino acids using molecular simulations and circular dichroism. Dalton Transactions 2024, 53 (31) , 13076-13086. https://doi.org/10.1039/D4DT01245C

Chemical Reviews

Cite this: Chem. Rev. 2024, 124, 14, 8740–8786
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.chemrev.4c00120
Published July 3, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

11k

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. NcAAs discussed in this review. (A) NcAAs incorporated via selective pressure incorporation (SPI), expressed protein ligation (EPL), or solid-phase peptide synthesis (SPPS). (B) NcAAs incorporated by GCE. The orthogonal translation system(s) used to incorporate each ncAA are listed. For several ncAAs, multiple incorporation techniques are discussed in this review, and these are also listed. DAP is incorporated as a precursor featuring a photocleavable group, which matures to DAP upon irradiation at 365 nm. 4-NH2Phe is incorporated as 4-AzPhe, which is then chemically reduced in situ to form 4-NH2Phe.

    Figure 2

    Figure 2. SPI of ncAAs. SPI employs an auxotrophic expression system to globally replace a target canonical amino acid (cAA) with a close structural analogue. An endogenous aaRS loads its cognate tRNA with the ncAA which is incorporated into proteins. Created with BioRender.com.

    Figure 3

    Figure 3. Strategies for the generation of ncAA-loaded tRNAs employ either chemoenzymatic methods (top left, PDB: 2C5U (102)) or Flexizymes (bottom left, PDB: 3CUN (103)). These ncAA-tRNAs can then be incorporated into a polypeptide chain using cell-free expression (CFE) systems (right). Created with BioRender.com.

    Figure 4

    Figure 4. Positive and negative selection processes can be used to engineer orthogonal aaRS-tRNA pairs to improve incorporation efficiency and/or specificity. The engineered aaRS catalyzes an aminoacylation reaction between its cognate tRNA and ncAA, with the ncAA added to the growing polypeptide chain during translation in response to a repurposed codon (e.g., the amber stop codon, UAG). Created with BioRender.com.

    Figure 5

    Figure 5. DAP incorporation into Valinomycin synthetase. (A) Genetically encoded (2S)-2-amino-3-([(2-[1-(6-nitrobenzo[d][1,3]dioxol-5-yl)ethyl]thio)ethoxy)carbonyl] ncAA is photodeprotected by irradiation at 365 nm to give DAP, which forms stable acyl-enzyme intermediates with an amide bond that is resistant to hydrolysis. (B) The active site of Valinomycin synthetase (protein shown as a gray cartoon, PDB: 6ECE (213)) with a noncanonical DAP nucleophile in position 2463 (atom-colored sticks, brown carbons) bound to a dodecadepsipeptide substrate (atom colored sticks, blue carbons). (C) Large structural differences are observed in the lid region of Valinomycin synthetase when bound to a dodecadepsipeptidyl intermediate (gray cartoon, PDB: 6ECE (213)) in comparison to a tetradepsipeptidyl intermediate (blue cartoon, PDB: 6ECD (213)).

    Figure 6

    Figure 6. Mechanistic studies on RNRs using ncAAs have shed light on the electron transfer pathway and enabled structural characterization of the active form of the multimer. (A) A cryogenic electron microscopy structure of RNR (PDB: 6W4X (231)) in its active α2β2 form was captured using a 2,3,5-F3Tyr122 mutation. The protein chains are shown as cartoons, and GDP and TPP are shown as red and gray spheres, respectively. (B) The mechanism of RNRs, which catalyze the conversion of nucleoside di- and triphosphates to deoxynucleotides. (236) TR = thioredoxin. (C) DEER experiments provided information on the relative distances between the Tyr122 radical in the unreacted α/β pair and radicals on an N3NDP mechanistic inhibitor or radicals trapped on 3-NH2Tyr.

    Figure 7

    Figure 7. Mechanistic proposal for the FtmOx1-catalyzed hydrogen atom transfer from Tyr68 to C26•.

    Figure 8

    Figure 8. 3-ClTyr incorporation into Ketosteroid Isomerase (KSI) to tune the active site electric field. (A) The mechanism of KSI. (B) The active site of KSI (PDB: 5KP1 (254)) with the ncAA 3-ClTyr in the active site, shown with orange carbons. The protein backbone is shown as a gray cartoon. Active site residues and the substrate and transition state analogue equilenin are shown as atom-colored sticks, with gray and blue carbons, respectively. (C) The product analogue 19-nortestosterone used for VSE experiments.

    Figure 9

    Figure 9. Active site of WT NiSOD (left) and a variant with a secondary amine backbone substitution (right).

    Figure 10

    Figure 10. Electron donation to the iron center affects ferryl reactivity. (top) Cytochrome P450s are capable of hydrogen atom abstraction by the intermediate Compound I. Increased electron donation through an ncAA selenolate ligand increases the rate compared to WT P450. (bottom) Heme peroxidase compound II is reduced through proton coupled electron transfer. His to MeHis substitution decreases the electron donation to the ferryl intermediate and reduces its proton affinity, slowing the rate of compound II reduction.

    Figure 11

    Figure 11. Anaerobic X-ray crystal structures of the active sites of Human Cysteine Dioxygenase (CDO, PDB: 6N43 (306)) and CDO Tyr157F2-Tyr (PDB: 6BPR (306)) in complex with the substrate cysteine and NO. CDO and CDO Tyr157F2-Tyr are shown as cartoons in blue and gray, respectively, with key active site residues and the substrate cysteine shown as atom-colored sticks with blue and gray carbon atoms. The noncanonical F2-Tyr157 is shown with orange carbon atoms.

    Figure 12

    Figure 12. NcAA-mediated noncovalent interactions influence enzyme stability. (A) SPI of 4-R-FPro in KlenTaq DNA polymerase switches many Pro puckers from endo to exo, as illustrated by the substitution of Pro555 (left, gray carbons) to 4-R-FPro555 (right, orange carbons) (PDB: 4DLG, 4DLE (335)). (B) Evolutionary trajectory of TFLeu-incorporating CAT (orange bars) starting from WT CAT (gray bar) against the half-life of enzyme inactivation at 60 °C. (C) Structures of T4 lysozyme with canonical Tyr18 (left, gray carbons) and noncanonical 3-ClTyr18 (right, orange carbons). Glu11 and Gly28 backbone atoms shown (white carbons). Halogen bond between Gly28 backbone oxygen and 3-ClTyr18 chlorine atom indicated with a dashed line (PDB: 1L63, (340) 5V7E (339)).

    Figure 13

    Figure 13. Covalent cross-links mediated by ncAAs. (A) Cross-links generated between cAAs (black) and ncAAs (orange). Cross-linking bonds shown in gray. Top left: canonical Cys-Cys cross-link. Top right: Cys-SbuTyr cross-link. Middle left: Cys-BpAla cross-link. Middle right: amino group-4-NCSPhe cross-link. Bottom left: Cys-O-2-BeTyr cross-link. Bottom right: Cys-4-CaaPhe cross-link. (B) Structures of Cys-O-2-BeTyr cross-link (left) and Cys-4-CaaPhe cross-link (right) in Mb(H64V,V68A), with Tm increases given by one and two cross-links indicated. ncAAs shown with orange carbons and Cys with white carbons (PDB: 7SPE, 7SPH (351)).

    Figure 14

    Figure 14. NcAA-mediated enzyme immobilization. (A) Schematic representation of nonspecific enzyme immobilization, mediated by cross-linking at multiple reactive surface residues (gray circles), resulting in multiple enzyme orientations relative to the solid support, as well as enzyme–enzyme cross-linking leading to multilayer immobilization. (B) Schematic representation of site-specific enzyme immobilization, mediated by a ncAA (orange circles) incorporated site specifically, resulting in a monolayer with a single defined enzyme orientation. (C) Immobilization chemistries utilizing ncAAs (orange). From top to bottom: CuAAC, SPAAC, DOPhe–amine coupling, tetrazine-sTCO Diels–Alder cycloaddition, 3-NH2Tyr-acryloyl Diels–Alder cycloaddition, Glaser–Hay alkynyl coupling, and 4-SHPhe-BODIPY coupling.

    Figure 15

    Figure 15. Introduction of 4-AcPhe into PikC, a CYP450 enzyme, enabled biosynthetic reprogramming through allowing C(sp3)–H oxidation to occur in the absence of an amino-sugar moiety (brown).

    Figure 16

    Figure 16. Incorporation of ncAAs at various positions within P450BM3 alters the oxidation product distributions for (S)-ibuprofen-OMe and (+)-nootkatone substrates.

    Figure 17

    Figure 17. Peroxidases with MeHis proximal ligands. (A) An overlay of the crystal structures of APX2 (PDB: 1OAG (382)) and APX2 MeHis163 (PDB: 5L86 (381)). Key active site residues and the heme are shown as atom-colored sticks with gray and blue carbons, respectively. MeHis is shown with brown carbons. (B) TTN achieved by APX2 and APX2 MeHis. (C) The catalytic efficiency toward guaiacol (2-methoxyphenol) oxidation for Mb variants and horseradish peroxidase (HRP). (D) An overlay of the crystal structures of Mb (PDB: 1A6K (383)) and Mb MeHis93 (PDB: 5OJ9 (384)). The protein backbones are shown as cartoons, and key active site residues and the heme are shown as atom-colored sticks with gray and blue carbons, respectively. MeHis93 is shown with brown carbon atoms.

    Figure 18

    Figure 18. Biocatalytic cyclopropanations by Mb* MeHis93. (A) The bridged ion carbenoid intermediate observed by X-ray crystallography (PDB: 6F17 (386)). A 2FO–FC map contoured at 1.5 σ is shown around the bridged carbenoid intermediate and the iron atom. (B) The cyclopropanation reaction catalyzed by engineered Mbs. (C) The non-native cofactor and MeHis ligand used to expand the scope of biocatalytic cyclopropanations. (388)

    Figure 19

    Figure 19. Introduction of 4-AzPhe into selected sites of formate dehydrogenase (FDH) and mannitol dehydrogenase (MNDH) created bioorthogonal handles for SPAAC conjugation to either a heterobifunctional linker harboring a tetrazine handle or an alternative linker with a cyclooctene handle (PDB: 3WR5, (407) 1LJ8 (408)). FDH and MNDH are shown as gray and blue cartoons, respectively. The sites of 4-AzPhe incorporation are shown as red spheres.

    Figure 20

    Figure 20. Introduction of a photocaged ncAA into a DNA polymerase through GCE occludes the active site, preventing the diffusion of nucleotides for extension. Brief irradiation with UV light cleaves the O-NB moiety to reveal the catalytic Tyr and restore polymerase activity. Created with BioRender.com.

    Figure 21

    Figure 21. Photoresponsive ncAAs used in the allosteric light regulation of ImGPS. AzoPhe undergoes light induced reversible E/Z isomerizations enabling on–off switching of HisH activity.

    Figure 22

    Figure 22. Introduction of a pair of BpyAlas into Pfu POP (PDB: 5T88 (458)) enabled inhibition of protease activity when incubated in divalent metal salts. Metal binding of the noncanonical ligands holds POP in a closed inactive conformation, which can be released through chelation of metal ions with EDTA addition, thereby allowing reversible allosteric control of biocatalyst activity. Created with BioRender.com.

    Figure 23

    Figure 23. Catalytic metal-coordinating ncAAs. (A) Crystal structure of dimeric LmrR, with the positions Val15, Met89, and Trp96 in the binding pocket shown with blue carbons (PDB: 3F8B (474)). (B) BpyAla-coordinated Cu(II) complex which activates 1-(1-methyl-1H-imidazol-2-yl)but-2-en-1-one toward nucleophilic attack. (C) Schemes of vinylogous Friedel–Crafts alkylations (top) and α,β-unsaturated 2-acyl pyridine hydrations (bottom) catalyzed by BpyAla-Cu(II) or 3-HqAla-Cu(II) metalloenzymes.

    Figure 24

    Figure 24. 4-AzPhe-anchored metalloenzymes. (A) BCN-Derivatised dirhodium complex. OAc = acetate anion. (B) Crystal structure of POP, with positions of 4-AzPhe incorporation (orange spheres) and pore-opening alanine mutations (blue spheres) shown (PDB: 5T88 (479)). (C) Schemes of styrene cyclopropanations (top) and the diazo cross-coupling cascade (bottom) catalyzed by POP variants containing 4-AzPhe-tethered dirhodium complexes.

    Figure 25

    Figure 25. Nucleophilic catalysis utilizing MeHis. (A) Scheme of ester hydrolysis, showing the reactive covalent intermediate formed between the substrate and MeHis23 (orange). (B) Structure of OE1.3, with MeHis23 (orange carbons) and sites of mutations installed during evolution (blue spheres) shown (PDB: 6Q7Q (486)). (C) Scheme highlighting the proton transfer role of Glu26 (gray) in the evolved MBHase BHMeHis1.8. Intermediates 2 (left) and 3 (right) are shown, covalently bound to MeHis23 (orange).

    Figure 26

    Figure 26. Nucleophilic catalysis utilizing 4-NH2Phe. (A) Scheme of hydrazone (X = N) and oxime (X = O) formations catalyzed by 4-NH2Phe (orange) incorporated into LmrR, with the covalent adduct formed by the carbonyl substrate and 4-NH2Phe15 shown. (B) Scheme of vinylogous Friedel–Crafts alkylations catalyzed by LmrR_V15_4-NH2Phe_RGN, with the activated imine intermediate formed between 4-NH2Phe15 (orange) and the aldehyde substrate shown. At the end of the reaction time NaBH4 is added to reduce the enzymatic product to the corresponding alcohol (right).

    Figure 27

    Figure 27. [2 + 2] Photocycloadditions catalyzed by BpAla. (A) Schemes of intramolecular [2 + 2] photocycloadditions of derivatized quinolones (top) and indoles (bottom). X = O or C, n = 1 or 2. (B) Crystal structure of EnT1.3 with product (green carbons) bound between BpAla (orange carbons), Trp244, and His287 (blue carbons) (PDB: 7ZP7 (28)).

    Figure 28

    Figure 28. Metal-dependent ncAA-incorporating photoenzymes. (A) Chromophore autocatalytically generated in sfYFP and in PSP2, which incorporates BpAla (orange side chain) at position 66. (B) Structure of PSP2, with a chromophore shown (backbone indicated with gray carbons, BpAla side chain with orange carbons). The Cys95 site of nickel–terpyridine complex ligation is shown in dark gray (PDB: 5YR3 (506)). (C) Scheme of dehalogenation reactions catalyzed by BpAla-incorporating PSP2T2 or by BpyAla-incorporating Mb. X = Cl, Br, or I. (D) Structure of Mb incorporating BpyAla (orange carbons) and with an iridium photocatalyst (green carbons) ligated to Cys45 (gray carbons) (PDB: 7YLK (516)).

  • References


    This article references 539 other publications.

    1. 1
      Punekar, N. S. ENZYMES: Catalysis, Kinetics and Mechanisms; Springer Nature Singapore, 2018. DOI: 10.1007/978-981-13-0785-0
    2. 2
      Richter, M. Functional Diversity of Organic Molecule Enzyme Cofactors. Nat. Prod. Rep. 2013, 30 (10), 13241345,  DOI: 10.1039/c3np70045c
    3. 3
      Wagner, C. R.; Benkovic, S. J. Site Directed Mutagenesis: a Tool for Enzyme Mechanism Dissection. Trends Biotechnol. 1990, 8, 263270,  DOI: 10.1016/0167-7799(90)90189-5
    4. 4
      Arnold, F. H. Directed Evolution: Bringing New Chemistry to Life. Angew. Chem. Int. Ed. 2018, 57 (16), 41434148,  DOI: 10.1002/anie.201708408
    5. 5
      Turner, N. J. Directed Evolution Drives the Next Generation of Biocatalysts. Nat. Chem. Biol. 2009, 5, 567573,  DOI: 10.1038/nchembio.203
    6. 6
      Dauparas, J.; Anishchenko, I.; Bennett, N.; Bai, H.; Ragotte, R. J.; Milles, L. F.; Wicky, B. I. M.; Courbet, A.; de Haas, R. J.; Bethel, N. Robust Deep Learning-Based Protein Sequence Design Using ProteinMPNN. Science 2022, 378 (6615), 4956,  DOI: 10.1126/science.add2187
    7. 7
      Watson, J. L.; Juergens, D.; Bennett, N. R.; Trippe, B. L.; Yim, J.; Eisenach, H. E.; Ahern, W.; Borst, A. J.; Ragotte, R. J.; Milles, L. F. De Novo Design of Protein Structure and Function with RFdiffusion. Nature 2023, 620, 10891100,  DOI: 10.1038/s41586-023-06415-8
    8. 8
      Sumida, K. H.; Núñez-Franco, R.; Kalvet, I.; Pellock, S. J.; Wicky, B. I. M.; Milles, L. F.; Dauparas, J.; Wang, J.; Kipnis, Y.; Jameson, N. Improving Protein Expression, Stability, and Function with ProteinMPNN. J. Am. Chem. Soc. 2024, 146 (3), 20542061,  DOI: 10.1021/jacs.3c10941
    9. 9
      Gargiulo, S.; Soumillion, P. Directed Evolution for Enzyme Development in Biocatalysis. Curr. Opin. Chem. Biol. 2021, 61, 107113,  DOI: 10.1016/j.cbpa.2020.11.006
    10. 10
      Planas-Iglesias, J.; Marques, S. M.; Pinto, G. P.; Musil, M.; Stourac, J.; Damborsky, J.; Bednar, D. Computational Design of Enzymes for Biotechnological Applications. Biotechnol. Adv. 2021, 47, 107696,  DOI: 10.1016/j.biotechadv.2021.107696
    11. 11
      Świderek, K.; Tuñón, I.; Moliner, V.; Bertran, J. Computational Strategies for the Design of New Enzymatic Functions. Arch. Biochem. Biophys. 2015, 582, 6879,  DOI: 10.1016/j.abb.2015.03.013
    12. 12
      Renata, H.; Wang, Z. J.; Arnold, F. H. Expanding the Enzyme Universe: Accessing Non-Natural Reactions by Mechanism-Guided Directed Evolution. Angew. Chem. Int. Ed. 2015, 54 (11), 33513367,  DOI: 10.1002/anie.201409470
    13. 13
      Pagar, A. D.; Patil, M. D.; Flood, D. T.; Yoo, T. H.; Dawson, P. E.; Yun, H. Recent Advances in Biocatalysis with Chemical Modification and Expanded Amino Acid Alphabet. Chem. Rev. 2021, 121 (10), 61736245,  DOI: 10.1021/acs.chemrev.0c01201
    14. 14
      Zhao, J.; Burke, A. J.; Green, A. P. Enzymes with Noncanonical Amino Acids. Curr. Opin. Chem. Biol. 2020, 55, 136144,  DOI: 10.1016/j.cbpa.2020.01.006
    15. 15
      Huang, Y.; Liu, T. Therapeutic Applications of Genetic Code Expansion. Synth. Syst. Biotechnol. 2018, 3 (3), 150158,  DOI: 10.1016/j.synbio.2018.09.003
    16. 16
      Icking, L.-S.; Riedlberger, A. M.; Krause, F.; Widder, J.; Frederiksen; Anne, S.; Stockert, F.; Spädt, M.; Edel, N.; Armbruster, D.; Forlani, G. iNClusive: a Database Collecting Useful Information on Non-Canonical Amino Acids and Their Incorporation into Proteins for Easier Genetic Code Expansion Implementation. Nucleic Acids Res. 2024, 52, D476,  DOI: 10.1093/nar/gkad1090
    17. 17
      Chatterjee, A.; Guo, J.; Lee, H. S.; Schultz, P. G. A Genetically Encoded Fluorescent Probe in Mammalian Cells. J. Am. Chem. Soc. 2013, 135 (34), 1254012543,  DOI: 10.1021/ja4059553
    18. 18
      Evans, H. T.; Benetatos, J.; van Roijen, M.; Bodea, L. G.; Götz, J. Decreased Synthesis of Ribosomal Proteins in Tauopathy Revealed by Non-Canonical Amino Acid Labelling. EMBO J. 2019, 38 (13), e101174  DOI: 10.15252/embj.2018101174
    19. 19
      Schultz, K. C.; Supekova, L.; Ryu, Y.; Xie, J.; Perera, R.; Schultz, P. G. A Genetically Encoded Infrared Probe. J. Am. Chem. Soc. 2006, 128 (43), 1398413985,  DOI: 10.1021/ja0636690
    20. 20
      Goettig, P.; Koch, N. G.; Budisa, N. Non-Canonical Amino Acids in Analyses of Protease Structure and Function. Int. J. Mol. Sci. 2023, 24 (18), 14035,  DOI: 10.3390/ijms241814035
    21. 21
      Tinzl, M.; Hilvert, D. Trapping Transient Protein Species by Genetic Code Expansion. ChemBioChem 2021, 22 (1), 9299,  DOI: 10.1002/cbic.202000523
    22. 22
      Birch-Price, Z.; Taylor, C. J.; Ortmayer, M.; Green, A. P. Engineering Enzyme Activity Using an Expanded Amino Acid Alphabet. Protein Eng. Des. Sel. 2023, 36, gzac013,  DOI: 10.1093/protein/gzac013
    23. 23
      Giger, S.; Buller, R. Advances in Noncanonical Amino Acid Incorporation for Enzyme Engineering Applications. CHIMIA 2023, 77 (6), 395402,  DOI: 10.2533/chimia.2023.395
    24. 24
      Hayashi, T.; Hilvert, D.; Green, A. P. Engineered Metalloenzymes with Non-Canonical Coordination Environments. Chem. Eur. J. 2018, 24 (46), 1182111830,  DOI: 10.1002/chem.201800975
    25. 25
      Lugtenburg, T.; Gran-Scheuch, A.; Drienovská, I. Non-Canonical Amino Acids as a Tool for the Thermal Stabilization of Enzymes. Protein Eng. Des. Sel. 2023, 36, gzad003,  DOI: 10.1093/protein/gzad003
    26. 26
      Mirts, E. N.; Bhagi-Damodaran, A.; Lu, Y. Understanding and Modulating Metalloenzymes with Unnatural Amino Acids, Non-Native Metal Ions, and Non-Native Metallocofactors. Acc. Chem. Res. 2019, 52 (4), 935944,  DOI: 10.1021/acs.accounts.9b00011
    27. 27
      Drienovská, I.; Roelfes, G. Expanding the Enzyme Universe with Genetically Encoded Unnatural Amino Acids. Nat. Catal. 2020, 3 (3), 193202,  DOI: 10.1038/s41929-019-0410-8
    28. 28
      Trimble, J. S.; Crawshaw, R.; Hardy, F. J.; Levy, C. W.; Brown, M. J. B.; Fuerst, D. E.; Heyes, D. J.; Obexer, R.; Green, A. P. A Designed Photoenzyme for Enantioselective [2 + 2] Cycloadditions. Nature 2022, 611 (7937), 709714,  DOI: 10.1038/s41586-022-05335-3
    29. 29
      Boutureira, O.; Bernardes, G. J. L. Advances in Chemical Protein Modification. Chem. Rev. 2015, 115 (5), 21742195,  DOI: 10.1021/cr500399p
    30. 30
      Pfleiderer, G. Chemical Modification of Proteins. In Review Articles, including those from an International Conference held in Bielefeld, FR of Germany, June 1–2, 1984 , 2021; Walter de Gruyter GmbH & Co KG, p 207.  DOI: 10.1515/9783112393048-011
    31. 31
      Gunnoo, S. B.; Madder, A. Chemical Protein Modification through Cysteine. ChemBioChem 2016, 17 (7), 529553,  DOI: 10.1002/cbic.201500667
    32. 32
      Bischak, C. G.; Longhi, S.; Snead, D. M.; Costanzo, S.; Terrer, E.; Londergan, C. H. Probing Structural Transitions in the Intrinsically Disordered C-Terminal Domain of the Measles Virus Nucleoprotein by Vibrational Spectroscopy of Cyanylated Cysteines. Biophys. J. 2010, 99 (5), 16761683,  DOI: 10.1016/j.bpj.2010.06.060
    33. 33
      Fafarman, A. T.; Webb, L. J.; Chuang, J. I.; Boxer, S. G. Site-Specific Conversion of Cysteine Thiols into Thiocyanate Creates an IR Probe for Electric Fields in Proteins. J. Am. Chem. Soc. 2006, 128 (41), 1335613357,  DOI: 10.1021/ja0650403
    34. 34
      Weeks, C. L.; Jo, H.; Kier, B.; DeGrado, W. F.; Spiro, T. G. Cysteine-Linked Aromatic Nitriles as UV Resonance Raman Probes of Protein Structure. J. Raman Spectrosc. 2012, 43 (9), 12441249,  DOI: 10.1002/jrs.3167
    35. 35
      Boutureira, O.; Bernardes, G. J.; D’Hooge, F.; Davis, B. G. Direct Radiolabelling of Proteins at Cysteine Using [18 F]-Fluorosugars. Chem. Commun. 2011, 47 (36), 1001010012,  DOI: 10.1039/c1cc13524d
    36. 36
      Kaiser, E. T.; Lawrence, D. S. Chemical Mutation of Enzyme Active Sites. Science 1984, 226 (4674), 505511,  DOI: 10.1126/science.6238407
    37. 37
      Levine, H. L.; Kaiser, E. Oxidation of Dihydronicotinamides by Flavopapain. J. Am. Chem. Soc. 1978, 100 (24), 76707677,  DOI: 10.1021/ja00492a040
    38. 38
      Levine, H. L.; Nakagawa, Y.; Kaiser, E. Flavopapain: Synthesis and Properties of Semi-Synthetic Enzymes. Biochem. Biophys. Res. Commun. 1977, 76 (1), 6470,  DOI: 10.1016/0006-291X(77)91668-0
    39. 39
      Mayer, C.; Gillingham, D. G.; Ward, T. R.; Hilvert, D. An Artificial Metalloenzyme for Olefin Metathesis. Chem. Commun. 2011, 47 (44), 1206812070,  DOI: 10.1039/c1cc15005g
    40. 40
      den Heeten, R.; Muñoz, B. K.; Popa, G.; Laan, W.; Kamer, P. C. J. Synthesis of Hybrid Transition-Metalloproteins via Thiol-Selective Covalent Anchoring of Rh-Phosphine and Ru-Phenanthroline Complexes. Dalton Trans. 2010, 39 (36), 84778483,  DOI: 10.1039/c0dt00239a
    41. 41
      Deuss, P. J.; Popa, G.; Botting, C. H.; Laan, W.; Kamer, P. C. J. Highly Efficient and Site-Selective Phosphane Modification of Proteins through Hydrazone Linkage: Development of Artificial Metalloenzymes. Angew. Chem. Int. Ed. 2010, 49 (31), 53155317,  DOI: 10.1002/anie.201002174
    42. 42
      Jarvis, A. G.; Obrecht, L.; Deuss, P. J.; Laan, W.; Gibson, E. K.; Wells, P. P.; Kamer, P. C. J. Enzyme Activity by Design: An Artificial Rhodium Hydroformylase for Linear Aldehydes. Angew. Chem. Int. Ed. 2017, 56 (44), 1359613600,  DOI: 10.1002/anie.201705753
    43. 43
      Merrifield, R. B. Solid Phase Peptide Synthesis. I. The Synthesis of a Tetrapeptide. J. Am. Chem. Soc. 1963, 85 (14), 21492154,  DOI: 10.1021/ja00897a025
    44. 44
      Eckert, D. M.; Malashkevich, V. N.; Hong, L. H.; Carr, P. A.; Kim, P. S. Inhibiting HIV-1 Entry: Discovery of D-Peptide Inhibitors that Target the gp41 Coiled-Coil Pocket. Cell 1999, 99 (1), 103115,  DOI: 10.1016/S0092-8674(00)80066-5
    45. 45
      Kent, S.; Sohma, Y.; Liu, S.; Bang, D.; Pentelute, B.; Mandal, K. Through The Looking Glass - a New World of Proteins Enabled by Chemical Synthesis. J. Pept. Sci. 2012, 18 (7), 428436,  DOI: 10.1002/psc.2421
    46. 46
      Schumacher, T. N. M.; Mayr, L. M.; Minor, D. L.; Milhollen, M. A.; Burgess, M. W.; Kim, P. S. Identification of D-Peptide Ligands Through Mirror-Image Phage Display. Science 1996, 271 (5257), 18541857,  DOI: 10.1126/science.271.5257.1854
    47. 47
      Pech, A.; Achenbach, J.; Jahnz, M.; Schülzchen, S.; Jarosch, F.; Bordusa, F.; Klussmann, S. A Thermostable D-Polymerase for Mirror-Image PCR. Nucleic Acids Res. 2017, 45 (7), 39974005,  DOI: 10.1093/nar/gkx079
    48. 48
      Jackson, D. Y.; Burnier, J.; Quan, C.; Stanley, M.; Tom, J.; Wells, J. A. A Designed Peptide Ligase for Total Synthesis of Ribonuclease A with Unnatural Catalytic Residues. Science 1994, 266 (5183), 243247,  DOI: 10.1126/science.7939659
    49. 49
      Kaiser, E. Synthetic Approaches to Biologically Active Peptides and Proteins Including Enzymes. Acc. Chem. Res. 1989, 22 (2), 4754,  DOI: 10.1021/ar00158a001
    50. 50
      Schnölzer, M.; Kent, S. B. Constructing Proteins by Dovetailing Unprotected Synthetic Peptides: Backbone-Engineered HIV Protease. Science 1992, 256 (5054), 221225,  DOI: 10.1126/science.1566069
    51. 51
      Bode, J. W. Chemical Protein Synthesis with the α-Ketoacid-Hydroxylamine Ligation. Acc. Chem. Res. 2017, 50 (9), 21042115,  DOI: 10.1021/acs.accounts.7b00277
    52. 52
      Liu, H.; Li, X. Serine/Threonine Ligation: Origin, Mechanistic Aspects, and Applications. Acc. Chem. Res. 2018, 51 (7), 16431655,  DOI: 10.1021/acs.accounts.8b00151
    53. 53
      Muir, T. W. Semisynthesis of Proteins by Expressed Protein Ligation. Annu. Rev. Biochem. 2003, 72 (1), 249289,  DOI: 10.1146/annurev.biochem.72.121801.161900
    54. 54
      Antos, J. M.; Truttmann, M. C.; Ploegh, H. L. Recent Advances in Sortase-Catalyzed Ligation Methodology. Curr. Opin. Struct. Biol. 2016, 38, 111118,  DOI: 10.1016/j.sbi.2016.05.021
    55. 55
      Chen, S.-Y.; Cressman, S.; Mao, F.; Shao, H.; Low, D. W.; Beilan, H. S.; Cagle, E. N.; Carnevali, M.; Gueriguian, V.; Keogh, P. J. Synthetic Erythropoietic Proteins: Tuning Biological Performance by Site-Specific Polymer Attachment. Chem. Biol. 2005, 12 (3), 371383,  DOI: 10.1016/j.chembiol.2005.01.017
    56. 56
      Kochendoerfer, G. G.; Chen, S.-Y.; Mao, F.; Cressman, S.; Traviglia, S.; Shao, H.; Hunter, C. L.; Low, D. W.; Cagle, E. N.; Carnevali, M. Design and Chemical Synthesis of a Homogeneous Polymer-Modified Erythropoiesis Protein. Science 2003, 299 (5608), 884887,  DOI: 10.1126/science.1079085
    57. 57
      Kent, S. B. H. Bringing the Science of Proteins into the Realm of Organic Chemistry: Total Chemical Synthesis of SEP (Synthetic Erythropoiesis Protein). Angew. Chem. Int. Ed. 2013, 52 (46), 1198811996,  DOI: 10.1002/anie.201304116
    58. 58
      Liu, S.; Pentelute, B. L.; Kent, S. B. H. Convergent Chemical Synthesis of [Lysine24, 38, 83] Human Erythropoietin. Angew. Chem. Int. Ed. 2012, 51 (4), 993999,  DOI: 10.1002/anie.201106060
    59. 59
      Baca, M.; Kent, S. B. Catalytic Contribution of Flap-Substrate Hydrogen Bonds in ″HIV-1 Protease″ Explored by Chemical Synthesis. Proc. Natl. Acad. Sci. U.S.A. 1993, 90 (24), 1163811642,  DOI: 10.1073/pnas.90.24.11638
    60. 60
      Beadle, J. D.; Knuhtsen, A.; Hoose, A.; Raubo, P.; Jamieson, A. G.; Shipman, M. Solid-Phase Synthesis of Oxetane Modified Peptides. Org. Lett. 2017, 19 (12), 33033306,  DOI: 10.1021/acs.orglett.7b01466
    61. 61
      Abdildinova, A.; Kurth, M. J.; Gong, Y.-D. Solid-Phase Synthesis of Peptidomimetics with Peptide Backbone Modifications. Asian J. Org. Chem. 2021, 10 (9), 23002317,  DOI: 10.1002/ajoc.202100264
    62. 62
      Flavell, R. R.; Muir, T. W. Expressed Protein Ligation (EPL) in the Study of Signal Transduction, Ion Conduction, And Chromatin Biology. Acc. Chem. Res. 2009, 42 (1), 107116,  DOI: 10.1021/ar800129c
    63. 63
      Valiyaveetil, F. I.; Leonetti, M.; Muir, T. W.; MacKinnon, R. Ion Selectivity in a Semisynthetic K+ Channel Locked in the Conductive Conformation. Science 2006, 314 (5801), 10041007,  DOI: 10.1126/science.1133415
    64. 64
      Vázquez, M. E.; Nitz, M.; Stehn, J.; Yaffe, M. B.; Imperiali, B. Fluorescent Caged Phosphoserine Peptides as Probes to Investigate Phosphorylation-Dependent Protein Associations. J. Am. Chem. Soc. 2003, 125 (34), 1015010151,  DOI: 10.1021/ja0351847
    65. 65
      Müller, M. M.; Kries, H.; Csuhai, E.; Kast, P.; Hilvert, D. Design, Selection, and Characterization of a Split Chorismate Mutase. Protein Sci. 2010, 19 (5), 10001010,  DOI: 10.1002/pro.377
    66. 66
      Choi, Y.; Shin, S. H.; Jung, H.; Kwon, O.; Seo, J. K.; Kee, J.-M. Specific Fluorescent Probe for Protein Histidine Phosphatase Activity. ACS sensors 2019, 4 (4), 10551062,  DOI: 10.1021/acssensors.9b00242
    67. 67
      Sainlos, M.; Imperiali, B. Tools For Investigating Peptide-Protein Interactions: Peptide Incorporation of Environment-Sensitive Fluorophores Through SPPS-Based ’Building Block’ Approach. Nat. Protoc. 2007, 2 (12), 32103218,  DOI: 10.1038/nprot.2007.443
    68. 68
      Wu, Y.; Tam, W.-S.; Chau, H.-F.; Kaur, S.; Thor, W.; Aik, W. S.; Chan, W.-L.; Zweckstetter, M.; Wong, K.-L. Solid-Phase Fluorescent BODIPY-Peptide Synthesis via in situ Dipyrrin Construction. Chem. Sci. 2020, 11 (41), 1126611273,  DOI: 10.1039/D0SC04849F
    69. 69
      Kienhöfer, A.; Kast, P.; Hilvert, D. Selective Stabilization of the Chorismate Mutase Transition State by a Positively Charged Hydrogen Bond Donor. J. Am. Chem. Soc. 2003, 125 (11), 32063207,  DOI: 10.1021/ja0341992
    70. 70
      Roy, R. S.; Imperiali, B. Pyridoxamine-Amino Acid Chimeras in Semisynthetic Aminotransferase Mimics. Protein eng. 1997, 10 (6), 691698,  DOI: 10.1093/protein/10.6.691
    71. 71
      Lopez, G.; Anderson, J. C. Synthetic Auxotrophs with Ligand-Dependent Essential Genes for a BL21 (DE3) Biosafety Strain. ACS Synth. Biol. 2015, 4 (12), 12791286,  DOI: 10.1021/acssynbio.5b00085
    72. 72
      Budisa, N. Amino Acids and Codons - Code Organization and Protein Structure. Engineering the Genetic Code 2005, 6689,  DOI: 10.1002/3527607188.ch4
    73. 73
      Budisa, N.; Steipe, B.; Demange, P.; Eckerskorn, C.; Kellermann, J.; Huber, R. High-Level Biosynthetic Substitution of Methionine in Proteins by Its Analogs 2-Aminohexanoic Acid, Selenomethionine, Telluromethionine and Ethionine in Escherichia coli. Eur. J. Biochem. 1995, 230 (2), 788796,  DOI: 10.1111/j.1432-1033.1995.0788h.x
    74. 74
      Cowie, D. B.; Cohen, G. N. Biosynthesis by Escherichia coli of Active Altered Proteins Containing Selenium Instead of Sulfur. Biochim. Biophys. Acta 1957, 26 (2), 252261,  DOI: 10.1016/0006-3002(57)90003-3
    75. 75
      Wong, J. Membership Mutation of the Genetic Code: Loss of Fitness by Tryptophan. Proc. Natl. Acad. Sci. U.S.A. 1983, 80 (20), 63036306,  DOI: 10.1073/pnas.80.20.6303
    76. 76
      Dóring, V.; Mootz, H. D.; Nangle, L. A.; Hendrickson, T. L.; de Crécy-Lagard, V.; Schimmel, P.; Marliere, P. Enlarging the Amino Acid Set of Escherichia coli by Infiltration of the Valine Coding Pathway. Science 2001, 292 (5516), 501504,  DOI: 10.1126/science.1057718
    77. 77
      Datta, D.; Wang, P.; Carrico, I. S.; Mayo, S. L.; Tirrell, D. A. A Designed Phenylalanyl-tRNA Synthetase Variant Allows Efficient in Vivo Incorporation of Aryl Ketone Functionality into Proteins. J. Am. Chem. Soc. 2002, 124 (20), 56525653,  DOI: 10.1021/ja0177096
    78. 78
      Sharma, N.; Furter, R.; Kast, P.; Tirrell, D. A. Efficient Introduction of Aryl Bromide Functionality into Proteins in vivo. FEBS Lett. 2000, 467 (1), 3740,  DOI: 10.1016/S0014-5793(00)01120-0
    79. 79
      Kiick, K. L.; van Hest, J. C.; Tirrell, D. A. Expanding the Scope of Protein Biosynthesis by Altering the Methionyl-tRNA Synthetase Activity of a Bacterial Expression Host. Angew. Chem. Int. Ed. 2000, 39 (12), 21482152,  DOI: 10.1002/1521-3773(20000616)39:12<2148::AID-ANIE2148>3.0.CO;2-7
    80. 80
      Hendrickson, W. A.; Horton, J. R.; LeMaster, D. M. Selenomethionyl Proteins Produced for Analysis by Multiwavelength Anomalous Diffraction (MAD): a Vehicle for Direct Determination of Three-Dimensional Structure. EMBO J. 1990, 9 (5), 16651672,  DOI: 10.1002/j.1460-2075.1990.tb08287.x
    81. 81
      Boles, J. O.; Lewinski, K.; Kunkle, M.; Odom, J. D.; Dunlap, R. B.; Lebioda, L.; Hatada, M. Bio-Incorporation of Telluromethionine into Buried Residues of Dihydrofolate Reductase. Nat. Struct. Biol. 1994, 1 (5), 283284,  DOI: 10.1038/nsb0594-283
    82. 82
      Strub, M. P.; Hoh, F.; Sanchez, J. F.; Strub, J. M.; Böck, A.; Aumelas, A.; Dumas, C. Selenomethionine and Selenocysteine Double Labeling Strategy for Crystallographic Phasing. Structure 2003, 11 (11), 13591367,  DOI: 10.1016/j.str.2003.09.014
    83. 83
      Budisa, N.; Karnbrock, W.; Steinbacher, S.; Humm, A.; Prade, L.; Neuefeind, T.; Moroder, L.; Huber, R. Bioincorporation of Telluromethionine into Proteins: A Promising New Approach for X-Ray Structure Analysis of Proteins. J. Mol. Biol. 1997, 270 (4), 616623,  DOI: 10.1006/jmbi.1997.1132
    84. 84
      Bae, J. H.; Alefelder, S.; Kaiser, J. T.; Friedrich, R.; Moroder, L.; Huber, R.; Budisa, N. Incorporation of β-Selenolo[3,2-b]Pyrrolyl-Alanine into Proteins for Phase Determination in Protein X-Ray Crystallography. J. Mol. Biol. 2001, 309 (4), 925936,  DOI: 10.1006/jmbi.2001.4699
    85. 85
      Minks, C.; Huber, R.; Moroder, L.; Budisa, N. Atomic Mutations at the Single Tryptophan Residue of Human Recombinant Annexin V: Effects on Structure, Stability, and Activity. Biochemistry 1999, 38 (33), 1064910659,  DOI: 10.1021/bi990580g
    86. 86
      Renner, C.; Alefelder, S.; Bae, J. H.; Budisa, N.; Huber, R.; Moroder, L. Fluoroprolines as Tools for Protein Design and Engineering. Angew. Chem. Int. Ed. 2001, 40 (5), 923925,  DOI: 10.1002/1521-3773(20010302)40:5<923::AID-ANIE923>3.0.CO;2-#
    87. 87
      Tang, Y.; Ghirlanda, G.; Petka, W. A.; Nakajima, T.; DeGrado, W. F.; Tirrell, D. A. Fluorinated Coiled-Coil Proteins Prepared in Vivo Display Enhanced Thermal and Chemical Stability. Angew. Chem. Int. Ed. 2001, 40 (8), 14941496,  DOI: 10.1002/1521-3773(20010417)40:8<1494::AID-ANIE1494>3.0.CO;2-X
    88. 88
      Tang, Y.; Tirrell, D. A. Biosynthesis of a Highly Stable Coiled-Coil Protein Containing Hexafluoroleucine in an Engineered Bacterial Host. J. Am. Chem. Soc. 2001, 123 (44), 1108911090,  DOI: 10.1021/ja016652k
    89. 89
      Wang, P.; Tang, Y.; Tirrell, D. A. Incorporation of Trifluoroisoleucine into Proteins in Vivo. J. Am. Chem. Soc. 2003, 125 (23), 69006906,  DOI: 10.1021/ja0298287
    90. 90
      Montclare, J. K.; Tirrell, D. A. Evolving Proteins of Novel Composition. Angew. Chem. Int. Ed. 2006, 45 (27), 45184521,  DOI: 10.1002/anie.200600088
    91. 91
      Hyun Bae, J.; Rubini, M.; Jung, G.; Wiegand, G.; Seifert, M. H.J.; Azim, M.K.; Kim, J.-S.; Zumbusch, A.; Holak, T. A.; Moroder, L.; Huber, R.; Budisa, N. Expansion of the Genetic Code Enables Design of a Novel “Gold” Class of Green Fluorescent Proteins. J. Mol. Biol. 2003, 328 (5), 10711081,  DOI: 10.1016/S0022-2836(03)00364-4
    92. 92
      Budisa, N.; Rubini, M.; Bae, J. H.; Weyher, E.; Wenger, W.; Golbik, R.; Huber, R.; Moroder, L. Global Replacement of Tryptophan with Aminotryptophans Generates Non-Invasive Protein-Based Optical pH Sensors. Angew. Chem. Int. Ed. 2002, 41 (21), 40664069,  DOI: 10.1002/1521-3773(20021104)41:21<4066::AID-ANIE4066>3.0.CO;2-6
    93. 93
      Duewel, H. S.; Daub, E.; Robinson, V.; Honek, J. F. Elucidation of Solvent Exposure, Side-Chain Reactivity, and Steric Demands of the Trifluoromethionine Residue in a Recombinant Protein. Biochemistry 2001, 40 (44), 1316713176,  DOI: 10.1021/bi011381b
    94. 94
      Seifert, M. H. J.; Ksiazek, D.; Azim, M. K.; Smialowski, P.; Budisa, N.; Holak, T. A. Slow Exchange in the Chromophore of a Green Fluorescent Protein Variant. J. Am. Chem. Soc. 2002, 124 (27), 79327942,  DOI: 10.1021/ja0257725
    95. 95
      Bann, J. G.; Pinkner, J.; Hultgren, S. J.; Frieden, C. Real-Time and Equilibrium 19F-NMR Studies Reveal the Role of Domain-Domain Interactions in the Folding of the Chaperone PapD. Proc. Natl. Acad. Sci. U.S.A. 2002, 99 (2), 709714,  DOI: 10.1073/pnas.022649599
    96. 96
      Deming, T. J.; Fournier, M. J.; Mason, T. L.; Tirrell, D. A. Biosynthetic Incorporation and Chemical Modification of Alkene Functionality in Genetically Engineered Polymers. Journal of Macromolecular Science, Part A 1997, 34 (10), 21432150,  DOI: 10.1080/10601329708010331
    97. 97
      Kiick, K. L.; Saxon, E.; Tirrell, D. A.; Bertozzi, C. R. Incorporation of Azides into Recombinant Proteins for Chemoselective Modification by the Staudinger Ligation. Proc. Natl. Acad. Sci. U.S.A. 2002, 99 (1), 1924,  DOI: 10.1073/pnas.012583299
    98. 98
      Kothakota, S.; Mason, T. L.; Tirrell, D. A.; Fournier, M. J. Biosynthesis of a Periodic Protein Containing 3-Thienylalanine: A Step Toward Genetically Engineered Conducting Polymers. J. Am. Chem. Soc. 1995, 117 (1), 536537,  DOI: 10.1021/ja00106a064
    99. 99
      Link, A. J.; Tirrell, D. A. Cell Surface Labeling of Escherichia coli via Copper(I)-Catalyzed [3 + 2] Cycloaddition. J. Am. Chem. Soc. 2003, 125 (37), 1116411165,  DOI: 10.1021/ja036765z
    100. 100
      van Hest, J. C. M.; Kiick, K. L.; Tirrell, D. A. Efficient Incorporation of Unsaturated Methionine Analogues into Proteins in Vivo. J. Am. Chem. Soc. 2000, 122 (7), 12821288,  DOI: 10.1021/ja992749j
    101. 101
      van Hest, J. C. M.; Tirrell, D. A. Efficient introduction of alkene functionality into proteins in vivo. FEBS Lett. 1998, 428 (1–2), 6870,  DOI: 10.1016/S0014-5793(98)00489-X
    102. 102
      Omari, K. E.; Ren, J.; Bird, L. E.; Bona, M. K.; Klarmann, G.; LeGrice, S. F. J.; Stammers, D. K. Molecular Architecture and Ligand Recognition Determinants for T4 RNA Ligase. J. Biol. Chem. 2006, 281 (3), 15731579,  DOI: 10.1074/jbc.M509658200
    103. 103
      Xiao, H.; Murakami, H.; Suga, H.; Ferré-D’Amaré, A. R. Structural Basis of Specific tRNA Aminoacylation by a Small in vitro Selected Ribozyme. Nature 2008, 454 (7202), 358361,  DOI: 10.1038/nature07033
    104. 104
      Noren, C. J.; Anthony-Cahill, S. J.; Griffith, M. C.; Schultz, P. G. A General Method for Site-specific Incorporation of Unnatural Amino Acids into Proteins. Science 1989, 244 (4901), 182188,  DOI: 10.1126/science.2649980
    105. 105
      Goto, Y.; Katoh, T.; Suga, H. Flexizymes for Genetic Code Reprogramming. Nat. Protoc. 2011, 6 (6), 779790,  DOI: 10.1038/nprot.2011.331
    106. 106
      Shimizu, Y.; Inoue, A.; Tomari, Y.; Suzuki, T.; Yokogawa, T.; Nishikawa, K.; Ueda, T. Cell-Free Translation Reconstituted with Purified Components. Nat. Biotechnol. 2001, 19 (8), 751755,  DOI: 10.1038/90802
    107. 107
      Chapeville, F.; Lipmann, F.; Ehrenstein, G. v.; Weisblum, B.; Ray, W. J.; Benzer, S. On the Role of Soluble Ribonucleic Acid in Coding for Amino Acids. Proc. Natl. Acad. Sci. U.S.A. 1962, 48 (6), 10861092,  DOI: 10.1073/pnas.48.6.1086
    108. 108
      Fahnestock, S.; Rich, A. Ribosome-Catalyzed Polyester Formation. Science 1971, 173 (3994), 340343,  DOI: 10.1126/science.173.3994.340
    109. 109
      Hecht, S. M.; Alford, B. L.; Kuroda, Y.; Kitano, S. “Chemical Aminoacylation” of tRNA’s. J. Biol. Chem. 1978, 253 (13), 45174520,  DOI: 10.1016/S0021-9258(17)30417-9
    110. 110
      Katoh, T.; Goto, Y.; Passioura, T.; Suga, H. Development of Flexizyme Aminoacylation Ribozymes and Their Applications. Ribozymes 2021, 519543,  DOI: 10.1002/9783527814527.ch20
    111. 111
      Morimoto, J.; Hayashi, Y.; Iwasaki, K.; Suga, H. Flexizymes: Their Evolutionary History and the Origin of Catalytic Function. Acc. Chem. Res. 2011, 44 (12), 13591368,  DOI: 10.1021/ar2000953
    112. 112
      Murakami, H.; Ohta, A.; Ashigai, H.; Suga, H. A Highly Flexible tRNA Acylation Method for Non-Natural Polypeptide Synthesis. Nat. Methods 2006, 3 (5), 357359,  DOI: 10.1038/nmeth877
    113. 113
      Katoh, T.; Suga, H. In Vitro Genetic Code Reprogramming for the Expansion of Usable Noncanonical Amino Acids. Annu. Rev. Biochem. 2022, 91 (1), 221243,  DOI: 10.1146/annurev-biochem-040320-103817
    114. 114
      Goto, Y.; Suga, H. The RaPID Platform for the Discovery of Pseudo-Natural Macrocyclic Peptides. Acc. Chem. Res. 2021, 54 (18), 36043617,  DOI: 10.1021/acs.accounts.1c00391
    115. 115
      Schultz, P. Expanding the Genetic Code. Protein Sci. 2023, 32 (1), e4488  DOI: 10.1002/pro.4488
    116. 116
      Neumann, H.; Wang, K.; Davis, L.; Garcia-Alai, M.; Chin, J. W. Encoding Multiple Unnatural Amino Acids via Evolution of a Quadruplet-Decoding Ribosome. Nature 2010, 464 (7287), 441444,  DOI: 10.1038/nature08817
    117. 117
      Robertson, W. E.; Funke, L. F. H.; de la Torre, D.; Fredens, J.; Elliott, T. S.; Spinck, M.; Christova, Y.; Cervettini, D.; Böge, F. L.; Liu, K. C. Sense Codon Reassignment Enables Viral Resistance and Encoded Polymer Synthesis. Science 2021, 372 (6546), 10571062,  DOI: 10.1126/science.abg3029
    118. 118
      Wang, L.; Brock, A.; Herberich, B.; Schultz, P. G. Expanding the Genetic Code of Escherichia coli. Science 2001, 292 (5516), 498500,  DOI: 10.1126/science.1060077
    119. 119
      Dunkelmann, D. L.; Piedrafita, C.; Dickson, A.; Liu, K. C.; Elliott, T. S.; Fiedler, M.; Bellini, D.; Zhou, A.; Cervettini, D.; Chin, J. W. Adding α,α-Disubstituted and β-Linked Monomers to the Genetic Code of an Organism. Nature 2024, 625 (7995), 603610,  DOI: 10.1038/s41586-023-06897-6
    120. 120
      Bryson, D. I. Continuous Directed Evolution of Aminoacyl-tRNA Synthetases. Nat. Chem. Biol. 2017, 13, 1253,  DOI: 10.1038/nchembio.2474
    121. 121
      Krahn, N.; Tharp, J. M.; Crnković, A.; Söll, D. Chapter Twelve - Engineering Aminoacyl-tRNA Synthetases for Use in Synthetic Biology. In The Enzymes; Ribas de Pouplana, L., Kaguni, L. S., Eds.; Academic Press, 2020; Vol. 48, pp 351395. DOI: 10.1016/bs.enz.2020.06.004
    122. 122
      Amiram, M.; Haimovich, A. D.; Fan, C.; Wang, Y.-S.; Aerni, H.-R.; Ntai, I.; Moonan, D. W.; Ma, N. J.; Rovner, A. J.; Hong, S. H. Evolution of Translation Machinery in Recoded Bacteria Enables Multi-Site Incorporation of Nonstandard Amino Acids. Nat. Biotechnol. 2015, 33 (12), 12721279,  DOI: 10.1038/nbt.3372
    123. 123
      Hohl, A.; Karan, R.; Akal, A.; Renn, D.; Liu, X.; Ghorpade, S.; Groll, M.; Rueping, M.; Eppinger, J. Engineering a Polyspecific Pyrrolysyl-tRNA Synthetase by a High Throughput FACS Screen. Sci. Rep. 2019, 9 (1), 11971,  DOI: 10.1038/s41598-019-48357-0
    124. 124
      Wang, L.; Xie, J.; Schultz, P. G. Expanding the Genetic Code. Annu. Rev. Biophys. Biomol. Struct. 2006, 35, 225249,  DOI: 10.1146/annurev.biophys.35.101105.121507
    125. 125
      Santoro, S.; Wang, L.; Herberich, B.; King, D. S.; Schultz, P. G. An Efficient System for the Evolution of Aminoacyl-tRNA Synthetase Specificity. Nat. Biotechnol. 2002, 20, 10441048,  DOI: 10.1038/nbt742
    126. 126
      Dumas, A.; Lercher, L.; Spicer, C. D.; Davis, B. G. Designing Logical Codon Reassignment - Expanding the Chemistry in Biology. Chem. Sci. 2015, 6 (1), 5069,  DOI: 10.1039/C4SC01534G
    127. 127
      Wan, W.; Tharp, J. M.; Liu, W. R. Pyrrolysyl-tRNA Synthetase: An Ordinary Enzyme but an Outstanding Genetic Code Expansion Tool. Biochim. Biophys. Acta Proteins Proteomics 2014, 1844 (6), 10591070,  DOI: 10.1016/j.bbapap.2014.03.002
    128. 128
      Ryu, Y.; Schultz, P. G. Efficient Incorporation of Unnatural Amino Acids into Proteins in Escherichia coli. Nat. Methods 2006, 3 (4), 263265,  DOI: 10.1038/nmeth864
    129. 129
      Srinivasan, G.; James, C. M.; Krzycki, J. A. Pyrrolysine Encoded by UAG in Archaea: Charging of a UAG-Decoding Specialized tRNA. Science 2002, 296 (5572), 14591462,  DOI: 10.1126/science.1069588
    130. 130
      Hao, B.; Gong, W.; Ferguson, T. K.; James, C. M.; Krzycki, J. A.; Chan, M. K. A New UAG-Encoded Residue in the Structure of a Methanogen Methyltransferase. Science 2002, 296 (5572), 14621466,  DOI: 10.1126/science.1069556
    131. 131
      Chin, J. W. Expanding and Reprogramming the Genetic Code of Cells and Animals. Annu. Rev. Biochem. 2014, 83 (1), 379408,  DOI: 10.1146/annurev-biochem-060713-035737
    132. 132
      Ranaghan, K. E.; Hung, J. E.; Bartlett, G. J.; Mooibroek, T. J.; Harvey, J. N.; Woolfson, D. N.; van der Donk, W. A.; Mulholland, A. J. A Catalytic Role for Methionine Revealed by a Combination of Computation and Experiments on Phosphite Dehydrogenase. Chem. Sci. 2014, 5 (6), 21912199,  DOI: 10.1039/C3SC53009D
    133. 133
      Ekanayake, K. B.; Mahawaththa, M. C.; Qianzhu, H.; Abdelkader, E. H.; George, J.; Ullrich, S.; Murphy, R. B.; Fry, S. E.; Johansen-Leete, J.; Payne, R. J. Probing Ligand Binding Sites on Large Proteins by Nuclear Magnetic Resonance Spectroscopy of Genetically Encoded Non-Canonical Amino Acids. J. Med. Chem. 2023, 66 (7), 52895304,  DOI: 10.1021/acs.jmedchem.3c00222
    134. 134
      Ellman, J. A.; Volkman, B. F.; Mendel, D.; Schulz, P. G.; Wemmer, D. E. Site-Specific Isotopic Labeling of Proteins for NMR Studies. J. Am. Chem. Soc. 1992, 114 (20), 79597961,  DOI: 10.1021/ja00046a080
    135. 135
      Schmidt, M. J.; Borbas, J.; Drescher, M.; Summerer, D. A Genetically Encoded Spin Label for Electron Paramagnetic Resonance Distance Measurements. J. Am. Chem. Soc. 2014, 136 (4), 12381241,  DOI: 10.1021/ja411535q
    136. 136
      Fafarman, A. T.; Boxer, S. G. Nitrile Bonds as Infrared Probes of Electrostatics in Ribonuclease S. J. Phys. Chem. B 2010, 114 (42), 1353613544,  DOI: 10.1021/jp106406p
    137. 137
      Xie, J.; Wang, L.; Wu, N.; Brock, A.; Spraggon, G.; Schultz, P. G. The Site-Specific Incorporation of p-Iodo-L-Phenylalanine into Proteins for Structure Determination. Nat. Biotechnol. 2004, 22 (10), 12971301,  DOI: 10.1038/nbt1013
    138. 138
      Summerer, D.; Chen, S.; Wu, N.; Deiters, A.; Chin, J. W.; Schultz, P. G. A Genetically Encoded Fluorescent Amino Acid. Proc. Natl. Acad. Sci. U.S.A. 2006, 103 (26), 97859789,  DOI: 10.1073/pnas.0603965103
    139. 139
      Lang, K.; Davis, L.; Wallace, S.; Mahesh, M.; Cox, D. J.; Blackman, M. L.; Fox, J. M.; Chin, J. W. Genetic Encoding of Bicyclononynes and trans-Cyclooctenes for Site-Specific Protein Labeling in Vitro and in Live Mammalian Cells via Rapid Fluorogenic Diels-Alder Reactions. J. Am. Chem. Soc. 2012, 134 (25), 1031710320,  DOI: 10.1021/ja302832g
    140. 140
      Plass, T.; Milles, S.; Koehler, C.; Schultz, C.; Lemke, E. A. Genetically Encoded Copper-Free Click Chemistry. Angew. Chem. Int. Ed. 2011, 50 (17), 38783881,  DOI: 10.1002/anie.201008178
    141. 141
      Brustad, E. M.; Lemke, E. A.; Schultz, P. G.; Deniz, A. A. A General and Efficient Method for the Site-Specific Dual-Labeling of Proteins for Single Molecule Fluorescence Resonance Energy Transfer. J. Am. Chem. Soc. 2008, 130 (52), 1766417665,  DOI: 10.1021/ja807430h
    142. 142
      Loving, G.; Imperiali, B. A Versatile Amino Acid Analogue of the Solvatochromic Fluorophore 4-N,N-Dimethylamino-1,8-naphthalimide: A Powerful Tool for the Study of Dynamic Protein Interactions. J. Am. Chem. Soc. 2008, 130 (41), 1363013638,  DOI: 10.1021/ja804754y
    143. 143
      Mendes, K. R.; Martinez, J. A.; Kantrowitz, E. R. Asymmetric Allosteric Signaling in Aspartate Transcarbamoylase. ACS Chem. Biol. 2010, 5 (5), 499506,  DOI: 10.1021/cb9003207
    144. 144
      Dean, S. F.; Whalen, K. L.; Spies, M. A. Biosynthesis of a Novel Glutamate Racemase Containing a Site-Specific 7-Hydroxycoumarin Amino Acid: Enzyme-Ligand Promiscuity Revealed at the Atomistic Level. ACS Cent. Sci. 2015, 1 (7), 364373,  DOI: 10.1021/acscentsci.5b00211
    145. 145
      Wang, J.; Xie, J.; Schultz, P. G. A Genetically Encoded Fluorescent Amino Acid. J. Am. Chem. Soc. 2006, 128 (27), 87388739,  DOI: 10.1021/ja062666k
    146. 146
      Li, M.; Peng, T. Genetic Encoding of a Fluorescent Noncanonical Amino Acid as a FRET Donor for the Analysis of Deubiquitinase Activities. In Genetically Incorporated Non-Canonical Amino Acids: Methods and Protocols; Tsai, Y.-H., Elsässer, S. J., Eds.; Springer: US, 2023; pp 55- 67. DOI: 10.1007/978-1-0716-3251-2_4
    147. 147
      Miyake-Stoner, S. J.; Miller, A. M.; Hammill, J. T.; Peeler, J. C.; Hess, K. R.; Mehl, R. A.; Brewer, S. H. Probing Protein Folding Using Site-Specifically Encoded Unnatural Amino Acids as FRET Donors with Tryptophan. Biochemistry 2009, 48 (25), 59535962,  DOI: 10.1021/bi900426d
    148. 148
      Bergfors, T. M. Protein crystallization. Internat’l University Line 2009. ISBN: 978–0-9720774–4-6.
    149. 149
      Sakamoto, K.; Murayama, K.; Oki, K.; Iraha, F.; Kato-Murayama, M.; Takahashi, M.; Ohtake, K.; Kobayashi, T.; Kuramitsu, S.; Shirouzu, M. Genetic Encoding of 3-Iodo-l-Tyrosine in Escherichia coli for Single-Wavelength Anomalous Dispersion Phasing in Protein Crystallography. Structure 2009, 17 (3), 335344,  DOI: 10.1016/j.str.2009.01.008
    150. 150
      Lee, H. S.; Spraggon, G.; Schultz, P. G.; Wang, F. Genetic Incorporation of a Metal-Ion Chelating Amino Acid into Proteins as a Biophysical Probe. J. Am. Chem. Soc. 2009, 131 (7), 24812483,  DOI: 10.1021/ja808340b
    151. 151
      Nogly, P.; Weinert, T.; James, D.; Carbajo, S.; Ozerov, D.; Furrer, A.; Gashi, D.; Borin, V.; Skopintsev, P.; Jaeger, K. Retinal Isomerization in Bacteriorhodopsin Captured by a Femtosecond X-Ray Laser. Science 2018, 361 (6398), eaat0094  DOI: 10.1126/science.aat0094
    152. 152
      Tenboer, J.; Basu, S.; Zatsepin, N.; Pande, K.; Milathianaki, D.; Frank, M.; Hunter, M.; Boutet, S.; Williams, G. J.; Koglin, J. E. Time-Resolved Serial Crystallography Captures High-Resolution Intermediates of Photoactive Yellow Protein. Science 2014, 346 (6214), 12421246,  DOI: 10.1126/science.1259357
    153. 153
      Suga, M.; Akita, F.; Sugahara, M.; Kubo, M.; Nakajima, Y.; Nakane, T.; Yamashita, K.; Umena, Y.; Nakabayashi, M.; Yamane, T. Light-Induced Structural Changes and the Site of O = O Bond Formation in PSII Caught by XFEL. Nature 2017, 543 (7643), 131135,  DOI: 10.1038/nature21400
    154. 154
      Dods, R.; Båth, P.; Morozov, D.; Gagnér, V. A.; Arnlund, D.; Luk, H. L.; Kübel, J.; Maj, M.; Vallejos, A.; Wickstrand, C. Ultrafast Structural Changes within a Photosynthetic Reaction Centre. Nature 2021, 589 (7841), 310314,  DOI: 10.1038/s41586-020-3000-7
    155. 155
      Chapman, H. N. X-Ray Free-Electron Lasers for the Structure and Dynamics of Macromolecules. Annu. Rev. Biochem. 2019, 88 (1), 3558,  DOI: 10.1146/annurev-biochem-013118-110744
    156. 156
      Kern, J.; Chatterjee, R.; Young, I. D.; Fuller, F. D.; Lassalle, L.; Ibrahim, M.; Gul, S.; Fransson, T.; Brewster, A. S.; Alonso-Mori, R. Structures of the Intermediates of Kok’s Photosynthetic Water Oxidation Clock. Nature 2018, 563 (7731), 421425,  DOI: 10.1038/s41586-018-0681-2
    157. 157
      Nango, E.; Royant, A.; Kubo, M.; Nakane, T.; Wickstrand, C.; Kimura, T.; Tanaka, T.; Tono, K.; Song, C.; Tanaka, R. A Three-Dimensional Movie of Structural Changes in Bacteriorhodopsin. Science 2016, 354 (6319), 15521557,  DOI: 10.1126/science.aah3497
    158. 158
      Liu, X.; Liu, P.; Li, H.; Xu, Z.; Jia, L.; Xia, Y.; Yu, M.; Tang, W.; Zhu, X.; Chen, C. Excited-State Intermediates in a Designer Protein Encoding a Phototrigger Caught by an X-Ray Free-Electron Laser. Nat. Chem. 2022, 14 (9), 10541060,  DOI: 10.1038/s41557-022-00992-3
    159. 159
      Hosaka, T.; Katsura, K.; Ishizuka-Katsura, Y.; Hanada, K.; Ito, K.; Tomabechi, Y.; Inoue, M.; Akasaka, R.; Takemoto, C.; Shirouzu, M. Crystal Structure of an Archaeal Tyrosyl-tRNA Synthetase Bound to Photocaged L-Tyrosine and Its Potential Application to Time-Resolved X-ray Crystallography. Int. J. Mol. Sci. 2022, 23 (18), 10399,  DOI: 10.3390/ijms231810399
    160. 160
      Markley, J. L.; Putter, I.; Jardetzky, O. High-Resolution Nuclear Magnetic Resonance Spectra of Selectively Deuterated Staphylococcal Nuclease. Science 1968, 161 (3847), 12491251,  DOI: 10.1126/science.161.3847.1249
    161. 161
      Deiters, A.; Geierstanger, B. H.; Schultz, P. G. Site-Specific in vivo Labeling of Proteins for NMR Studies. ChemBioChem 2005, 6 (1), 5558,  DOI: 10.1002/cbic.200400319
    162. 162
      Jones, D. H.; Cellitti, S. E.; Hao, X.; Zhang, Q.; Jahnz, M.; Summerer, D.; Schultz, P. G.; Uno, T.; Geierstanger, B. H. Site-Specific Labeling of Proteins with NMR-Active Unnatural Amino Acids. J. Biomol. NMR 2010, 46 (1), 89100,  DOI: 10.1007/s10858-009-9365-4
    163. 163
      Abdelkader, E. H.; Qianzhu, H.; Huber, T.; Otting, G. Genetic Encoding of 7-Aza-l-tryptophan: Isoelectronic Substitution of a Single CH-Group in a Protein for a Nitrogen Atom for Site-Selective Isotope Labeling. ACS Sensors 2023, 8 (11), 44024406,  DOI: 10.1021/acssensors.3c01904
    164. 164
      Cellitti, S. E.; Jones, D. H.; Lagpacan, L.; Hao, X.; Zhang, Q.; Hu, H.; Brittain, S. M.; Brinker, A.; Caldwell, J.; Bursulaya, B. In vivo Incorporation of Unnatural Amino Acids to Probe Structure, Dynamics, and Ligand Binding in a Large Protein by Nuclear Magnetic Resonance Spectroscopy. J. Am. Chem. Soc. 2008, 130 (29), 92689281,  DOI: 10.1021/ja801602q
    165. 165
      Lampe, J. N.; Floor, S. N.; Gross, J. D.; Nishida, C. R.; Jiang, Y.; Trnka, M. J.; Ortiz de Montellano, P. R. Ligand-Induced Conformational Heterogeneity of Cytochrome P450 CYP119 Identified by 2D NMR Spectroscopy with the Unnatural Amino Acid (13)C-p-Methoxyphenylalanine. J. Am. Chem. Soc. 2008, 130 (48), 1616816169,  DOI: 10.1021/ja8071463
    166. 166
      Hull, W. E.; Sykes, B. D. Fluorotyrosine Alkaline Phosphatase. Fluorine-19 Nuclear Magnetic Resonance Relaxation Times and Molecular Motion of the Individual Fluorotyrosines. Biochemistry 1974, 13 (17), 34313437,  DOI: 10.1021/bi00714a002
    167. 167
      Gamcsik, M. P.; Gerig, J. T. NMR Studies of Fluorophenylalanine-Containing Carbonic Anhydrase. FEBS Lett. 1986, 196 (1), 7174,  DOI: 10.1016/0014-5793(86)80216-2
    168. 168
      Jackson, J. C.; Hammill, J. T.; Mehl, R. A. Site-Specific Incorporation of a 19F-Amino Acid into Proteins as an NMR Probe for Characterizing Protein Structure and Reactivity. J. Am. Chem. Soc. 2007, 129 (5), 11601166,  DOI: 10.1021/ja064661t
    169. 169
      Hammill, J. T.; Miyake-Stoner, S.; Hazen, J. L.; Jackson, J. C.; Mehl, R. A. Preparation of Site-Specifically Labeled Fluorinated Proteins for 19F-NMR Structural Characterization. Nat. Protoc. 2007, 2 (10), 26012607,  DOI: 10.1038/nprot.2007.379
    170. 170
      Wacks, D. B.; Schachman, H. K. 19F Nuclear Magnetic Resonance Studies of Fluorotyrosine-Labeled Aspartate Transcarbamoylase. Properties of the Enzyme and Its Catalytic and Regulatory Subunits. J. Biol. Chem. 1985, 260 (21), 1165111658,  DOI: 10.1016/S0021-9258(17)39080-4
    171. 171
      Gerig, J. T. Fluorine NMR of Proteins. Prog. Nucl. Magn. Reson. Spectrosc. 1994, 26, 293370,  DOI: 10.1016/0079-6565(94)80009-X
    172. 172
      Furter, R. Expansion of the Genetic Code: Site-Directed p-Fluoro-Phenylalanine Incorporation in Escherichia coli. Protein Sci. 1998, 7 (2), 419426,  DOI: 10.1002/pro.5560070223
    173. 173
      Kim, H.-W.; Perez, J. A.; Ferguson, S. J.; Campbell, I. D. The Specific Incorporation of Labelled Aromatic Amino Acids into Proteins through Growth of Bacteria in the Presence of Glyphosate. FEBS Lett. 1990, 272 (1–2), 3436,  DOI: 10.1016/0014-5793(90)80442-L
    174. 174
      Niu, W.; Shu, Q.; Chen, Z.; Mathews, S.; Di Cera, E.; Frieden, C. The Role of Zn2+ on the Structure and Stability of Murine Adenosine Deaminase. J. Phys. Chem. B 2010, 114 (49), 1615616165,  DOI: 10.1021/jp106041v
    175. 175
      Ruben, E. A.; Gandhi, P. S.; Chen, Z.; Koester, S. K.; DeKoster, G. T.; Frieden, C.; Di Cera, E. 19F NMR Reveals the Conformational Properties of Free Thrombin and Its Zymogen Precursor Prethrombin-2. J. Biol. Chem. 2020, 295 (24), 82278235,  DOI: 10.1074/jbc.RA120.013419
    176. 176
      Duewel, H.; Daub, E.; Robinson, V.; Honek, J. F. Incorporation of Trifluoromethionine into a Phage Lysozyme: Implications and a New Marker for Use in Protein 19F NMR. Biochemistry 1997, 36 (11), 34043416,  DOI: 10.1021/bi9617973
    177. 177
      Cleve, P.; Robinson, V.; Duewel, H. S.; Honek, J. F. Difluoromethionine as a Novel 19F NMR Structural Probe for Internal Amino Acid Packing in Proteins. J. Am. Chem. Soc. 1999, 121 (37), 84758478,  DOI: 10.1021/ja9911418
    178. 178
      Holzberger, B.; Rubini, M.; Möller, H. M.; Marx, A. A Highly Active DNA Polymerase with a Fluorous Core. Angew. Chem. Int. Ed. 2010, 49 (7), 13241327,  DOI: 10.1002/anie.200905978
    179. 179
      Orton, H. W.; Qianzhu, H.; Abdelkader, E. H.; Habel, E. I.; Tan, Y. J.; Frkic, R. L.; Jackson, C. J.; Huber, T.; Otting, G. Through-Space Scalar 19F-19F Couplings between Fluorinated Noncanonical Amino Acids for the Detection of Specific Contacts in Proteins. J. Am. Chem. Soc. 2021, 143 (46), 1958719598,  DOI: 10.1021/jacs.1c10104
    180. 180
      Miao, Q.; Nitsche, C.; Orton, H.; Overhand, M.; Otting, G.; Ubbink, M. Paramagnetic Chemical Probes for Studying Biological Macromolecules. Chem. Rev. 2022, 122 (10), 95719642,  DOI: 10.1021/acs.chemrev.1c00708
    181. 181
      Fanucci, G. E.; Cafiso, D. S. Recent advances and applications of site-directed spin labeling. Curr. Opin. Struct. Biol. 2006, 16 (5), 644653,  DOI: 10.1016/j.sbi.2006.08.008
    182. 182
      Braun, T.; Drescher, M.; Summerer, D. Expanding the Genetic Code for Site-Directed Spin-Labeling. Int. J. Mol. Sci. 2019, 20 (2), 373,  DOI: 10.3390/ijms20020373
    183. 183
      Kálai, T.; Fleissner, M. R.; Jeko, J.; Hubbell, W. L.; Hideg, K. Synthesis of New Spin Labels for Cu-Free Click Conjugation. Tetrahedron Lett. 2011, 52 (21), 27472749,  DOI: 10.1016/j.tetlet.2011.03.077
    184. 184
      Kugele, A.; Braun, T. S.; Widder, P.; Williams, L.; Schmidt, M. J.; Summerer, D.; Drescher, M. Site-Directed Spin Labelling of Proteins by Suzuki-Miyaura Coupling via a Genetically Encoded Aryliodide Amino Acid. Chem. Commun. 2019, 55 (13), 19231926,  DOI: 10.1039/C8CC09325C
    185. 185
      Jana, S.; Evans, E. G. B.; Jang, H. S.; Zhang, S.; Zhang, H.; Rajca, A.; Gordon, S. E.; Zagotta, W. N.; Stoll, S.; Mehl, R. A. Ultrafast Bioorthogonal Spin-Labeling and Distance Measurements in Mammalian Cells Using Small, Genetically Encoded Tetrazine Amino Acids. J. Am. Chem. Soc. 2023, 145 (27), 1460814620,  DOI: 10.1021/jacs.3c00967
    186. 186
      Fleissner, M. R.; Brustad, E. M.; Kálai, T.; Altenbach, C.; Cascio, D.; Peters, F. B.; Hideg, K.; Peuker, S.; Schultz, P. G.; Hubbell, W. L. Site-Directed Spin Labeling of a Genetically Encoded Unnatural Amino Acid. Proc. Natl. Acad. Sci. U.S.A. 2009, 106 (51), 2163721642,  DOI: 10.1073/pnas.0912009106
    187. 187
      Nguyen, D. P.; Lusic, H.; Neumann, H.; Kapadnis, P. B.; Deiters, A.; Chin, J. W. Genetic Encoding and Labeling of Aliphatic Azides and Alkynes in Recombinant Proteins via a Pyrrolysyl-tRNA Synthetase/tRNACUA Pair and Click Chemistry. J. Am. Chem. Soc. 2009, 131 (25), 87208721,  DOI: 10.1021/ja900553w
    188. 188
      Chin, J. W.; Santoro, S. W.; Martin, A. B.; King, D. S.; Wang, L.; Schultz, P. G. Addition of p-Azido-l-phenylalanine to the Genetic Code of Escherichia coli. J. Am. Chem. Soc. 2002, 124 (31), 90269027,  DOI: 10.1021/ja027007w
    189. 189
      Kucher, S.; Korneev, S.; Tyagi, S.; Apfelbaum, R.; Grohmann, D.; Lemke, E. A.; Klare, J. P.; Steinhoff, H.-J.; Klose, D. Orthogonal Spin Labeling Using Click Chemistry for in vitro and in vivo Applications. J. Magn. Reson. 2017, 275, 3845,  DOI: 10.1016/j.jmr.2016.12.001
    190. 190
      Abdelkader, E. H.; Feintuch, A.; Yao, X.; Adams, L. A.; Aurelio, L.; Graham, B.; Goldfarb, D.; Otting, G. Protein Conformation by EPR Spectroscopy Using Gadolinium Tags Clicked to Genetically Encoded p-Azido-L-Phenylalanine. Chem. Commun. 2015, 51 (88), 1589815901,  DOI: 10.1039/C5CC07121F
    191. 191
      Loh, C. T.; Ozawa, K.; Tuck, K. L.; Barlow, N.; Huber, T.; Otting, G.; Graham, B. Lanthanide Tags for Site-Specific Ligation to an Unnatural Amino Acid and Generation of Pseudocontact Shifts in Proteins. Bioconjug. Chem. 2013, 24 (2), 260268,  DOI: 10.1021/bc300631z
    192. 192
      Mahawaththa, M. C.; Lee, M. D.; Giannoulis, A.; Adams, L. A.; Feintuch, A.; Swarbrick, J. D.; Graham, B.; Nitsche, C.; Goldfarb, D.; Otting, G. Small Neutral Gd(iii) Tags for Distance Measurements in Proteins by Double Electron-Electron Resonance Experiments. Phys. Chem. Chem. Phys. 2018, 20 (36), 2353523545,  DOI: 10.1039/C8CP03532F
    193. 193
      Yang, H.; Yang, S.; Kong, J.; Dong, A.; Yu, S. Obtaining Information About Protein Secondary Structures in Aqueous Solution Using Fourier Transform IR Spectroscopy. Nat. Protoc. 2015, 10 (3), 382396,  DOI: 10.1038/nprot.2015.024
    194. 194
      Lorenz-Fonfria, V. A. Infrared Difference Spectroscopy of Proteins: From Bands to Bonds. Chem. Rev. 2020, 120 (7), 34663576,  DOI: 10.1021/acs.chemrev.9b00449
    195. 195
      Chung, J. K.; Thielges, M. C.; Fayer, M. D. Dynamics of the Folded and Unfolded Villin Headpiece (HP35) Measured with Ultrafast 2D IR Vibrational Echo Spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 2011, 108 (9), 35783583,  DOI: 10.1073/pnas.1100587108
    196. 196
      Chung, J. K.; Thielges, M. C.; Fayer, M. D. Conformational Dynamics and Stability of HP35 Studied with 2D IR Vibrational Echoes. J. Am. Chem. Soc. 2012, 134 (29), 1211812124,  DOI: 10.1021/ja303017d
    197. 197
      Urbanek, D. C.; Vorobyev, D. Y.; Serrano, A. L.; Gai, F.; Hochstrasser, R. M. The Two-Dimensional Vibrational Echo of a Nitrile Probe of the Villin HP35 Protein. J. Phys. Chem. Lett. 2010, 1 (23), 33113315,  DOI: 10.1021/jz101367d
    198. 198
      Bagchi, S.; Boxer, S. G.; Fayer, M. D. Ribonuclease S Dynamics Measured Using a Nitrile Label with 2D IR Vibrational Echo Spectroscopy. J. Phys. Chem. B 2012, 116 (13), 40344042,  DOI: 10.1021/jp2122856
    199. 199
      Tharp, J. M.; Wang, Y.-S.; Lee, Y.-J.; Yang, Y.; Liu, W. R. Genetic Incorporation of Seven ortho-Substituted Phenylalanine Derivatives. ACS Chem. Biol. 2014, 9 (4), 884890,  DOI: 10.1021/cb400917a
    200. 200
      van Wilderen, L. J. G. W.; Kern-Michler, D.; Müller-Werkmeister, H. M.; Bredenbeck, J. Vibrational Dynamics and Solvatochromism of the Label SCN in Various Solvents and Hemoglobin by Time Dependent IR and 2D-IR Spectroscopy. Phys. Chem. Chem. Phys. 2014, 16 (36), 1964319653,  DOI: 10.1039/C4CP01498G
    201. 201
      Bloem, R.; Koziol, K.; Waldauer, S. A.; Buchli, B.; Walser, R.; Samatanga, B.; Jelesarov, I.; Hamm, P. Ligand Binding Studied by 2D IR Spectroscopy Using the Azidohomoalanine Label. J. Phys. Chem. B 2012, 116 (46), 1370513712,  DOI: 10.1021/jp3095209
    202. 202
      Thielges, M. C.; Axup, J. Y.; Wong, D.; Lee, H. S.; Chung, J. K.; Schultz, P. G.; Fayer, M. D. Two-Dimensional IR Spectroscopy of Protein Dynamics Using Two Vibrational Labels: A Site-Specific Genetically Encoded Unnatural Amino Acid and an Active Site Ligand. J. Phys. Chem. B 2011, 115 (38), 1129411304,  DOI: 10.1021/jp206986v
    203. 203
      Ye, S.; Zaitseva, E.; Caltabiano, G.; Schertler, G. F. X.; Sakmar, T. P.; Deupi, X.; Vogel, R. Tracking G-Protein-Coupled Receptor Activation Using Genetically Encoded Infrared Probes. Nature 2010, 464 (7293), 13861389,  DOI: 10.1038/nature08948
    204. 204
      Le Sueur, A. L.; Horness, R. E.; Thielges, M. C. Applications of Two-Dimensional Infrared Spectroscopy. Analyst 2015, 140 (13), 43364349,  DOI: 10.1039/C5AN00558B
    205. 205
      Smith, E. E.; Linderman, B. Y.; Luskin, A. C.; Brewer, S. H. Probing Local Environments with the Infrared Probe: l-4-Nitrophenylalanine. J. Phys. Chem. B 2011, 115 (10), 23802385,  DOI: 10.1021/jp109288j
    206. 206
      Kooter, I. M.; Moguilevsky, N.; Bollen, A.; van der Veen, L. A.; Otto, C.; Dekker, H. L.; Wever, R. The Sulfonium Ion Linkage in Myeloperoxidase: direct spectroscopic detection by isotopic labeling and effect of mutation. J. Biol. Chem. 1999, 274 (38), 2679426802,  DOI: 10.1074/jbc.274.38.26794
    207. 207
      Thielges, M. C.; Case, D. A.; Romesberg, F. E. Carbon-Deuterium Bonds as Probes of Dihydrofolate Reductase. J. Am. Chem. Soc. 2008, 130 (20), 65976603,  DOI: 10.1021/ja0779607
    208. 208
      Le Sueur, A. L.; Schaugaard, R. N.; Baik, M.-H.; Thielges, M. C. Methionine Ligand Interaction in a Blue Copper Protein Characterized by Site-Selective Infrared Spectroscopy. J. Am. Chem. Soc. 2016, 138 (22), 71877193,  DOI: 10.1021/jacs.6b03916
    209. 209
      Horness, R. E.; Basom, E. J.; Thielges, M. C. Site-Selective Characterization of Src Homology 3 Domain Molecular Recognition with Cyanophenylalanine Infrared Probes. Anal. Methods 2015, 7 (17), 72347241,  DOI: 10.1039/C5AY00523J
    210. 210
      Basom, E. J.; Maj, M.; Cho, M.; Thielges, M. C. Site-Specific Characterization of Cytochrome P450cam Conformations by Infrared Spectroscopy. Anal. Chem. 2016, 88 (12), 65986606,  DOI: 10.1021/acs.analchem.6b01520
    211. 211
      Bagchi, S.; Fried, S. D.; Boxer, S. G. A Solvatochromic Model Calibrates Nitriles’ Vibrational Frequencies to Electrostatic Fields. J. Am. Chem. Soc. 2012, 134 (25), 1037310376,  DOI: 10.1021/ja303895k
    212. 212
      Wang, L.; Zhang, J.; Han, M.-J.; Zhang, L.; Chen, C.; Huang, A.; Xie, R.; Wang, G.; Zhu, J.; Wang, Y. A Genetically Encoded Two-Dimensional Infrared Probe for Enzyme Active-Site Dynamics. Angew. Chem. Int. Ed. 2021, 60 (20), 1114311147,  DOI: 10.1002/anie.202016880
    213. 213
      Huguenin-Dezot, N.; Alonzo, D. A.; Heberlig, G. W.; Mahesh, M.; Nguyen, D. P.; Dornan, M. H.; Boddy, C. N.; Schmeing, T. M.; Chin, J. W. Trapping Biosynthetic Acyl-Enzyme Intermediates with Encoded 2,3-Diaminopropionic Acid. Nature 2019, 565 (7737), 112117,  DOI: 10.1038/s41586-018-0781-z
    214. 214
      Tang, S.; Beattie, A. T.; Kafkova, L.; Petris, G.; Huguenin-Dezot, N.; Fiedler, M.; Freeman, M.; Chin, J. W. Mechanism-Based Traps Enable Protease and Hydrolase Substrate Discovery. Nature 2022, 602 (7898), 701707,  DOI: 10.1038/s41586-022-04414-9
    215. 215
      Dorman, G.; Prestwich, G. D. Benzophenone Photophores in Biochemistry. Biochemistry 1994, 33 (19), 56615673,  DOI: 10.1021/bi00185a001
    216. 216
      Chin, J. W.; Martin, A. B.; King, D. S.; Wang, L.; Schultz, P. G. Addition of a Photocrosslinking Amino Acid to the Genetic Code of Escherichia coli. Proc. Natl. Acad. Sci. U.S.A. 2002, 99 (17), 1102011024,  DOI: 10.1073/pnas.172226299
    217. 217
      Farrell, I. S.; Toroney, R.; Hazen, J. L.; Mehl, R. A.; Chin, J. W. Photo-Cross-Linking Interacting Proteins with a Genetically Encoded Benzophenone. Nat. Methods 2005, 2 (5), 377384,  DOI: 10.1038/nmeth0505-377
    218. 218
      Chin, J. W.; Schultz, P. G. In Vivo Photocrosslinking with Unnatural Amino Acid Mutagenesis. ChemBioChem 2002, 3 (11), 11351137,  DOI: 10.1002/1439-7633(20021104)3:11<1135::AID-CBIC1135>3.0.CO;2-M
    219. 219
      Mori, H.; Ito, K. Different Modes of SecY-SecA Interactions Revealed by Site-Directed in vivo Photo-Cross-Linking. Proc. Natl. Acad. Sci. U.S.A. 2006, 103 (44), 1615916164,  DOI: 10.1073/pnas.0606390103
    220. 220
      Majmudar, C. Y.; Lee, L. W.; Lancia, J. K.; Nwokoye, A.; Wang, Q.; Wands, A. M.; Wang, L.; Mapp, A. K. Impact of Nonnatural Amino Acid Mutagenesis on the in Vivo Function and Binding Modes of a Transcriptional Activator. J. Am. Chem. Soc. 2009, 131 (40), 1424014242,  DOI: 10.1021/ja904378z
    221. 221
      Liu, C.; Young, A. L.; Starling-Windhof, A.; Bracher, A.; Saschenbrecker, S.; Rao, B. V.; Rao, K. V.; Berninghausen, O.; Mielke, T.; Hartl, F. U. Coupled Chaperone Action in Folding and Assembly of Hexadecameric Rubisco. Nature 2010, 463 (7278), 197202,  DOI: 10.1038/nature08651
    222. 222
      Ye, Z.; Bair, M.; Desai, H.; Williams, G. J. A Photocrosslinking Assay for Reporting Protein Interactions in Polyketide and Fatty Acid Synthases. Mol. BioSyst. 2011, 7 (11), 31523156,  DOI: 10.1039/c1mb05270e
    223. 223
      Ye, Z.; Williams, G. J. Mapping a Ketosynthase:Acyl Carrier Protein Binding Interface via Unnatural Amino Acid-Mediated Photo-Cross-Linking. Biochemistry 2014, 53 (48), 74947502,  DOI: 10.1021/bi500936u
    224. 224
      Pavic, K.; Rios, P.; Dzeyk, K.; Koehler, C.; Lemke, E. A.; Köhn, M. Unnatural Amino Acid Mutagenesis Reveals Dimerization As a Negative Regulatory Mechanism of VHR’s Phosphatase Activity. ACS Chem. Biol. 2014, 9 (7), 14511459,  DOI: 10.1021/cb500240n
    225. 225
      Morozov, Y. I.; Agaronyan, K.; Cheung, A. C. M.; Anikin, M.; Cramer, P.; Temiakov, D. A Novel Intermediate in Transcription Initiation by Human Mitochondrial RNA Polymerase. Nucleic Acids Res. 2014, 42 (6), 38843893,  DOI: 10.1093/nar/gkt1356
    226. 226
      Wong, H. E.; Kwon, I. Effects of Non-Natural Amino Acid Incorporation into the Enzyme Core Region on Enzyme Structure and Function. Int. J. Mol. Sci. 2015, 16 (9), 2273522753,  DOI: 10.3390/ijms160922735
    227. 227
      Rappaport, F.; Boussac, A.; Force, D. A.; Peloquin, J.; Brynda, M.; Sugiura, M.; Un, S.; Britt, R. D.; Diner, B. A. Probing the Coupling between Proton and Electron Transfer in Photosystem II Core Complexes Containing a 3-Fluorotyrosine. J. Am. Chem. Soc. 2009, 131 (12), 44254433,  DOI: 10.1021/ja808604h
    228. 228
      Schlesinger, S.; Schlesinger, M. J. The Effect of Amino Acid Analogues on Alkaline Phosphatase Formation in Escherichia coli K-12: I. Substitution of Triazolealanine for Histidine. J. Biol. Chem. 1967, 242 (14), 33693372,  DOI: 10.1016/S0021-9258(18)95919-3
    229. 229
      Fried, S. D.; Bagchi, S.; Boxer, S. G. Extreme Electric Fields Power Catalysis in the Active Site of Ketosteroid Isomerase. Science 2014, 346 (6216), 15101514,  DOI: 10.1126/science.1259802
    230. 230
      Ortmayer, M.; Hardy, F. J.; Quesne, M. G.; Fisher, K.; Levy, C.; Heyes, D. J.; Catlow, C. R. A.; de Visser, S. P.; Rigby, S. E. J.; Hay, S.; Green, A. P. A Noncanonical Tryptophan Analogue Reveals an Active Site Hydrogen Bond Controlling Ferryl Reactivity in a Heme Peroxidase. JACS Au 2021, 1 (7), 913918,  DOI: 10.1021/jacsau.1c00145
    231. 231
      Kang, G.; Taguchi, A. T.; Stubbe, J.; Drennan, C. L. Structure of a Trapped Radical Transfer Pathway within a Ribonucleotide Reductase Holocomplex. Science 2020, 368 (6489), 424427,  DOI: 10.1126/science.aba6794
    232. 232
      Beiboer, S. H. W.; Berg, B. v. d.; Dekker, N.; Cox, R. C.; Verheij, H. M. Incorporation of an Unnatural Amino Acid in the Active Site of Porcine Pancreatic Phospholipase A2. Substitution of Histidine by l,2,4-Triazole-3-Alanine Yields an Enzyme with High Activity at Acidic pH. Protein Eng. Des. Sel. 1996, 9 (4), 345352,  DOI: 10.1093/protein/9.4.345
    233. 233
      Soumillion, P.; Fastrez, J. Incorporation of 1,2,4-Triazole-3-Alanine into a Mutant of Phage Lambda Lysozyme Containing a Single Histidine. Protein Eng. Des. Sel. 1998, 11 (3), 213217,  DOI: 10.1093/protein/11.3.213
    234. 234
      Blatter, N.; Prokup, A.; Deiters, A.; Marx, A. Modulating the pKa of a Tyrosine in KlenTaq DNA Polymerase that Is Crucial for Abasic Site Bypass by in Vivo Incorporation of a Non-canonical Amino Acid. Chem Bio Chem 2014, 15, 17351737,  DOI: 10.1002/cbic.201400051
    235. 235
      Obeid, S.; Blatter, N.; Kranaster, R.; Schnur, A.; Diederichs, K.; Welte, W.; Marx, A. Replication Through an Abasic DNA Lesion: Structural Basis for Adenine Selectivity. EMBO J. 2010, 29 (10), 1738,  DOI: 10.1038/emboj.2010.64
    236. 236
      Greene, B. L.; Kang, G.; Cui, C.; Bennati, M.; Nocera, D. G.; Drennan, C. L.; Stubbe, J. Ribonucleotide Reductases: Structure, Chemistry, and Metabolism Suggest New Therapeutic Targets. Annu. Rev. Biochem. 2020, 89 (1), 4575,  DOI: 10.1146/annurev-biochem-013118-111843
    237. 237
      Uhlin, U.; Eklund, H. Structure of Ribonucleotide Reductase Protein R1. Nature 1994, 370 (6490), 533539,  DOI: 10.1038/370533a0
    238. 238
      Reece, S. Y.; Seyedsayamdost, M. R.; Stubbe, J.; Nocera, D. G. Electron Transfer Reactions of Fluorotyrosyl Radicals. J. Am. Chem. Soc. 2006, 128 (42), 1365413655,  DOI: 10.1021/ja0636688
    239. 239
      Seyedsayamdost, M. R.; Yee, C. S.; Reece, S. Y.; Nocera, D. G.; Stubbe, J. pH Rate Profiles of FnY356-R2s (n = 2, 3, 4) in Escherichia coli Ribonucleotide Reductase: Evidence that Y356 Is a Redox-Active Amino Acid along the Radical Propagation Pathway. J. Am. Chem. Soc. 2006, 128 (5), 15621568,  DOI: 10.1021/ja055927j
    240. 240
      Seyedsayamdost, M. R.; Xie, J.; Chan, C. T. Y.; Schultz, P. G.; Stubbe, J. Site-Specific Insertion of 3-Aminotyrosine into Subunit α2 of E. coli Ribonucleotide Reductase: Direct Evidence for Involvement of Y730 and Y731 in Radical Propagation. J. Am. Chem. Soc. 2007, 129 (48), 1506015071,  DOI: 10.1021/ja076043y
    241. 241
      Seyedsayamdost, M. R.; Chan, C. T. Y.; Mugnaini, V.; Stubbe, J.; Bennati, M. PELDOR Spectroscopy with DOPA-β2 and NH2Y-α2s: Distance Measurements between Residues Involved in the Radical Propagation Pathway of E. coli Ribonucleotide Reductase. J. Am. Chem. Soc. 2007, 129 (51), 1574815749,  DOI: 10.1021/ja076459b
    242. 242
      Seyedsayamdost, M. R.; Stubbe, J. Site-Specific Replacement of Y356 with 3,4-Dihydroxyphenylalanine in the β2 Subunit of E. coli Ribonucleotide Reductase. J. Am. Chem. Soc. 2006, 128 (8), 25222523,  DOI: 10.1021/ja057776q
    243. 243
      Seyedsayamdost, M. R.; Stubbe, J. Forward and Reverse Electron Transfer with the Y356DOPA-β2 Heterodimer of E. coli Ribonucleotide Reductase. J. Am. Chem. Soc. 2007, 129 (8), 22262227,  DOI: 10.1021/ja0685607
    244. 244
      Yokoyama, K.; Smith, A. A.; Corzilius, B.; Griffin, R. G.; Stubbe, J. Equilibration of Tyrosyl Radicals (Y356•, Y731•, Y730•) in the Radical Propagation Pathway of the Escherichia coli Class Ia Ribonucleotide Reductase. J. Am. Chem. Soc. 2011, 133 (45), 1842018432,  DOI: 10.1021/ja207455k
    245. 245
      Lin, C.-Y.; Muñoz Hernández, A. L.; Laremore, T. N.; Silakov, A.; Krebs, C.; Boal, A. K.; Bollinger, J. M., Jr Use of Noncanonical Tyrosine Analogues to Probe Control of Radical Intermediates during Endoperoxide Installation by Verruculogen Synthase (FtmOx1). ACS Catal. 2022, 12 (12), 69686979,  DOI: 10.1021/acscatal.2c01037
    246. 246
      Taylor, A.; Heyes, D. J.; Scrutton, N. S. Catalysis by Nature’s photoenzymes. Curr. Opin. Struct. Biol. 2022, 77, 102491,  DOI: 10.1016/j.sbi.2022.102491
    247. 247
      Taylor, A.; Zhang, S.; Johannissen, L. O.; Sakuma, M.; Phillips, R. S.; Green, A. P.; Hay, S.; Heyes, D. J.; Scrutton, N. S. Mechanistic Implications of the Ternary Complex Structural Models for the Photoenzyme Protochlorophyllide Oxidoreductase. FEBS J. 2024, 291, 1404,  DOI: 10.1111/febs.17025
    248. 248
      Horowitz, S.; Adhikari, U.; Dirk, L. M. A.; Del Rizzo, P. A.; Mehl, R. A.; Houtz, R. L.; Al-Hashimi, H. M.; Scheiner, S.; Trievel, R. C. Manipulating Unconventional CH-Based Hydrogen Bonding in a Methyltransferase via Noncanonical Amino Acid Mutagenesis. ACS Chem. Biol. 2014, 9 (8), 16921697,  DOI: 10.1021/cb5001185
    249. 249
      Kirsh, J. M.; Weaver, J. B.; Boxer, S. G.; Kozuch, J. Comprehensive Analysis of Nitrile Probe IR Shifts and Intensities in Proteins: Experiment and Critical Evaluation of Simulations. ChemRxiv 2024,  DOI: 10.26434/chemrxiv-2023-j935v-v2
    250. 250
      Weaver, J. B.; Kozuch, J.; Kirsh, J. M.; Boxer, S. G. Nitrile Infrared Intensities Characterize Electric Fields and Hydrogen Bonding in Protic, Aprotic, and Protein Environments. J. Am. Chem. Soc. 2022, 144 (17), 75627567,  DOI: 10.1021/jacs.2c00675
    251. 251
      Zheng, C.; Mao, Y.; Kozuch, J.; Atsango, A. O.; Ji, Z.; Markland, T. E.; Boxer, S. G. A Two-Directional Vibrational Probe Reveals Different Electric Field Orientations in Solution and an Enzyme Active Site. Nat. Chem. 2022, 14 (8), 891897,  DOI: 10.1038/s41557-022-00937-w
    252. 252
      Kedzierski, P.; Zaczkowska, M.; Sokalski, W. A. Extreme Catalytic Power of Ketosteroid Isomerase Related to the Reversal of Proton Dislocations in Hydrogen-Bond Network. J. Phys. Chem. B 2020, 124 (18), 36613666,  DOI: 10.1021/acs.jpcb.0c01489
    253. 253
      Pollack, R. M. Enzymatic Mechanisms for Catalysis of Enolization: Ketosteroid Isomerase. Bioorg. Chem. 2004, 32 (5), 341353,  DOI: 10.1016/j.bioorg.2004.06.005
    254. 254
      Wu, Y.; Boxer, S. G. A Critical Test of the Electrostatic Contribution to Catalysis with Noncanonical Amino Acids in Ketosteroid Isomerase. J. Am. Chem. Soc. 2016, 138 (36), 1189011895,  DOI: 10.1021/jacs.6b06843
    255. 255
      Faraldos, J. A.; Antonczak, A. K.; González, V.; Fullerton, R.; Tippmann, E. M.; Allemann, R. K. Probing Eudesmane Cation-π Interactions in Catalysis by Aristolochene Synthase with Non-canonical Amino Acids. J. Am. Chem. Soc. 2011, 133 (35), 1390613909,  DOI: 10.1021/ja205927u
    256. 256
      Morikubo, N.; Fukuda, Y.; Ohtake, K.; Shinya, N.; Kiga, D.; Sakamoto, K.; Asanuma, M.; Hirota, H.; Yokoyama, S.; Hoshino, T. Cation-π Interaction in the Polyolefin Cyclization Cascade Uncovered by Incorporating Unnatural Amino Acids into the Catalytic Sites of Squalene Cyclase. J. Am. Chem. Soc. 2006, 128 (40), 1318413194,  DOI: 10.1021/ja063358p
    257. 257
      Herbst, R. W.; Guce, A.; Bryngelson, P. A.; Higgins, K. A.; Ryan, K. C.; Cabelli, D. E.; Garman, S. C.; Maroney, M. J. Role of Conserved Tyrosine Residues in NiSOD Catalysis: A Case of Convergent Evolution. Biochemistry 2009, 48 (15), 33543369,  DOI: 10.1021/bi802029t
    258. 258
      Campeciño, J. O.; Dudycz, L. W.; Tumelty, D.; Berg, V.; Cabelli, D. E.; Maroney, M. J. A Semisynthetic Strategy Leads to Alteration of the Backbone Amidate Ligand in the NiSOD Active Site. J. Am. Chem. Soc. 2015, 137 (28), 90449052,  DOI: 10.1021/jacs.5b03629
    259. 259
      Evans, R. M.; Krahn, N.; Murphy, B. J.; Lee, H.; Armstrong, F. A.; Söll, D. Selective Cysteine-To-Selenocysteine Changes in a [NiFe]-Hydrogenase Confirm a Special Position for Catalysis and Oxygen Tolerance. Proc. Natl. Acad. Sci. U.S.A. 2021, 118 (13), e2100921118  DOI: 10.1073/pnas.2100921118
    260. 260
      Mukai, T.; Sevostyanova, A.; Suzuki, T.; Fu, X.; Söll, D. A Facile Method for Producing Selenocysteine-Containing Proteins. Angew. Chem. Int. Ed. 2018, 57 (24), 72157219,  DOI: 10.1002/anie.201713215
    261. 261
      Snider, G. W.; Ruggles, E.; Khan, N.; Hondal, R. J. Selenocysteine Confers Resistance to Inactivation by Oxidation in Thioredoxin Reductase: Comparison of Selenium and Sulfur Enzymes. Biochemistry 2013, 52 (32), 54725481,  DOI: 10.1021/bi400462j
    262. 262
      Wu, Z. P.; Hilvert, D. Selenosubtilisin as a Glutathione Peroxidase Mimic. J. Am. Chem. Soc. 1990, 112 (14), 56475648,  DOI: 10.1021/ja00170a043
    263. 263
      Hardy, F. J.; Ortmayer, M.; Green, A. P.; Noble, C. E. M.; Anderson, J. L. R. Recent Advances in Understanding, Enhancing and Creating Heme Peroxidases. In Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier, 2020. DOI: 10.1016/B978-0-08-102688-5.00021-0
    264. 264
      Poulos, T. L. Heme Enzyme Structure and Function. Chem. Rev. 2014, 114, 39193962,  DOI: 10.1021/cr400415k
    265. 265
      Behan, R. K.; Green, M. T. On the Status of Ferryl Protonation. J. Inorg. Biochem. 2006, 100 (4), 448459,  DOI: 10.1016/j.jinorgbio.2005.12.019
    266. 266
      Sivaramakrishnan, S.; Ouellet, H.; Du, J.; McLean, K. J.; Medzihradszky, K. F.; Dawson, J. H.; Munro, A. W.; Ortiz de Montellano, P. R. A Novel Intermediate in the Reaction of Seleno CYP119 with m-Chloroperbenzoic Acid. Biochemistry 2011, 50 (14), 30143024,  DOI: 10.1021/bi101728y
    267. 267
      Aldag, C.; Gromov, I. A.; García-Rubio, I.; Von Koenig, K.; Schlichting, I.; Jaun, B.; Hilvert, D. Probing the Role of the Proximal Heme Ligand in Cytochrome P450cam by Recombinant Incorporation of Selenocysteine. Proc. Natl. Acad. Sci. U.S.A. 2009, 106 (14), 54815486,  DOI: 10.1073/pnas.0810503106
    268. 268
      Jiang, Y.; Sivaramakrishnan, S.; Hayashi, T.; Cohen, S.; Moënne-Loccoz, P.; Shaik, S.; Ortiz de Montellano, P. R. Calculated and Experimental Spin State of Seleno Cytochrome P450. Angew. Chem. Int. Ed. 2009, 48 (39), 71937195,  DOI: 10.1002/anie.200901485
    269. 269
      Onderko, E. L.; Silakov, A.; Yosca, T. H.; Green, M. T. Characterization of a Selenocysteine-Ligated P450 Compound I Reveals Direct Link Between Electron Donation and Reactivity. Nat. Chem. 2017, 9 (7), 623628,  DOI: 10.1038/nchem.2781
    270. 270
      Sivaramakrishnan, S.; Ouellet, H.; Matsumura, H.; Guan, S.; Moënne-Loccoz, P.; Burlingame, A. L.; Ortiz De Montellano, P. R. Proximal Ligand Electron Donation and Reactivity of the Cytochrome P450 Ferric-Peroxo Anion. J. Am. Chem. Soc. 2012, 134 (15), 66736684,  DOI: 10.1021/ja211499q
    271. 271
      Ortmayer, M.; Fisher, K.; Basran, J.; Wolde-Michael, E. M.; Heyes, D. J.; Levy, C.; Lovelock, S. L.; Anderson, J. L. R.; Raven, E. L.; Hay, S. Rewiring the “ Push-Pull ” Catalytic Machinery of a Heme Enzyme Using an Expanded Genetic Code. ACS Catal. 2020, 10, 27352746,  DOI: 10.1021/acscatal.9b05129
    272. 272
      Xiao, H.; Peters, F. B.; Yang, P.-Y.; Reed, S.; Chittuluru, J. R.; Schultz, P. G. Genetic Incorporation of Histidine Derivatives Using an Engineered Pyrrolysyl-tRNA Synthetase. ACS Chem. Biol. 2014, 9 (5), 10921096,  DOI: 10.1021/cb500032c
    273. 273
      Martin, C.; Zhang, Y. The Diverse Functions of Histone Lysine Methylation. Nat. Rev. Mol. Cell Biol. 2005, 6 (11), 838849,  DOI: 10.1038/nrm1761
    274. 274
      Niu, W.; Guo, J. Co-translational Installation of Posttranslational Modifications by Non-canonical Amino Acid Mutagenesis. ChemBioChem 2023, 24 (9), e202300039  DOI: 10.1002/cbic.202300039
    275. 275
      Rust, H. L.; Subramanian, V.; West, G. M.; Young, D. D.; Schultz, P. G.; Thompson, P. R. Using Unnatural Amino Acid Mutagenesis To Probe the Regulation of PRMT1. ACS Chem. Biol. 2014, 9 (3), 649655,  DOI: 10.1021/cb400859z
    276. 276
      Neumann, H.; Neumann-Staubitz, P.; Witte, A.; Summerer, D. Epigenetic Chromatin Modification by Amber Suppression Technology. Curr. Opin. Chem. Biol. 2018, 45, 19,  DOI: 10.1016/j.cbpa.2018.01.017
    277. 277
      Wang, Z. A.; Cole, P. A. The Chemical Biology of Reversible Lysine Post-translational Modifications. Cell Chem. Biol. 2020, 27 (8), 953969,  DOI: 10.1016/j.chembiol.2020.07.002
    278. 278
      Wang, T.; Zhou, Q.; Li, F.; Yu, Y.; Yin, X.; Wang, J. Genetic Incorporation of Nε-Formyllysine, a New Histone Post-translational Modification. ChemBioChem 2015, 16 (10), 14401442,  DOI: 10.1002/cbic.201500170
    279. 279
      Cao, L.; Liu, J.; Ghelichkhani, F.; Rozovsky, S.; Wang, L. Genetic Incorporation of ε-N-Benzoyllysine by Engineering Methanomethylophilus alvus Pyrrolysyl-tRNA Synthetase. ChemBioChem 2021, 22 (15), 25302534,  DOI: 10.1002/cbic.202100218
    280. 280
      Ren, C.; Wu, Q.; Xiao, R.; Ji, Y.; Yang, X.; Zhang, Z.; Qin, H.; Ma, J.-A.; Xuan, W. Expanding the Scope of Genetically Encoded Lysine Post-Translational Modifications with Lactylation, β-Hydroxybutyrylation and Lipoylation. ChemBioChem 2022, 23 (18), e202200302  DOI: 10.1002/cbic.202200302
    281. 281
      Nguyen, D. P.; Garcia Alai, M. M.; Kapadnis, P. B.; Neumann, H.; Chin, J. W. Genetically Encoding Nϵ-Methyl-l-lysine in Recombinant Histones. J. Am. Chem. Soc. 2009, 131 (40), 1419414195,  DOI: 10.1021/ja906603s
    282. 282
      Wang, Z. A.; Liu, W. R. Proteins with Site-Specific Lysine Methylation. Chem. Eur. J. 2017, 23 (49), 1173211737,  DOI: 10.1002/chem.201701655
    283. 283
      Wang, Y.-S.; Wu, B.; Wang, Z.; Huang, Y.; Wan, W.; Russell, W. K.; Pai, P.-J.; Moe, Y. N.; Russell, D. H.; Liu, W. R. A Genetically Encoded Photocaged Nε-Methyl-L-Lysine. Mol. BioSyst. 2010, 6 (9), 15571560,  DOI: 10.1039/c002155e
    284. 284
      Neumann, H.; Peak-Chew, S. Y.; Chin, J. W. Genetically Encoding Nε-Acetyllysine in Recombinant Proteins. Nat. Chem. Biol. 2008, 4 (4), 232234,  DOI: 10.1038/nchembio.73
    285. 285
      Gattner, M. J.; Vrabel, M.; Carell, T. Synthesis of ε-N-Propionyl-, ε-N-Butyryl-, and ε-N-Crotonyl-Lysine Containing Histone H3 Using the Pyrrolysine System. Chem. Commun. 2013, 49 (4), 379381,  DOI: 10.1039/C2CC37836A
    286. 286
      Wilkins, B. J.; Hahn, L. E.; Heitmüller, S.; Frauendorf, H.; Valerius, O.; Braus, G. H.; Neumann, H. Genetically Encoding Lysine Modifications on Histone H4. ACS Chem. Biol. 2015, 10 (4), 939944,  DOI: 10.1021/cb501011v
    287. 287
      Xiao, H.; Xuan, W.; Shao, S.; Liu, T.; Schultz, P. G. Genetic Incorporation of ε-N-2-Hydroxyisobutyryl-Lysine into Recombinant Histones. ACS Chem. Biol. 2015, 10 (7), 15991603,  DOI: 10.1021/cb501055h
    288. 288
      Kim, C. H.; Kang, M.; Kim, H. J.; Chatterjee, A.; Schultz, P. G. Site-Specific Incorporation of ε-N-Crotonyllysine into Histones. Angew. Chem. Int. Ed. 2012, 51 (29), 72467249,  DOI: 10.1002/anie.201203349
    289. 289
      Tian, H.; Yang, J.; Guo, A.-D.; Ran, Y.; Yang, Y.-Z.; Yang, B.; Huang, R.; Liu, H.; Chen, X.-H. Genetically Encoded Benzoyllysines Serve as Versatile Probes for Interrogating Histone Benzoylation and Interactions in Living Cells. ACS Chem. Biol. 2021, 16 (11), 25602569,  DOI: 10.1021/acschembio.1c00614
    290. 290
      Fatema, N.; Fan, C. Studying Lysine Acetylation of Citric Acid Cycle Enzymes by Genetic Code Expansion. Mol. Microbiol. 2023, 119 (5), 551559,  DOI: 10.1111/mmi.15052
    291. 291
      Araujo, J.; Ottinger, S.; Venkat, S.; Gan, Q.; Fan, C. Studying Acetylation of Aconitase Isozymes by Genetic Code Expansion. Front. Chem. 2022, 10, 1,  DOI: 10.3389/fchem.2022.862483
    292. 292
      Venkat, S.; Chen, H.; Stahman, A.; Hudson, D.; McGuire, P.; Gan, Q.; Fan, C. Characterizing Lysine Acetylation of Isocitrate Dehydrogenase in Escherichia coli. J. Mol. Biol. 2018, 430 (13), 19011911,  DOI: 10.1016/j.jmb.2018.04.031
    293. 293
      Venkat, S.; Gregory, C.; Sturges, J.; Gan, Q.; Fan, C. Studying the Lysine Acetylation of Malate Dehydrogenase. J. Mol. Biol. 2017, 429 (9), 13961405,  DOI: 10.1016/j.jmb.2017.03.027
    294. 294
      Venkat, S.; Chen, H.; McGuire, P.; Stahman, A.; Gan, Q.; Fan, C. Characterizing Lysine Acetylation of Escherichia coli Type II Citrate Synthase. FEBS J. 2019, 286 (14), 27992808,  DOI: 10.1111/febs.14845
    295. 295
      Wright, D. E.; Altaany, Z.; Bi, Y.; Alperstein, Z.; O’Donoghue, P. Acetylation Regulates Thioredoxin Reductase Oligomerization and Activity. Antioxid. Redox Signal. 2018, 29 (4), 377388,  DOI: 10.1089/ars.2017.7082
    296. 296
      Rogerson, D. T.; Sachdeva, A.; Wang, K.; Haq, T.; Kazlauskaite, A.; Hancock, S. M.; Huguenin-Dezot, N.; Muqit, M. M. K.; Fry, A. M.; Bayliss, R. Efficient Genetic Encoding of Phosphoserine and its Nonhydrolyzable Analog. Nat. Chem. Biol. 2015, 11 (7), 496503,  DOI: 10.1038/nchembio.1823
    297. 297
      Venkat, S.; Sturges, J.; Stahman, A.; Gregory, C.; Gan, Q.; Fan, C. Genetically Incorporating Two Distinct Post-translational Modifications into One Protein Simultaneously. ACS Synth. Biol. 2018, 7 (2), 689695,  DOI: 10.1021/acssynbio.7b00408
    298. 298
      Zang, J.; Chen, Y.; Liu, C.; Lin, S. Probing the Role of Aurora Kinase A Threonylation with Site-Specific Lysine Threonylation. ACS Chem. Biol. 2023, 18 (4), 674678,  DOI: 10.1021/acschembio.1c00682
    299. 299
      Wan, N.; Wang, N.; Yu, S.; Zhang, H.; Tang, S.; Wang, D.; Lu, W.; Li, H.; Delafield, D. G.; Kong, Y. Cyclic Immonium Ion of Lactyllysine Reveals Widespread Lactylation in the Human Proteome. Nat. Methods 2022, 19 (7), 854864,  DOI: 10.1038/s41592-022-01523-1
    300. 300
      Whittaker, J. W. Free Radical Catalysis by Galactose Oxidase. Chem. Rev. 2003, 103 (6), 23472364,  DOI: 10.1021/cr020425z
    301. 301
      Polyakov, K. M.; Boyko, K. M.; Tikhonova, T. V.; Slutsky, A.; Antipov, A. N.; Zvyagilskaya, R. A.; Popov, A. N.; Bourenkov, G. P.; Lamzin, V. S.; Popov, V. O. High-Resolution Structural Analysis of a Novel Octaheme Cytochrome c Nitrite Reductase from the Haloalkaliphilic Bacterium Thioalkalivibrio nitratireducens. J. Mol. Biol. 2009, 389 (5), 846862,  DOI: 10.1016/j.jmb.2009.04.037
    302. 302
      Ye, S.; Wu, X. a.; Wei, L.; Tang, D.; Sun, P.; Bartlam, M.; Rao, Z. An Insight into the Mechanism of Human Cysteine Dioxygenase: key roles of the thioether-bonded tyrosine-cysteine cofactor. J. Biol. Chem. 2007, 282 (5), 33913402,  DOI: 10.1074/jbc.M609337200
    303. 303
      Zhou, Q.; Hu, M.; Zhang, W.; Jiang, L.; Perrett, S.; Zhou, J.; Wang, J. Probing the Function of the Tyr-Cys Cross-Link in Metalloenzymes by the Genetic Incorporation of 3-Methylthiotyrosine. Angew. Chem. Int. Ed. 2013, 52 (4), 12031207,  DOI: 10.1002/anie.201207229
    304. 304
      Dominy, J. E.; Hwang, J.; Guo, S.; Hirschberger, L. L.; Zhang, S.; Stipanuk, M. H. Synthesis of Amino Acid Cofactor in Cysteine Dioxygenase Is Regulated by Substrate and Represents a Novel Post-translational Regulation of Activity. J. Biol. Chem. 2008, 283 (18), 1218812201,  DOI: 10.1074/jbc.M800044200
    305. 305
      Li, J.; Griffith, W. P.; Davis, I.; Shin, I.; Wang, J.; Li, F.; Wang, Y.; Wherritt, D. J.; Liu, A. Cleavage of a Carbon-Fluorine Bond by an Engineered Cysteine Dioxygenase. Nat. Chem. Biol. 2018, 14 (9), 853860,  DOI: 10.1038/s41589-018-0085-5
    306. 306
      Li, J.; Koto, T.; Davis, I.; Liu, A. Probing the Cys-Tyr Cofactor Biogenesis in Cysteine Dioxygenase by the Genetic Incorporation of Fluorotyrosine. Biochemistry 2019, 58 (17), 22182227,  DOI: 10.1021/acs.biochem.9b00006
    307. 307
      Chen, L.; Naowarojna, N.; Song, H.; Wang, S.; Wang, J.; Deng, Z.; Zhao, C.; Liu, P. Use of a Tyrosine Analogue To Modulate the Two Activities of a Nonheme Iron Enzyme OvoA in Ovothiol Biosynthesis, Cysteine Oxidation versus Oxidative C-S Bond Formation. J. Am. Chem. Soc. 2018, 140 (13), 46044612,  DOI: 10.1021/jacs.7b13628
    308. 308
      Trumpower, B. L.; Gennis, R. B. Energy Transduction By Cytochrome Complexes In Mitochondrial And Bacterial Respiration: The Enzymology of Coupling Electron Transfer Reactions to Transmembrane Proton Translocation. Annu. Rev. Biochem. 1994, 63 (1), 675716,  DOI: 10.1146/annurev.bi.63.070194.003331
    309. 309
      Kim, E.; Chufán, E. E.; Kamaraj, K.; Karlin, K. D. Synthetic Models for Heme-Copper Oxidases. Chem. Rev. 2004, 104 (2), 10771134,  DOI: 10.1021/cr0206162
    310. 310
      Miner, K. D.; Mukherjee, A.; Gao, Y.-G.; Null, E. L.; Petrik, I. D.; Zhao, X.; Yeung, N.; Robinson, H.; Lu, Y. A Designed Functional Metalloenzyme that Reduces O2 to H2O with Over One Thousand Turnovers. Angew. Chem. Int. Ed. 2012, 51 (23), 55895592,  DOI: 10.1002/anie.201201981
    311. 311
      Liu, X.; Yu, Y.; Hu, C.; Zhang, W.; Lu, Y.; Wang, J. Significant Increase of Oxidase Activity through the Genetic Incorporation of a Tyrosine-Histidine Cross-Link in a Myoglobin Model of Heme-Copper Oxidase. Angew. Chem. Int. Ed. 2012, 51 (18), 43124316,  DOI: 10.1002/anie.201108756
    312. 312
      Yu, Y.; Lv, X.; Li, J.; Zhou, Q.; Cui, C.; Hosseinzadeh, P.; Mukherjee, A.; Nilges, M. J.; Wang, J.; Lu, Y. Defining the Role of Tyrosine and Rational Tuning of Oxidase Activity by Genetic Incorporation of Unnatural Tyrosine Analogs. J. Am. Chem. Soc. 2015, 137 (14), 45944597,  DOI: 10.1021/ja5109936
    313. 313
      Yu, Y.; Zhou, Q.; Wang, L.; Liu, X.; Zhang, W.; Hu, M.; Dong, J.; Li, J.; Lv, X.; Ouyang, H. Significant Improvement of Oxidase Activity Through the Genetic Incorporation of a Redox-Active Unnatural Amino Acid. Chem. Sci. 2015, 6 (7), 38813885,  DOI: 10.1039/C5SC01126D
    314. 314
      Rigoldi, F.; Donini, S.; Redaelli, A.; Parisini, E.; Gautieri, A. Review: Engineering of Thermostable Enzymes for Industrial Applications. APL Bioengineering 2018, 2 (1), 011501,  DOI: 10.1063/1.4997367
    315. 315
      Sheldon, R. A.; Basso, A.; Brady, D. New Frontiers in Enzyme Immobilisation: Robust Biocatalysts for a Circular Bio-Based Economy. Chem. Soc. Rev. 2021, 50 (10), 58505862,  DOI: 10.1039/D1CS00015B
    316. 316
      Baker, P. J.; Montclare, J. K. Enhanced Refoldability and Thermoactivity of Fluorinated Phosphotriesterase. ChemBioChem 2011, 12 (12), 18451848,  DOI: 10.1002/cbic.201100221
    317. 317
      Deepankumar, K.; Shon, M.; Nadarajan, S. P.; Shin, G.; Mathew, S.; Ayyadurai, N.; Kim, B.-G.; Choi, S.-H.; Lee, S.-H.; Yun, H. Enhancing Thermostability and Organic Solvent Tolerance of ω-Transaminase through Global Incorporation of Fluorotyrosine. Adv. Synth. Catal. 2014, 356 (5), 993998,  DOI: 10.1002/adsc.201300706
    318. 318
      Ohtake, K.; Mukai, T.; Iraha, F.; Takahashi, M.; Haruna, K.-i.; Date, M.; Yokoyama, K.; Sakamoto, K. Engineering an Automaturing Transglutaminase with Enhanced Thermostability by Genetic Code Expansion with Two Codon Reassignments. ACS Synth. Biol. 2018, 7 (9), 21702176,  DOI: 10.1021/acssynbio.8b00157
    319. 319
      Politzer, P.; Murray, J. S. Halogen Bonding: An Interim Discussion. ChemPhysChem 2013, 14 (2), 278294,  DOI: 10.1002/cphc.201200799
    320. 320
      Scholfield, M. R.; Ford, M. C.; Carlsson, A.-C. C.; Butta, H.; Mehl, R. A.; Ho, P. S. Structure-Energy Relationships of Halogen Bonds in Proteins. Biochemistry 2017, 56 (22), 27942802,  DOI: 10.1021/acs.biochem.7b00022
    321. 321
      Acevedo-Rocha, C. G.; Hoesl, M. G.; Nehring, S.; Royter, M.; Wolschner, C.; Wiltschi, B.; Antranikian, G.; Budisa, N. Non-Canonical Amino Acids as a Useful Synthetic Biological Tool for Lipase-Catalysed Reactions in Hostile Environments. Catal. Sci. Technol. 2013, 3 (5), 11981201,  DOI: 10.1039/c3cy20712a
    322. 322
      Mendel, D.; Ellman, J. A.; Chang, Z.; Veenstra, D. L.; Kollman, P. A.; Schultz, P. G. Probing Protein Stability with Unnatural Amino Acids. Science 1992, 256 (5065), 17981802,  DOI: 10.1126/science.1615324
    323. 323
      Ismail, A. R.; Kashtoh, H.; Baek, K.-H. Temperature-Resistant and Solvent-Tolerant Lipases as Industrial Biocatalysts: Biotechnological Approaches and Applications. Int. J. Biol. Macromol. 2021, 187, 127142,  DOI: 10.1016/j.ijbiomac.2021.07.101
    324. 324
      Vivek, K.; Sandhia, G. S.; Subramaniyan, S. Extremophilic Lipases for Industrial Applications: A General Review. Biotechnol. Adv. 2022, 60, 108002,  DOI: 10.1016/j.biotechadv.2022.108002
    325. 325
      Budisa, N.; Wenger, W.; Wiltschi, B. Residue-Specific Global Fluorination of Candida antarctica Lipase B in Pichia pastoris. Mol. BioSyst. 2010, 6 (9), 16301639,  DOI: 10.1039/c002256j
    326. 326
      Merkel, L.; Schauer, M.; Antranikian, G.; Budisa, N. Parallel Incorporation of Different Fluorinated Amino Acids: On the Way to “Teflon” Proteins. ChemBioChem 2010, 11 (11), 15051507,  DOI: 10.1002/cbic.201000295
    327. 327
      Hoesl, M. G.; Acevedo-Rocha, C. G.; Nehring, S.; Royter, M.; Wolschner, C.; Wiltschi, B.; Budisa, N.; Antranikian, G. Lipase Congeners Designed by Genetic Code Engineering. ChemCatChem 2011, 3 (1), 213221,  DOI: 10.1002/cctc.201000253
    328. 328
      Kelly, S. A.; Pohle, S.; Wharry, S.; Mix, S.; Allen, C. C. R.; Moody, T. S.; Gilmore, B. F. Application of ω-Transaminases in the Pharmaceutical Industry. Chem. Rev. 2018, 118 (1), 349367,  DOI: 10.1021/acs.chemrev.7b00437
    329. 329
      Mathew, S.; Yun, H. ω-Transaminases for the Production of Optically Pure Amines and Unnatural Amino Acids. ACS Catal. 2012, 2 (6), 9931001,  DOI: 10.1021/cs300116n
    330. 330
      Won, Y.; Jeon, H.; Pagar, A. D.; Patil, M. D.; Nadarajan, S. P.; Flood, D. T.; Dawson, P. E.; Yun, H. In vivo Biosynthesis of Tyrosine Analogs and their Concurrent Incorporation into a Residue-Specific Manner for Enzyme Engineering. Chem. Commun. 2019, 55 (100), 1513315136,  DOI: 10.1039/C9CC08503C
    331. 331
      Votchitseva, Y. A.; Efremenko, E. N.; Varfolomeyev, S. D. Insertion of an Unnatural Amino Acid into the Protein Structure: Preparation and Properties of 3-Fluorotyrosine-Containing Organophosphate Hydrolase. Russ. Chem. Bull. 2006, 55 (2), 369374,  DOI: 10.1007/s11172-006-0262-7
    332. 332
      Castro, A. A. d.; Prandi, I. G.; Kuca, K.; Ramalho, T. C. Enzimas Degradantes de Organofosforados: Base Molecular e Perspectivas para Biorremediação Enzimática de Agroquímicos. Ciênc. Agrotec. 2017, 41 (5), 471,  DOI: 10.1590/1413-70542017415000417
    333. 333
      Makkar, R. S.; DiNovo, A. A.; Westwater, C.; Schofield, D. A. Enzyme-Mediated Bioremediation of Organophosphates using Stable Yeast Biocatalysts. J. Bioremed. Biodeg. 2013, 4 (182), 2,  DOI: 10.4172/2155-6199.1000182
    334. 334
      Holzberger, B.; Marx, A. Replacing 32 Proline Residues by a Noncanonical Amino Acid Results in a Highly Active DNA Polymerase. J. Am. Chem. Soc. 2010, 132 (44), 1570815713,  DOI: 10.1021/ja106525y
    335. 335
      Holzberger, B.; Obeid, S.; Welte, W.; Diederichs, K.; Marx, A. Structural Insights into the Potential of 4-Fluoroproline to Modulate Biophysical Properties of Proteins. Chem. Sci. 2012, 3 (10), 29242931,  DOI: 10.1039/c2sc20545a
    336. 336
      Panchenko, T.; Zhu, W. W.; Montclare, J. K. Influence of Global Fluorination on Chloramphenicol Acetyltransferase Activity and Stability. Biotechnol. Bioeng. 2006, 94 (5), 921930,  DOI: 10.1002/bit.20910
    337. 337
      Yang, C.-Y.; Renfrew, P. D.; Olsen, A. J.; Zhang, M.; Yuvienco, C.; Bonneau, R.; Montclare, J. K. Improved Stability and Half-Life of Fluorinated Phosphotriesterase Using Rosetta. ChemBioChem 2014, 15 (12), 17611764,  DOI: 10.1002/cbic.201402062
    338. 338
      Ohtake, K.; Yamaguchi, A.; Mukai, T.; Kashimura, H.; Hirano, N.; Haruki, M.; Kohashi, S.; Yamagishi, K.; Murayama, K.; Tomabechi, Y. Protein Stabilization Utilizing a Redefined Codon. Sci. Rep. 2015, 5 (1), 9762,  DOI: 10.1038/srep09762
    339. 339
      Carlsson, A. C.; Scholfield, M. R.; Rowe, R. K.; Ford, M. C.; Alexander, A. T.; Mehl, R. A.; Ho, P. S. Increasing Enzyme Stability and Activity through Hydrogen Bond-Enhanced Halogen Bonds. Biochemistry 2018, 57 (28), 41354147,  DOI: 10.1021/acs.biochem.8b00603
    340. 340
      Nicholson, H.; Anderson, D. E.; Dao Pin, S.; Matthews, B. W. Analysis of the Interaction Between Charged Side Chains and the Alpha-Helix Dipole Using Designed Thermostable Mutants of Phage T4 Lysozyme. Biochemistry 1991, 30 (41), 98169828,  DOI: 10.1021/bi00105a002
    341. 341
      Klink, T. A.; Woycechowsky, K. J.; Taylor, K. M.; Raines, R. T. Contribution of Disulfide Bonds to the Conformational Stability and Catalytic Activity Of Ribonuclease A. Eur. J. Biochem. 2000, 267 (2), 566572,  DOI: 10.1046/j.1432-1327.2000.01037.x
    342. 342
      Yin, X.; Hu, D.; Li, J.-F.; He, Y.; Zhu, T.-D.; Wu, M.-C. Contribution of Disulfide Bridges to the Thermostability of a Type A Feruloyl Esterase from Aspergillus usamii. PLOS ONE 2015, 10 (5), e0126864  DOI: 10.1371/journal.pone.0126864
    343. 343
      Gihaz, S.; Bash, Y.; Rush, I.; Shahar, A.; Pazy, Y.; Fishman, A. Bridges to Stability: Engineering Disulfide Bonds Towards Enhanced Lipase Biodiesel Synthesis. ChemCatChem 2020, 12 (1), 181192,  DOI: 10.1002/cctc.201901369
    344. 344
      Zhou, X.; Xu, Z.; Li, Y.; He, J.; Zhu, H. Improvement of the Stability and Activity of an LPMO Through Rational Disulfide Bonds Design. Front. bioeng. biotechnol. 2022,  DOI: 10.3389/fbioe.2021.815990
    345. 345
      Dombkowski, A. A.; Sultana, K. Z.; Craig, D. B. Protein Disulfide Engineering. FEBS Lett. 2014, 588 (2), 206212,  DOI: 10.1016/j.febslet.2013.11.024
    346. 346
      Liu, T.; Wang, Y.; Luo, X.; Li, J.; Reed, S. A.; Xiao, H.; Young, T. S.; Schultz, P. G. Enhancing Protein Stability with Extended Disulfide Bonds. Proc. Natl. Acad. Sci. U.S.A. 2016, 113 (21), 59105915,  DOI: 10.1073/pnas.1605363113
    347. 347
      Hecky, J.; Müller, K. M. Structural Perturbation and Compensation by Directed Evolution at Physiological Temperature Leads to Thermostabilization of β-Lactamase. Biochemistry 2005, 44 (38), 1264012654,  DOI: 10.1021/bi0501885
    348. 348
      Brown, N. G.; Pennington, J. M.; Huang, W.; Ayvaz, T.; Palzkill, T. Multiple Global Suppressors of Protein Stability Defects Facilitate the Evolution of Extended-Spectrum TEM β-Lactamases. J. Mol. Biol. 2010, 404 (5), 832846,  DOI: 10.1016/j.jmb.2010.10.008
    349. 349
      Moore, E. J.; Zorine, D.; Hansen, W. A.; Khare, S. D.; Fasan, R. Enzyme Stabilization via Computationally Guided Protein Stapling. Proc. Natl. Acad. Sci. U.S.A. 2017, 114 (47), 1247212477,  DOI: 10.1073/pnas.1708907114
    350. 350
      Bi, J.; Jing, X.; Wu, L.; Zhou, X.; Gu, J.; Nie, Y.; Xu, Y. Computational Design of Noncanonical Amino Acid-Based Thioether Staples at N/C-Terminal Domains of Multi-Modular Pullulanase for Thermostabilization in Enzyme Catalysis. Comput. Struct. Biotechnol. J. 2021, 19, 577585,  DOI: 10.1016/j.csbj.2020.12.043
    351. 351
      Iannuzzelli, J. A.; Bacik, J.-P.; Moore, E. J.; Shen, Z.; Irving, E. M.; Vargas, D. A.; Khare, S. D.; Ando, N.; Fasan, R. Tuning Enzyme Thermostability via Computationally Guided Covalent Stapling and Structural Basis of Enhanced Stabilization. Biochemistry 2022, 61 (11), 10411054,  DOI: 10.1021/acs.biochem.2c00033
    352. 352
      Gur, E.; Biran, D.; Gazit, E.; Ron, E. Z. In vivo Aggregation of a Single Enzyme Limits Growth of Escherichia coli at Elevated Temperatures. Mol. Microbiol. 2002, 46 (5), 13911397,  DOI: 10.1046/j.1365-2958.2002.03257.x
    353. 353
      Li, J. C.; Liu, T.; Wang, Y.; Mehta, A. P.; Schultz, P. G. Enhancing Protein Stability with Genetically Encoded Noncanonical Amino Acids. J. Am. Chem. Soc. 2018, 140 (47), 1599716000,  DOI: 10.1021/jacs.8b07157
    354. 354
      Xuan, W.; Li, J.; Luo, X.; Schultz, P. G. Genetic Incorporation of a Reactive Isothiocyanate Group into Proteins. Angew. Chem. Int. Ed. 2016, 55 (34), 1006510068,  DOI: 10.1002/anie.201604891
    355. 355
      Li, J. C.; Nastertorabi, F.; Xuan, W.; Han, G. W.; Stevens, R. C.; Schultz, P. G. A Single Reactive Noncanonical Amino Acid Is Able to Dramatically Stabilize Protein Structure. ACS Chem. Biol. 2019, 14 (6), 11501153,  DOI: 10.1021/acschembio.9b00002
    356. 356
      Deiters, A.; Cropp, T. A.; Summerer, D.; Mukherji, M.; Schultz, P. G. Site-Specific PEGylation of Proteins Containing Unnatural Amino Acids. Bioorg. Med. Chem. Lett. 2004, 14 (23), 57435745,  DOI: 10.1016/j.bmcl.2004.09.059
    357. 357
      Schoffelen, S.; Lambermon, M. H. L.; Eldijk, M. B. v.; Hest, J. C. M. v. Site-Specific Modification of Candida antarctica Lipase B via Residue-Specific Incorporation of a Non-Canonical Amino Acid. Bioconjug. Chem. 2008, 19 (6), 11271131,  DOI: 10.1021/bc800019v
    358. 358
      van Dongen, S. F. M.; Nallani, M.; Schoffelen, S.; Cornelissen, J. J. L. M.; Nolte, R. J. M.; van Hest, J. C. M. A Block Copolymer for Functionalisation of Polymersome Surfaces. Macromol. Rapid Commun. 2008, 29 (4), 321325,  DOI: 10.1002/marc.200700765
    359. 359
      Teeuwen, R. L. M.; van Berkel, S. S.; van Dulmen, T. H. H.; Schoffelen, S.; Meeuwissen, S. A.; Zuilhof, H.; de Wolf, F. A.; van Hest, J. C. M. “Clickable” Elastins: Elastin-Like Polypeptides Functionalized with Azide or Alkyne Groups. Chem. Commun. 2009, (27), 40224024,  DOI: 10.1039/b903903a
    360. 360
      Debets, M. F.; van Berkel, S. S.; Schoffelen, S.; Rutjes, F. P. J. T.; van Hest, J. C. M.; van Delft, F. L. Aza-Dibenzocyclooctynes for Fast and Efficient Enzyme PEGylation via Copper-Free (3 + 2) Cycloaddition. Chem. Commun. 2010, 46 (1), 9799,  DOI: 10.1039/B917797C
    361. 361
      Wilding, K. M.; Smith, A. K.; Wilkerson, J. W.; Bush, D. B.; Knotts, T. A. I. V.; Bundy, B. C. The Locational Impact of Site-Specific PEGylation: Streamlined Screening with Cell-Free Protein Expression and Coarse-Grain Simulation. ACS Synth. Biol. 2018, 7 (2), 510521,  DOI: 10.1021/acssynbio.7b00316
    362. 362
      Wu, J. C. Y.; Hutchings, C. H.; Lindsay, M. J.; Werner, C. J.; Bundy, B. C. Enhanced Enzyme Stability Through Site-Directed Covalent Immobilization. J. Biotechnol. 2015, 193, 8390,  DOI: 10.1016/j.jbiotec.2014.10.039
    363. 363
      Basso, A.; Serban, S. Industrial Applications of Immobilized Enzymes─A Review. Mol. Catal. 2019, 479, 110607,  DOI: 10.1016/j.mcat.2019.110607
    364. 364
      Hernandez, K.; Fernandez-Lafuente, R. Control of Protein Immobilization: Coupling Immobilization and Site-Directed Mutagenesis to Improve Biocatalyst or Biosensor Performance. Enzyme Microb. Technol. 2011, 48 (2), 107122,  DOI: 10.1016/j.enzmictec.2010.10.003
    365. 365
      Guan, D.; Kurra, Y.; Liu, W.; Chen, Z. A Click Chemistry Approach to Site-Specific Immobilization of a Small Laccase Enables Efficient Direct Electron Transfer in a Biocathode. Chem. Commun. 2015, 51 (13), 25222525,  DOI: 10.1039/C4CC09179E
    366. 366
      Lim, S. I.; Mizuta, Y.; Takasu, A.; Kim, Y. H.; Kwon, I. Site-Specific Bioconjugation of a Murine Dihydrofolate Reductase Enzyme by Copper(I)-Catalyzed Azide-Alkyne Cycloaddition with Retained Activity. PLOS ONE 2014, 9 (6), e98403  DOI: 10.1371/journal.pone.0098403
    367. 367
      Wang, A.; Du, F.; Pei, X.; Chen, C.; Wu, S. G.; Zheng, Y. Rational Immobilization of Lipase by Combining the Structure Analysis and Unnatural Amino Acid Insertion. J. Mol. Catal. B Enzym. 2016, 132, 5460,  DOI: 10.1016/j.molcatb.2016.06.015
    368. 368
      Li, H.; Yin, Y.; Wang, A.; Li, N.; Wang, R.; Zhang, J.; Chen, X.; Pei, X.; Xie, T. Stable Immobilization of Aldehyde Ketone Reductase Mutants Containing Nonstandard Amino Acids on an Epoxy Resin via Strain-Promoted Alkyne-Azide Cycloaddition. RSC Adv. 2020, 10 (5), 26242633,  DOI: 10.1039/C9RA09067C
    369. 369
      Umeda, A.; Thibodeaux, G. N.; Zhu, J.; Lee, Y.; Zhang, Z. J. Site-specific Protein Cross-Linking with Genetically Incorporated 3,4-Dihydroxy-L-Phenylalanine. ChemBioChem 2009, 10 (8), 13021304,  DOI: 10.1002/cbic.200900127
    370. 370
      Deepankumar, K.; Nadarajan, S. P.; Mathew, S.; Lee, S.-G.; Yoo, T. H.; Hong, E. Y.; Kim, B.-G.; Yun, H. Engineering Transaminase for Stability Enhancement and Site-Specific Immobilization through Multiple Noncanonical Amino Acids Incorporation. ChemCatChem 2015, 7 (3), 417421,  DOI: 10.1002/cctc.201402882
    371. 371
      Bednar, R. M.; Golbek, T. W.; Kean, K. M.; Brown, W. J.; Jana, S.; Baio, J. E.; Karplus, P. A.; Mehl, R. A. Immobilization of Proteins with Controlled Load and Orientation. ACS Appl. Mater. Interfaces 2019, 11 (40), 3639136398,  DOI: 10.1021/acsami.9b12746
    372. 372
      Blizzard, R. J.; Backus, D. R.; Brown, W.; Bazewicz, C. G.; Li, Y.; Mehl, R. A. Ideal Bioorthogonal Reactions Using A Site-Specifically Encoded Tetrazine Amino Acid. J. Am. Chem. Soc. 2015, 137 (32), 1004410047,  DOI: 10.1021/jacs.5b03275
    373. 373
      Switzer, H. J.; Howard, C. A.; Halonski, J. F.; Peairs, E. M.; Smith, N.; Zamecnik, M. P.; Verma, S.; Young, D. D. Employing Non-Canonical Amino Acids Towards the Immobilization of a Hyperthermophilic Enzyme to Increase Protein Stability. RSC Adv. 2023, 13 (13), 84968501,  DOI: 10.1039/D3RA00392B
    374. 374
      Ray, S.; Chand, S.; Zhang, Y.; Nussbaum, S.; Rajeshwar, K.; Perera, R. Implications of Active Site Orientation in Myoglobin for Direct Electron Transfer and Electrocatalysis Based on Monolayer and Multilayer Covalent Immobilization on Gold Electrodes. Electrochim. Acta 2013, 99, 8593,  DOI: 10.1016/j.electacta.2013.03.080
    375. 375
      Xia, L.; Han, M.-J.; Zhou, L.; Huang, A.; Yang, Z.; Wang, T.; Li, F.; Yu, L.; Tian, C.; Zang, Z. S-Click Reaction for Isotropic Orientation of Oxidases on Electrodes to Promote Electron Transfer at Low Potentials. Angew. Chem. Int. Ed. 2019, 58 (46), 1648016484,  DOI: 10.1002/anie.201909203
    376. 376
      Pan, Y.; Li, G.; Liu, R.; Guo, J.; Liu, Y.; Liu, M.; Zhang, X.; Chi, L.; Xu, K.; Wu, R. Unnatural Activities and Mechanistic Insights of Cytochrome P450 PikC Gained from Site-Specific Mutagenesis by Non-Canonical Amino Acids. Nature Commun. 2023, 14 (1), 1669,  DOI: 10.1038/s41467-023-37288-0
    377. 377
      Kolev, J. N.; Zaengle, J. M.; Ravikumar, R.; Fasan, R. Enhancing the Efficiency and Regioselectivity of P450 Oxidation Catalysts by Unnatural Amino Acid Mutagenesis. ChemBioChem 2014, 15 (7), 10011010,  DOI: 10.1002/cbic.201400060
    378. 378
      Ma, H.; Yang, X.; Lu, Z.; Liu, N.; Chen, Y. The “Gate Keeper” Role of Trp222 Determines the Enantiopreference of Diketoreductase toward 2-Chloro-1-Phenylethanone. PLOS ONE 2014, 9 (7), e103792  DOI: 10.1371/journal.pone.0103792
    379. 379
      Yu, Z.; Yu, H.; Tang, H.; Wang, Z.; Wu, J.; Yang, L.; Xu, G. Site-specifically Incorporated Non-Canonical Amino Acids into Pseudomonas alcaligenes Lipase to Hydrolyze L-menthol Propionate among the Eight Isomers. ChemCatChem 2021, 13 (11), 26912701,  DOI: 10.1002/cctc.202100358
    380. 380
      Drienovská, I.; Gajdoš, M.; Kindler, A.; Takhtehchian, M.; Darnhofer, B.; Birner-Gruenberger, R.; Dörr, M.; Bornscheuer, U. T.; Kourist, R. Folding Assessment of Incorporation of Noncanonical Amino Acids Facilitates Expansion of Functional-Group Diversity for Enzyme Engineering. Chem. Eur. J. 2020, 26 (54), 1233812342,  DOI: 10.1002/chem.202002077
    381. 381
      Green, A. P.; Hayashi, T.; Mittl, P. R. E.; Hilvert, D. A Chemically Programmed Proximal Ligand Enhances the Catalytic Properties of a Heme Enzyme. J. Am. Chem. Soc. 2016, 138 (35), 1134411352,  DOI: 10.1021/jacs.6b07029
    382. 382
      Sharp, K. H.; Mewies, M.; Moody, P. C. E.; Raven, E. L. Crystal Structure of the Ascorbate Peroxidase-Ascorbate Complex. Nat. Struct. Biol. 2003, 10 (4), 303307,  DOI: 10.1038/nsb913
    383. 383
      Vojtechovský, J.; Chu, K.; Berendzen, J.; Sweet, R. M.; Schlichting, I. Crystal Structures of Myoglobin-Ligand Complexes at Near-Atomic Resolution. Biophys. J. 1999, 77 (4), 21532174,  DOI: 10.1016/S0006-3495(99)77056-6
    384. 384
      Pott, M.; Hayashi, T.; Mori, T.; Mittl, P. R. E.; Green, A. P.; Hilvert, D. A Noncanonical Proximal Heme Ligand Affords an Efficient Peroxidase in a Globin Fold. J. Am. Chem. Soc. 2018, 140 (4), 15351543,  DOI: 10.1021/jacs.7b12621
    385. 385
      Matsuo, T.; Fukumoto, K.; Watanabe, T.; Hayashi, T. Precise Design of Artificial Cofactors for Enhancing Peroxidase Activity of Myoglobin: Myoglobin Mutant H64D Reconstituted with a ″Single-Winged Cofactor″ Is Equivalent to Native Horseradish Peroxidase in Oxidation Activity. Chem. Asian J. 2011, 6 (9), 24912499,  DOI: 10.1002/asia.201100107
    386. 386
      Hayashi, T.; Tinzl, M.; Mori, T.; Krengel, U.; Proppe, J.; Soetbeer, J.; Klose, D.; Jeschke, G.; Reiher, M.; Hilvert, D. Capture and Characterization of a Reactive Haem-Carbenoid Complex in an Artificial Metalloenzyme. Nat. Catal. 2018, 1 (8), 578584,  DOI: 10.1038/s41929-018-0105-6
    387. 387
      Pott, M.; Tinzl, M.; Hayashi, T.; Ota, Y.; Dunkelmann, D. L.; Mittl, P. R. E.; Hilvert, D. Noncanonical Heme Ligands Steer Carbene Transfer Reactivity in an Artificial Metalloenzyme. Angew. Chem. Int. Ed 2021, 60, 1506315068,  DOI: 10.1002/anie.202103437
    388. 388
      Carminati, D. M.; Fasan, R. Stereoselective Cyclopropanation of Electron-Deficient Olefins with a Cofactor Redesigned Carbene Transferase Featuring Radical Reactivity. ACS Catal. 2019, 9 (10), 96839697,  DOI: 10.1021/acscatal.9b02272
    389. 389
      Moore, E. J.; Fasan, R. Effect of Proximal Ligand Substitutions on the Carbene and Nitrene Transferase Activity of Myoglobin. Tetrahedron 2019, 75 (16), 23572363,  DOI: 10.1016/j.tet.2019.03.009
    390. 390
      Gan, F.; Liu, R.; Wang, F.; Schultz, P. G. Functional Replacement of Histidine in Proteins To Generate Noncanonical Amino Acid Dependent Organisms. J. Am. Chem. Soc. 2018, 140 (11), 38293832,  DOI: 10.1021/jacs.7b13452
    391. 391
      Chand, S.; Ray, S.; Yadav, P.; Samanta, S.; Pierce, B. S.; Perera, R. Abiological Catalysis by Myoglobin Mutant with a Genetically Incorporated Unnatural Amino Acid. Biochem. J. 2021, 478 (9), 17951808,  DOI: 10.1042/BCJ20210091
    392. 392
      Jackson, J. C.; Duffy, S. P.; Hess, K. R.; Mehl, R. A. Improving Nature’s Enzyme Active Site with Genetically Encoded Unnatural Amino Acids. J. Am. Chem. Soc. 2006, 128 (34), 1112411127,  DOI: 10.1021/ja061099y
    393. 393
      Grove, J. I.; Lovering, A. L.; Guise, C.; Race, P. R.; Wrighton, C. J.; White, S. A.; Hyde, E. I.; Searle, P. F. Generation of Escherichia coli Nitroreductase Mutants Conferring Improved Cell Sensitization to the Prodrug CB19541. Cancer Res. 2003, 63 (17), 55325537
    394. 394
      Ugwumba, I. N.; Ozawa, K.; Xu, Z.-Q.; Ely, F.; Foo, J.-L.; Herlt, A. J.; Coppin, C.; Brown, S.; Taylor, M. C.; Ollis, D. L. Improving a Natural Enzyme Activity through Incorporation of Unnatural Amino Acids. J. Am. Chem. Soc. 2011, 133 (2), 326333,  DOI: 10.1021/ja106416g
    395. 395
      Pagar, A. D.; Jeon, H.; Khobragade, T. P.; Sarak, S.; Giri, P.; Lim, S.; Yoo, T. H.; Ko, B. J.; Yun, H. Non-Canonical Amino Acid-Based Engineering of (R)-Amine Transaminase. Front. Chem. 2022,  DOI: 10.3389/fchem.2022.839636
    396. 396
      Savile, C. K.; Janey, J. M.; Mundorff, E. C.; Moore, J. C.; Tam, S.; Jarvis, W. R.; Colbeck, J. C.; Krebber, A.; Fleitz, F. J.; Brands, J. Biocatalytic Asymmetric Synthesis of Chiral Amines from Ketones Applied to Sitagliptin Manufacture. Science 2010, 329 (5989), 305309,  DOI: 10.1126/science.1188934
    397. 397
      Wilkinson, H. C.; Dalby, P. A. Fine-Tuning the Activity and Stability of an Evolved Enzyme Active-Site Through Noncanonical Amino-Acids. FEBS J. 2021, 288 (6), 19351955,  DOI: 10.1111/febs.15560
    398. 398
      Parsons, J. F.; Xiao, G.; Gilliland, G. L.; Armstrong, R. N. Enzymes Harboring Unnatural Amino Acids: Mechanistic and Structural Analysis of the Enhanced Catalytic Activity of a Glutathione Transferase Containing 5-Fluorotryptophan. Biochemistry 1998, 37 (18), 62866294,  DOI: 10.1021/bi980219e
    399. 399
      Dominguez, M. A., Jr; Thornton, K. C.; Melendez, M. G.; Dupureur, C. M. Differential Effects of Isomeric Incorporation of Fluorophenylalanines into PvuII Endonuclease. Proteins 2001, 45 (1), 5561,  DOI: 10.1002/prot.1123
    400. 400
      Hoesl, M. G.; Budisa, N. Expanding and Engineering the Genetic Code in a Single Expression Experiment. ChemBioChem 2011, 12, 552555,  DOI: 10.1002/cbic.201000586
    401. 401
      Cirino, P. C.; Tang, Y.; Takahashi, K.; Tirrell, D. A.; Arnold, F. H. Global Incorporation of Norleucine in Place of Methionine in Cytochrome P450 BM-3 Heme Domain Increases Peroxygenase Activity. Biotechnol. Bioeng. 2003, 83 (6), 729734,  DOI: 10.1002/bit.10718
    402. 402
      Xiao, H.; Nasertorabi, F.; Choi, S.-h.; Han, G. W.; Reed, S. A.; Stevens, R. C.; Schultz, P. G. Exploring the Potential Impact of an Expanded Genetic Code on Protein Function. Proc. Natl. Acad. Sci. U.S.A. 2015, 112 (22), 69616966,  DOI: 10.1073/pnas.1507741112
    403. 403
      Young, T. S.; Ahmad, I.; Yin, J. A.; Schultz, P. G. An Enhanced System for Unnatural Amino Acid Mutagenesis in E. coli. J. Mol. Biol. 2010, 395, 361374,  DOI: 10.1016/j.jmb.2009.10.030
    404. 404
      Schoffelen, S.; Beekwilder, J.; Debets, M. F.; Bosch, D.; Hest, J. C. M. v. Construction of a Multifunctional Enzyme Complex via the Strain-Promoted Azide-Alkyne Cycloaddition. Bioconjug. Chem. 2013, 24 (6), 987996,  DOI: 10.1021/bc400021j
    405. 405
      Lim, S. I.; Cho, J.; Kwon, I. Double Clicking for Site-Specific Coupling of Multiple Enzymes. Chem. Commun. 2015, 51 (71), 1360713610,  DOI: 10.1039/C5CC04611D
    406. 406
      Lim, S. I.; Yang, B.; Jung, Y.; Cha, J.; Cho, J.; Choi, E.-S.; Kim, Y. H.; Kwon, I. Controlled Orientation of Active Sites in a Nanostructured Multienzyme Complex. Sci. Rep. 2016, 6 (1), 39587,  DOI: 10.1038/srep39587
    407. 407
      Ha, J. M.; Jeon, S. T.; Yoon, H. J.; Lee, H. H. Formate Dehydrogenase. PDB 2014,  DOI: 10.2210/pdb3WR5/pdb
    408. 408
      Kavanagh, K. L.; Klimacek, M.; Nidetzky, B.; Wilson, D. K. Crystal Structure of Pseudomonas fluorescens Mannitol 2-Dehydrogenase Binary and Ternary Complexes: Specificity And Catalytic Mechanism. J. Biol. Chem. 2002, 277 (45), 4343343442,  DOI: 10.1074/jbc.M206914200
    409. 409
      Li, J.; Jia, S.; Chen, P. R. Diels-Alder Reaction-Triggered Bioorthogonal Protein Decaging in Living Cells. Nat. Chem. Biol. 2014, 10 (12), 10031005,  DOI: 10.1038/nchembio.1656
    410. 410
      Li, J.; Yu, J.; Zhao, J.; Wang, J.; Zheng, S.; Lin, S.; Chen, L.; Yang, M.; Jia, S.; Zhang, X. Palladium-Triggered Deprotection Chemistry for Protein Activation in Living Cells. Nat. Chem. 2014, 6 (4), 352361,  DOI: 10.1038/nchem.1887
    411. 411
      Wang, J.; Zheng, S.; Liu, Y.; Zhang, Z.; Lin, Z.; Li, J.; Zhang, G.; Wang, X.; Li, J.; Chen, P. R. Palladium-Triggered Chemical Rescue of Intracellular Proteins via Genetically Encoded Allene-Caged Tyrosine. J. Am. Chem. Soc. 2016, 138 (46), 1511815121,  DOI: 10.1021/jacs.6b08933
    412. 412
      Georgianna, W. E.; Lusic, H.; McIver, A. L.; Deiters, A. Photocleavable Polyethylene Glycol for the Light-Regulation of Protein Function. Bioconjug. Chem. 2010, 21 (8), 14041407,  DOI: 10.1021/bc100084n
    413. 413
      Wu, N.; Deiters, A.; Cropp, T. A.; King, D.; Schultz, P. G. A Genetically Encoded Photocaged Amino Acid. J. Am. Chem. Soc. 2004, 126 (44), 1430614307,  DOI: 10.1021/ja040175z
    414. 414
      Deiters, A.; Groff, D.; Ryu, Y.; Xie, J.; Schultz, P. G. A Genetically Encoded Photocaged Tyrosine. Angew. Chem. Int. Ed. 2006, 45 (17), 27282731,  DOI: 10.1002/anie.200600264
    415. 415
      Chou, C.; Young, D. D.; Deiters, A. A Light-Activated DNA Polymerase. Angew. Chem. Int. Ed. 2009, 48 (32), 59505953,  DOI: 10.1002/anie.200901115
    416. 416
      Chou, C.; Young, D. D.; Deiters, A. Photocaged T7 RNA Polymerase for the Light Activation of Transcription and Gene Function in Pro- and Eukaryotic Cells. ChemBioChem 2010, 11 (7), 972977,  DOI: 10.1002/cbic.201000041
    417. 417
      Chou, C.; Deiters, A. Light-Activated Gene Editing with a Photocaged Zinc-Finger Nuclease. Angew. Chem. Int. Ed. 2011, 50 (30), 6839,  DOI: 10.1002/anie.201101157
    418. 418
      Gautier, A.; Nguyen, D. P.; Lusic, H.; An, W.; Deiters, A.; Chin, J. W. Genetically Encoded Photocontrol of Protein Localization in Mammalian Cells. J. Am. Chem. Soc. 2010, 132 (12), 40864088,  DOI: 10.1021/ja910688s
    419. 419
      Gautier, A.; Deiters, A.; Chin, J. W. Light-Activated Kinases Enable Temporal Dissection of Signaling Networks in Living Cells. J. Am. Chem. Soc. 2011, 133 (7), 21242127,  DOI: 10.1021/ja1109979
    420. 420
      Hemphill, J.; Chou, C.; Chin, J. W.; Deiters, A. Genetically Encoded Light-Activated Transcription for Spatiotemporal Control of Gene Expression and Gene Silencing in Mammalian Cells. J. Am. Chem. Soc. 2013, 135 (36), 1343313439,  DOI: 10.1021/ja4051026
    421. 421
      Hemphill, J.; Borchardt, E. K.; Brown, K.; Asokan, A.; Deiters, A. Optical Control of CRISPR/Cas9 Gene Editing. J. Am. Chem. Soc. 2015, 137 (17), 56425645,  DOI: 10.1021/ja512664v
    422. 422
      Nguyen, D. P.; Mahesh, M.; Elsässer, S. J.; Hancock, S. M.; Uttamapinant, C.; Chin, J. W. Genetic Encoding of Photocaged Cysteine Allows Photoactivation of TEV Protease in Live Mammalian Cells. J. Am. Chem. Soc. 2014, 136 (6), 22402243,  DOI: 10.1021/ja412191m
    423. 423
      Uprety, R.; Luo, J.; Liu, J.; Naro, Y.; Samanta, S.; Deiters, A. Genetic Encoding of Caged Cysteine and Caged Homocysteine in Bacterial and Mammalian Cells. ChemBioChem 2014, 15 (12), 17931799,  DOI: 10.1002/cbic.201400073
    424. 424
      Yang, X.; Zhao, L.; Wang, Y.; Ji, Y.; Su, X. C.; Ma, J. A.; Xuan, W. Constructing Photoactivatable Protein with Genetically Encoded Photocaged Glutamic Acid. Angew. Chem. Int. Ed. 2023, 135 (40), e202308472  DOI: 10.1002/ange.202308472
    425. 425
      Givens, R. S.; Weber, J. F.; Jung, A. H.; Park, C.-H. New Photoprotecting Groups: Desyl and p-Hydroxyphenacyl Phosphate and Carboxylate Esters. In Methods in enzymology; Elsevier, 1998; Vol. 291, pp 129. DOI: 10.1016/s0076-6879(98)91004-7 .
    426. 426
      Mangubat-Medina, A. E.; Ball, Z. T. Triggering Biological Processes: Methods and Applications of Photocaged Peptides and Proteins. Chem. Soc. Rev. 2021, 50 (18), 1040310421,  DOI: 10.1039/D0CS01434F
    427. 427
      Grasso, K. T.; Singha Roy, S. J.; Osgood, A. O.; Yeo, M. J. R.; Soni, C.; Hillenbrand, C. M.; Ficaretta, E. D.; Chatterjee, A. A Facile Platform to Engineer Escherichia coli Tyrosyl-tRNA Synthetase Adds New Chemistries to the Eukaryotic Genetic Code, Including a Phosphotyrosine Mimic. ACS Cent. Sci. 2022, 8 (4), 483492,  DOI: 10.1021/acscentsci.1c01465
    428. 428
      Edwards, W. F.; Young, D. D.; Deiters, A. Light-Activated Cre Recombinase as a Tool for the Spatial and Temporal Control of Gene Function in Mammalian Cells. ACS Chem. Biol. 2009, 4 (6), 441445,  DOI: 10.1021/cb900041s
    429. 429
      Le Provost, F.; Lillico, S.; Passet, B.; Young, R.; Whitelaw, B.; Vilotte, J.-L. Zinc Finger Nuclease Technology Heralds a New Era in Mammalian Transgenesis. Trends Biotechnol. 2010, 28 (3), 134141,  DOI: 10.1016/j.tibtech.2009.11.007
    430. 430
      Porteus, M. H. Mammalian Gene Targeting with Designed Zinc Finger Nucleases. Molecular Therapy 2006, 13 (2), 438446,  DOI: 10.1016/j.ymthe.2005.08.003
    431. 431
      Porteus, M. H.; Carroll, D. Gene Targeting Using Zinc Finger Nucleases. Nat. Biotechnol. 2005, 23 (8), 967973,  DOI: 10.1038/nbt1125
    432. 432
      Urnov, F. D.; Rebar, E. J.; Holmes, M. C.; Zhang, H. S.; Gregory, P. D. Genome Editing with Engineered Zinc Finger Nucleases. Nat. Rev. Genet. 2010, 11 (9), 636646,  DOI: 10.1038/nrg2842
    433. 433
      Rémy, S.; Tesson, L.; Ménoret, S.; Usal, C.; Scharenberg, A. M.; Anegon, I. Zinc-Finger Nucleases: A Powerful Tool for Genetic Engineering of Animals. Transgenic Res. 2010, 19, 363371,  DOI: 10.1007/s11248-009-9323-7
    434. 434
      Li, Y.; Korolev, S.; Waksman, G. Crystal Structures of Open and Closed Forms of Binary and Ternary Complexes of the Large Fragment of Thermus aquaticus DNA Polymerase I: Structural Basis for Nucleotide Incorporation. EMBO J. 1998, 17 (24), 75147525,  DOI: 10.1093/emboj/17.24.7514
    435. 435
      Suzuki, M.; Baskin, D.; Hood, L.; Loeb, L. A. Random Mutagenesis of Thermus aquaticus DNA Polymerase I: Concordance of Immutable Sites in vivo with the Crystal Structure. Proc. Natl. Acad. Sci. U.S.A. 1996, 93 (18), 96709675,  DOI: 10.1073/pnas.93.18.9670
    436. 436
      Paul, N.; Shum, J.; Le, T. Hot Start PCR. RT-PCR Protocols, Second ed.; 2010; pp 301318. DOI: 10.1007/978-1-60761-629-0
    437. 437
      Cramer, P. Common Structural Features of Nucleic Acid Polymerases. Bioessays 2002, 24 (8), 724729,  DOI: 10.1002/bies.10127
    438. 438
      Banghart, M. R.; Mourot, A.; Fortin, D. L.; Yao, J. Z.; Kramer, R. H.; Trauner, D. Photochromic Blockers of Voltage-Gated Potassium Channels. Angew. Chem. Int. Ed. 2009, 48 (48), 90979101,  DOI: 10.1002/anie.200904504
    439. 439
      Beharry, A. A.; Wong, L.; Tropepe, V.; Woolley, G. A. Fluorescence Imaging of Azobenzene Photoswitching in vivo. Angew. Chem. Int. Ed. 2011, 50 (6), 13251327,  DOI: 10.1002/anie.201006506
    440. 440
      Bléger, D.; Schwarz, J.; Brouwer, A. M.; Hecht, S. o-Fluoroazobenzenes as Readily Synthesized Photoswitches Offering Nearly Quantitative Two-Way Isomerization with Visible Light. J. Am. Chem. Soc. 2012, 134 (51), 2059720600,  DOI: 10.1021/ja310323y
    441. 441
      Bonardi, F.; London, G.; Nouwen, N.; Feringa, B. L.; Driessen, A. J. Light-Induced Control of Protein Translocation by the SecYEG Complex. Angew. Chem. Int. Ed. 2010, 49 (40), 72347238,  DOI: 10.1002/anie.201002243
    442. 442
      Gorostiza, P.; Isacoff, E. Y. Nanoengineering Ion Channels for Optical Control. Physiology 2008, 23 (5), 238247,  DOI: 10.1152/physiol.00018.2008
    443. 443
      Knie, C.; Utecht, M.; Zhao, F.; Kulla, H.; Kovalenko, S.; Brouwer, A. M.; Saalfrank, P.; Hecht, S.; Bléger, D. ortho-Fluoroazobenzenes: Visible Light Switches with Very Long-Lived Z Isomers. Chem. Eur. J. 2014, 20 (50), 1649216501,  DOI: 10.1002/chem.201404649
    444. 444
      Liang, X.; Mochizuki, T.; Asanuma, H. A Supra-Photoswitch Involving Sandwiched DNA Base Pairs and Azobenzenes for Light-Driven Nanostructures and Nanodevices. Small 2009, 5 (15), 17611768,  DOI: 10.1002/smll.200900223
    445. 445
      Samanta, S.; Beharry, A. A.; Sadovski, O.; McCormick, T. M.; Babalhavaeji, A.; Tropepe, V.; Woolley, G. A. Photoswitching Azo Compounds in vivo with Red Light. J. Am. Chem. Soc. 2013, 135 (26), 97779784,  DOI: 10.1021/ja402220t
    446. 446
      Schierling, B.; Noel, A.-J.; Wende, W.; Hien, L. T.; Volkov, E.; Kubareva, E.; Oretskaya, T.; Kokkinidis, M.; Rompp, A.; Spengler, B.; Pingoud, A. Controlling the Enzymatic Activity of a Restriction Enzyme by Light. Proc. Natl. Acad. Sci. U.S.A. 2010, 107 (4), 13611366,  DOI: 10.1073/pnas.0909444107
    447. 447
      Zhang, F.; Zarrine-Afsar, A.; Al-Abdul-Wahid, M. S.; Prosser, R. S.; Davidson, A. R.; Woolley, G. A. Structure-Based Approach to the Photocontrol of Protein Folding. J. Am. Chem. Soc. 2009, 131 (6), 22832289,  DOI: 10.1021/ja807938v
    448. 448
      Bose, M.; Groff, D.; Xie, J.; Brustad, E.; Schultz, P. G. The Incorporation of a Photoisomerizable Amino Acid into Proteins in E. c oli. J. Am. Chem. Soc. 2006, 128 (2), 388389,  DOI: 10.1021/ja055467u
    449. 449
      Hoppmann, C.; Lacey, V. K.; Louie, G. V.; Wei, J.; Noel, J. P.; Wang, L. Genetically Encoding Photoswitchable Click Amino Acids in Escherichia coli and Mammalian Cells. Angew. Chem. Int. Ed. 2014, 53 (15), 39323936,  DOI: 10.1002/anie.201400001
    450. 450
      Hoppmann, C.; Maslennikov, I.; Choe, S.; Wang, L. In situ Formation of an Azo Bridge on Proteins Controllable by Visible Light. J. Am. Chem. Soc. 2015, 137 (35), 1121811221,  DOI: 10.1021/jacs.5b06234
    451. 451
      John, A. A.; Ramil, C. P.; Tian, Y.; Cheng, G.; Lin, Q. Synthesis and Site-Specific Incorporation of Red-Shifted Azobenzene Amino Acids into Proteins. Org. Lett. 2015, 17 (24), 62586261,  DOI: 10.1021/acs.orglett.5b03268
    452. 452
      Klippenstein, V.; Hoppmann, C.; Ye, S.; Wang, L.; Paoletti, P. Optocontrol of Glutamate Receptor Activity by Single Side-Chain Photoisomerization. Elife 2017, 6, e25808  DOI: 10.7554/eLife.25808
    453. 453
      Kneuttinger, A. C.; Straub, K.; Bittner, P.; Simeth, N. A.; Bruckmann, A.; Busch, F.; Rajendran, C.; Hupfeld, E.; Wysocki, V. H.; Horinek, D. Light Regulation of Enzyme Allostery through Photo-responsive Unnatural Amino Acids. Cell Chem. Biol. 2019, 26 (11), 15011514,  DOI: 10.1016/j.chembiol.2019.08.006
    454. 454
      Luo, J.; Samanta, S.; Convertino, M.; Dokholyan, N. V.; Deiters, A. Reversible and Tunable Photoswitching of Protein Function through Genetic Encoding of Azobenzene Amino Acids in Mammalian Cells. ChemBioChem 2018, 19 (20), 21782185,  DOI: 10.1002/cbic.201800226
    455. 455
      Kneuttinger, A. C.; Winter, M.; Simeth, N. A.; Heyn, K.; Merkl, R.; König, B.; Sterner, R. Artificial Light Regulation of an Allosteric Bienzyme Complex by a Photosensitive Ligand. ChemBioChem 2018, 19 (16), 17501757,  DOI: 10.1002/cbic.201800219
    456. 456
      Kneuttinger, A. C.; Rajendran, C.; Simeth, N. A.; Bruckmann, A.; König, B.; Sterner, R. Significance of the Protein Interface Configuration for Allostery in Imidazole Glycerol Phosphate Synthase. Biochemistry 2020, 59 (29), 27292742,  DOI: 10.1021/acs.biochem.0c00332
    457. 457
      Zubi, Y. S.; Seki, K.; Li, Y.; Hunt, A. C.; Liu, B.; Roux, B.; Jewett, M. C.; Lewis, J. C. Metal-Responsive Regulation of Enzyme Catalysis Using Genetically Encoded Chemical Switches. Nature Commun. 2022, 13 (1), 1864,  DOI: 10.1038/s41467-022-29239-y
    458. 458
      Yang, H.; Swartz, A. M.; Park, H. J.; Srivastava, P.; Ellis-Guardiola, K.; Upp, D. M.; Lee, G.; Belsare, K.; Gu, Y.; Zhang, C. Evolving Artificial Metalloenzymes via Random Mutagenesis. Nat. Chem. 2018, 10 (3), 318324,  DOI: 10.1038/nchem.2927
    459. 459
      Leveson-Gower, R. B.; Zhou, Z.; Drienovská, I.; Roelfes, G. Unlocking Iminium Catalysis in Artificial Enzymes to Create a Friedel-Crafts Alkylase. ACS Catal. 2021, 11 (12), 67636770,  DOI: 10.1021/acscatal.1c00996
    460. 460
      Mayer, C.; Dulson, C.; Reddem, E.; Thunnissen, A.-M. W. H.; Roelfes, G. Directed Evolution of a Designer Enzyme Featuring an Unnatural Catalytic Amino Acid. Angew. Chem. Int. Ed. 2019, 58 (7), 20832087,  DOI: 10.1002/anie.201813499
    461. 461
      Kaes, C.; Katz, A.; Hosseini, M. W. Bipyridine: The Most Widely Used Ligand. A Review of Molecules Comprising at Least Two 2,2‘-Bipyridine Units. Chem. Rev. 2000, 100 (10), 35533590,  DOI: 10.1021/cr990376z
    462. 462
      Xie, J.; Liu, W.; Schultz, P. G. A Genetically Encoded Bidentate, Metal-Binding Amino Acid. Angew. Chem. Int. Ed. 2007, 46 (48), 92399242,  DOI: 10.1002/anie.200703397
    463. 463
      Lee, H. S.; Schultz, P. G. Biosynthesis of a Site-Specific DNA Cleaving Protein. J. Am. Chem. Soc. 2008, 130 (40), 1319413195,  DOI: 10.1021/ja804653f
    464. 464
      Roelfes, G. LmrR: A Privileged Scaffold for Artificial Metalloenzymes. Acc. Chem. Res. 2019, 52 (3), 545556,  DOI: 10.1021/acs.accounts.9b00004
    465. 465
      Drienovská, I.; Rioz-Martínez, A.; Draksharapu, A.; Roelfes, G. Novel Artificial Metalloenzymes by in vivo Incorporation of Metal-Binding Unnatural Amino Acids. Chem. Sci. 2015, 6 (1), 770776,  DOI: 10.1039/C4SC01525H
    466. 466
      Drienovská, I.; Alonso-Cotchico, L.; Vidossich, P.; Lledós, A.; Maréchal, J.-D.; Roelfes, G. Design of an Enantioselective Artificial Metallo-Hydratase Enzyme Containing an Unnatural Metal-Binding Amino Acid. Chem. Sci. 2017, 8 (10), 72287235,  DOI: 10.1039/C7SC03477F
    467. 467
      Ségaud, N.; Drienovská, I.; Chen, J.; Browne, W. R.; Roelfes, G. Artificial Metalloproteins for Binding and Stabilization of a Semiquinone Radical. Inorg. Chem. 2017, 56 (21), 1329313299,  DOI: 10.1021/acs.inorgchem.7b02073
    468. 468
      Bersellini, M.; Roelfes, G. Multidrug Resistance Regulators (MDRs) as Scaffolds for the Design of Artificial Metalloenzymes. Org. Biomol. Chem. 2017, 15 (14), 30693073,  DOI: 10.1039/C7OB00390K
    469. 469
      Jung, S.-M.; Yang, M.; Song, W. J. Symmetry-Adapted Synthesis of Dicopper Oxidases with Divergent Dioxygen Reactivity. Inorg. Chem. 2022, 61 (31), 1243312441,  DOI: 10.1021/acs.inorgchem.2c01898
    470. 470
      Drienovská, I.; Scheele, R. A.; Gutiérrez de Souza, C.; Roelfes, G. A Hydroxyquinoline-Based Unnatural Amino Acid for the Design of Novel Artificial Metalloenzymes. ChemBioChem 2020, 21 (21), 30773081,  DOI: 10.1002/cbic.202000306
    471. 471
      Stein, A.; Liang, A. D.; Sahin, R.; Ward, T. R. Incorporation of Metal-Chelating Unnatural Amino Acids into Halotag for Allylic Deamination. J. Organomet. Chem. 2022, 962, 122272,  DOI: 10.1016/j.jorganchem.2022.122272
    472. 472
      Los, G. V.; Encell, L. P.; McDougall, M. G.; Hartzell, D. D.; Karassina, N.; Zimprich, C.; Wood, M. G.; Learish, R.; Ohana, R. F.; Urh, M. HaloTag: A Novel Protein Labeling Technology for Cell Imaging and Protein Analysis. ACS Chem. Biol. 2008, 3 (6), 373382,  DOI: 10.1021/cb800025k
    473. 473
      Coquière, D.; Bos, J.; Beld, J.; Roelfes, G. Enantioselective Artificial Metalloenzymes Based on a Bovine Pancreatic Polypeptide Scaffold. Angew. Chem. Int. Ed. 2009, 48 (28), 51595162,  DOI: 10.1002/anie.200901134
    474. 474
      Madoori, P. K.; Agustiandari, H.; Driessen, A. J. M.; Thunnissen, A. M. W. H. Structure of the Transcriptional Regulator LmrR and Its Mechanism of Multidrug Recognition. EMBO J. 2009, 28 (2), 156166,  DOI: 10.1038/emboj.2008.263
    475. 475
      Yu, Y.; Hu, C.; Xia, L.; Wang, J. Artificial Metalloenzyme Design with Unnatural Amino Acids and Non-Native Cofactors. ACS Catal. 2018, 8 (3), 18511863,  DOI: 10.1021/acscatal.7b03754
    476. 476
      Yang, H.; Srivastava, P.; Zhang, C.; Lewis, J. C. A General Method for Artificial Metalloenzyme Formation through Strain-Promoted Azide-Alkyne Cycloaddition. ChemBioChem 2014, 15 (2), 223227,  DOI: 10.1002/cbic.201300661
    477. 477
      Srivastava, P.; Yang, H.; Ellis-Guardiola, K.; Lewis, J. C. Engineering a Dirhodium Artificial Metalloenzyme for Selective Olefin Cyclopropanation. Nature Commun. 2015, 6 (1), 7789,  DOI: 10.1038/ncomms8789
    478. 478
      Upp, D. M.; Huang, R.; Li, Y.; Bultman, M. J.; Roux, B.; Lewis, J. C. Engineering Dirhodium Artificial Metalloenzymes for Diazo Coupling Cascade Reactions. Angew. Chem. Int. Ed. 2021, 60 (44), 2367223677,  DOI: 10.1002/anie.202107982
    479. 479
      Ellis-Guardiola, K.; Rui, H.; Beckner, R. L.; Srivastava, P.; Sukumar, N.; Roux, B.; Lewis, J. C. Crystal Structure and Conformational Dynamics of Pyrococcus furiosus Prolyl Oligopeptidase. Biochemistry 2019, 58 (12), 16161626,  DOI: 10.1021/acs.biochem.9b00031
    480. 480
      Brady, L.; Brzozowski, A. M.; Derewenda, Z. S.; Dodson, E.; Dodson, G.; Tolley, S.; Turkenburg, J. P.; Christiansen, L.; Huge-Jensen, B.; Norskov, L. A Serine Protease Triad Forms the Catalytic Centre of a Triacylglycerol Lipase. Nature 1990, 343 (6260), 767770,  DOI: 10.1038/343767a0
    481. 481
      Buller, A. R.; Townsend, C. A. Intrinsic Evolutionary Constraints on Protease Structure, Enzyme Acylation, and the Identity of the Catalytic Triad. Proc. Natl. Acad. Sci. U.S.A. 2013, 110 (8), E653-E661  DOI: 10.1073/pnas.1221050110
    482. 482
      Smith, A. J. T.; Müller, R.; Toscano, M. D.; Kast, P.; Hellinga, H. W.; Hilvert, D.; Houk, K. N. Structural Reorganization and Preorganization in Enzyme Active Sites: Comparisons of Experimental and Theoretically Ideal Active Site Geometries in the Multistep Serine Esterase Reaction Cycle. J. Am. Chem. Soc. 2008, 130 (46), 1536115373,  DOI: 10.1021/ja803213p
    483. 483
      Burton, A. J.; Thomson, A. R.; Dawson, W. M.; Brady, R. L.; Woolfson, D. N. Installing Hydrolytic Activity into a Completely De Novo Protein Framework. Nat. Chem. 2016, 8 (9), 837844,  DOI: 10.1038/nchem.2555
    484. 484
      Rajagopalan, S.; Wang, C.; Yu, K.; Kuzin, A. P.; Richter, F.; Lew, S.; Miklos, A. E.; Matthews, M. L.; Seetharaman, J.; Su, M. Design of Activated Serine-Containing Catalytic Triads with Atomic-Level Accuracy. Nat. Chem. Biol. 2014, 10 (5), 386391,  DOI: 10.1038/nchembio.1498
    485. 485
      Richter, F.; Blomberg, R.; Khare, S. D.; Kiss, G.; Kuzin, A. P.; Smith, A. J. T.; Gallaher, J.; Pianowski, Z.; Helgeson, R. C.; Grjasnow, A. Computational Design of Catalytic Dyads and Oxyanion Holes for Ester Hydrolysis. J. Am. Chem. Soc. 2012, 134 (39), 1619716206,  DOI: 10.1021/ja3037367
    486. 486
      Burke, A. J.; Lovelock, S. L.; Frese, A.; Crawshaw, R.; Ortmayer, M.; Dunstan, M.; Levy, C.; Green, A. P. Design and Evolution of an Enzyme with a Non-Canonical Organocatalytic Mechanism. Nature 2019, 570 (7760), 219223,  DOI: 10.1038/s41586-019-1262-8
    487. 487
      Bjelic, S.; Nivón, L. G.; Çelebi-Ölçüm, N.; Kiss, G.; Rosewall, C. F.; Lovick, H. M.; Ingalls, E. L.; Gallaher, J. L.; Seetharaman, J.; Lew, S. Computational Design of Enone-Binding Proteins with Catalytic Activity for the Morita-Baylis-Hillman Reaction. ACS Chem. Biol. 2013, 8 (4), 749757,  DOI: 10.1021/cb3006227
    488. 488
      Hutton, A. E.; Foster, J.; Crawshaw, R.; Hardy, F. J.; Johannissen, L. O.; Lister, T. M.; Gerard, E. F.; Birch-Price, Z.; Obexer, R.; Hay, S.; Green, A. P. A Non-Canonical Nucleophile Unlocks a New Mechanistic Pathway in a Designed Enzyme. Nat Commun 2024,  DOI: 10.1038/s41467-024-46123-z
    489. 489
      Crawshaw, R.; Crossley, A.; Johannissen, L.; Burke, A.; Hay, S.; Levy, C.; Baker, D.; Lovelock, S.; Green, A. Engineering an Efficient and Enantioselective Enzyme for the Morita-Baylis-Hillman Reaction. Nat. Chem. 2022, 14, 313,  DOI: 10.1038/s41557-021-00833-9
    490. 490
      Agten, S. M.; Dawson, P. E.; Hackeng, T. M. Oxime Conjugation in Protein Chemistry: From Carbonyl Incorporation to Nucleophilic Catalysis. J. Pept. Sci. 2016, 22 (5), 271279,  DOI: 10.1002/psc.2874
    491. 491
      Kölmel, D. K.; Kool, E. T. Oximes and Hydrazones in Bioconjugation: Mechanism and Catalysis. Chem. Rev. 2017, 117 (15), 1035810376,  DOI: 10.1021/acs.chemrev.7b00090
    492. 492
      Drienovská, I.; Mayer, C.; Dulson, C.; Roelfes, G. A Designer Enzyme for Hydrazone and Oxime Formation Featuring an Unnatural Catalytic Aniline Residue. Nat. Chem. 2018, 10 (9), 946952,  DOI: 10.1038/s41557-018-0082-z
    493. 493
      Ofori Atta, L.; Zhou, Z.; Roelfes, G. In vivo Biocatalytic Cascades Featuring an Artificial-Enzyme-Catalysed New-to-Nature Reaction. Angew. Chem. Int. Ed. 2023, 62 (1), e202214191  DOI: 10.1002/anie.202214191
    494. 494
      Leveson-Gower, R. B.; de Boer, R. M.; Roelfes, G. Tandem Friedel-Crafts-Alkylation-Enantioselective-Protonation by Artificial Enzyme Iminium Catalysis. ChemCatChem 2022, 14 (8), e202101875  DOI: 10.1002/cctc.202101875
    495. 495
      Zhou, Z.; Roelfes, G. Synergistic Catalysis in an Artificial Enzyme by Simultaneous Action of Two Abiological Catalytic Sites. Nat. Catal. 2020, 3 (3), 289294,  DOI: 10.1038/s41929-019-0420-6
    496. 496
      Zhou, Z.; Roelfes, G. Synergistic Catalysis of Tandem Michael Addition/Enantioselective Protonation Reactions by an Artificial Enzyme. ACS Catal. 2021, 11 (15), 93669369,  DOI: 10.1021/acscatal.1c02298
    497. 497
      Gran-Scheuch, A.; Bonandi, E.; Drienovská, I. Expanding the Genetic Code: Incorporation of Functional Secondary Amines via Stop Codon Suppression. ChemCatChem 2024, 16, e202301004  DOI: 10.1002/cctc.202301004
    498. 498
      Longwitz, L.; Leveson-Gower, R. B.; Rozeboom, H. J.; Thunnissen, A.-M. W. H.; Roelfes, G. Boron Catalysis in a Designer Enzyme. Nature 2024, 629, 824,  DOI: 10.1038/s41586-024-07391-3
    499. 499
      Garrido-Castro, A. F.; Maestro, M. C.; Alemán, J. Asymmetric Induction in Photocatalysis - Discovering a New Side to Light-Driven Chemistry. Tetrahedron Lett. 2018, 59 (14), 12861294,  DOI: 10.1016/j.tetlet.2018.02.040
    500. 500
      Yao, W.; Bazan-Bergamino, E. A.; Ngai, M.-Y. Asymmetric Photocatalysis Enabled by Chiral Organocatalysts. ChemCatChem 2022, 14 (1), e202101292  DOI: 10.1002/cctc.202101292
    501. 501
      Heyes, D. J.; Lakavath, B.; Hardman, S. J. O.; Sakuma, M.; Hedison, T. M.; Scrutton, N. S. Photochemical Mechanism of Light-Driven Fatty Acid Photodecarboxylase. ACS Catal. 2020, 10 (12), 66916696,  DOI: 10.1021/acscatal.0c01684
    502. 502
      Heyes, D. J.; Zhang, S.; Taylor, A.; Johannissen, L. O.; Hardman, S. J. O.; Hay, S.; Scrutton, N. S. Photocatalysis as the ‘Master Switch’ of Photomorphogenesis in Early Plant Development. Nature Plants 2021, 7 (3), 268276,  DOI: 10.1038/s41477-021-00866-5
    503. 503
      Tan, C.; Liu, Z.; Li, J.; Guo, X.; Wang, L.; Sancar, A.; Zhong, D. The Molecular Origin of High DNA-Repair Efficiency by Photolyase. Nature Commun. 2015, 6 (1), 7302,  DOI: 10.1038/ncomms8302
    504. 504
      Black, M. J.; Biegasiewicz, K. F.; Meichan, A. J.; Oblinsky, D. G.; Kudisch, B.; Scholes, G. D.; Hyster, T. K. Asymmetric Redox-Neutral Radical Cyclization Catalysed by Flavin-Dependent ‘Ene’-Reductases. Nat. Chem. 2020, 12 (1), 7175,  DOI: 10.1038/s41557-019-0370-2
    505. 505
      Emmanuel, M. A.; Greenberg, N. R.; Oblinsky, D. G.; Hyster, T. K. Accessing Non-Natural Reactivity by Irradiating Nicotinamide-Dependent Enzymes with Light. Nature 2016, 540 (7633), 414417,  DOI: 10.1038/nature20569
    506. 506
      Liu, X.; Kang, F.; Hu, C.; Wang, L.; Xu, Z.; Zheng, D.; Gong, W.; Lu, Y.; Ma, Y.; Wang, J. A Genetically Encoded Photosensitizer Protein Facilitates the Rational Design of a Miniature Photocatalytic CO2-Reducing Enzyme. Nat. Chem. 2018, 10 (12), 12011206,  DOI: 10.1038/s41557-018-0150-4
    507. 507
      Siegel, J. B.; Zanghellini, A.; Lovick, H. M.; Kiss, G.; Lambert, A. R.; St. Clair, J. L.; Gallaher, J. L.; Hilvert, D.; Gelb, M. H.; Stoddard, B. L.; Houk, K. N.; Michael, F. E.; Baker, D. Computational Design of an Enzyme Catalyst for a Stereoselective Bimolecular Diels-Alder Reaction. Science 2010, 329, 309313,  DOI: 10.1126/science.1190239
    508. 508
      Sun, N.; Huang, J.; Qian, J.; Zhou, T.-P.; Guo, J.; Tang, L.; Zhang, W.; Deng, Y.; Zhao, W.; Wu, G. Enantioselective [2 + 2]-Cycloadditions with Triplet Photoenzymes. Nature 2022, 611 (7937), 715720,  DOI: 10.1038/s41586-022-05342-4
    509. 509
      Allen, A. R.; Noten, E. A.; Stephenson, C. R. J. Aryl Transfer Strategies Mediated by Photoinduced Electron Transfer. Chem. Rev. 2022, 122 (2), 26952751,  DOI: 10.1021/acs.chemrev.1c00388
    510. 510
      Dadashi-Silab, S.; Doran, S.; Yagci, Y. Photoinduced Electron Transfer Reactions for Macromolecular Syntheses. Chem. Rev. 2016, 116 (17), 1021210275,  DOI: 10.1021/acs.chemrev.5b00586
    511. 511
      Reid, B. G.; Flynn, G. C. Chromophore Formation in Green Fluorescent Protein. Biochemistry 1997, 36 (22), 67866791,  DOI: 10.1021/bi970281w
    512. 512
      Fu, Y.; Huang, J.; Wu, Y.; Liu, X.; Zhong, F.; Wang, J. Biocatalytic Cross-Coupling of Aryl Halides with a Genetically Engineered Photosensitizer Artificial Dehalogenase. J. Am. Chem. Soc. 2021, 143 (2), 617622,  DOI: 10.1021/jacs.0c10882
    513. 513
      Gu, Y.; Ellis-Guardiola, K.; Srivastava, P.; Lewis, J. C. Preparation, Characterization, and Oxygenase Activity of a Photocatalytic Artificial Enzyme. ChemBioChem 2015, 16 (13), 18801883,  DOI: 10.1002/cbic.201500165
    514. 514
      Zubi, Y. S.; Liu, B.; Gu, Y.; Sahoo, D.; Lewis, J. C. Controlling the Optical and Catalytic Properties of Artificial Metalloenzyme Photocatalysts Using Chemogenetic Engineering. Chem. Sci. 2022, 13 (5), 14591468,  DOI: 10.1039/D1SC05792H
    515. 515
      Liu, B.; Zubi, Y. S.; Lewis, J. C. Iridium(iii) Polypyridine Artificial Metalloenzymes with Tunable Photophysical Properties: A New Platform for Visible Light Photocatalysis in Aqueous Solution. Dalton Trans. 2023, 52 (16), 50345038,  DOI: 10.1039/D3DT00932G
    516. 516
      Lee, J.; Song, W. J. Photocatalytic C-O Coupling Enzymes That Operate via Intramolecular Electron Transfer. J. Am. Chem. Soc. 2023, 145 (9), 52115221,  DOI: 10.1021/jacs.2c12226
    517. 517
      Mills, J. H.; Sheffler, W.; Ener, M. E.; Almhjell, P. J.; Oberdorfer, G.; Pereira, J. H.; Parmeggiani, F.; Sankaran, B.; Zwart, P. H.; Baker, D. Computational Design of a Homotrimeric Metalloprotein with a Trisbipyridyl Core. Proc. Natl. Acad. Sci. U.S.A. 2016, 113 (52), 1501215017,  DOI: 10.1073/pnas.1600188113
    518. 518
      Duan, H.-Z.; Hu, C.; Li, Y.-L.; Wang, S.-H.; Xia, Y.; Liu, X.; Wang, J.; Chen, Y.-X. Genetically Encoded Phosphine Ligand for Metalloprotein Design. J. Am. Chem. Soc. 2022, 144 (50), 2283122837,  DOI: 10.1021/jacs.2c09683
    519. 519
      Beattie, A. T.; Dunkelmann, D. L.; Chin, J. W. Quintuply Orthogonal Pyrrolysyl-tRNA Synthetase/tRNAPyl Pairs. Nat. Chem. 2023, 15 (7), 948959,  DOI: 10.1038/s41557-023-01232-y
    520. 520
      Dunkelmann, D. L.; Willis, J. C. W.; Beattie, A. T.; Chin, J. W. Engineered Triply Orthogonal Pyrrolysyl-tRNA Synthetase/tRNA Pairs Enable the Genetic Encoding of Three Distinct Non-Canonical Amino Acids. Nat. Chem. 2020, 12 (6), 535544,  DOI: 10.1038/s41557-020-0472-x
    521. 521
      Italia, J. S.; Addy, P. S.; Erickson, S. B.; Peeler, J. C.; Weerapana, E.; Chatterjee, A. Mutually Orthogonal Nonsense-Suppression Systems and Conjugation Chemistries for Precise Protein Labeling at up to Three Distinct Sites. J. Am. Chem. Soc. 2019, 141 (15), 62046212,  DOI: 10.1021/jacs.8b12954
    522. 522
      Hashimoto, K.; Fischer, E. C.; Romesberg, F. E. Efforts toward Further Integration of an Unnatural Base Pair into the Biology of a Semisynthetic Organism. J. Am. Chem. Soc. 2021, 143 (23), 86038607,  DOI: 10.1021/jacs.1c03860
    523. 523
      Ledbetter, M. P.; Karadeema, R. J.; Romesberg, F. E. Reprograming the Replisome of a Semisynthetic Organism for the Expansion of the Genetic Alphabet. J. Am. Chem. Soc. 2018, 140 (2), 758765,  DOI: 10.1021/jacs.7b11488
    524. 524
      Stucki, A.; Vallapurackal, J.; Ward, T. R.; Dittrich, P. S. Droplet Microfluidics and Directed Evolution of Enzymes: An Intertwined Journey. Angew. Chem. Int. Ed. 2021, 60 (46), 2436824387,  DOI: 10.1002/anie.202016154
    525. 525
      Wicky, B. I. M.; Milles, L. F.; Courbet, A.; Ragotte, R. J.; Dauparas, J.; Kinfu, E.; Tipps, S.; Kibler, R. D.; Baek, M.; DiMaio, F. Hallucinating Symmetric Protein Assemblies. Science 2022, 378 (6615), 5661,  DOI: 10.1126/science.add1964
    526. 526
      Wang, J.; Lisanza, S.; Juergens, D.; Tischer, D.; Watson, J. L.; Castro, K. M.; Ragotte, R.; Saragovi, A.; Milles, L. F.; Baek, M. Scaffolding Protein Functional Sites Using Deep Learning. Science 2022, 377 (6604), 387394,  DOI: 10.1126/science.abn2100
    527. 527
      Krishna, R.; Wang, J.; Ahern, W.; Sturmfels, P.; Venkatesh, P.; Kalvet, I.; Lee, G. R.; Morey-Burros, F. S.; Anishchenko, I.; Humphreys, I. R. Generalized Biomolecular Modeling and Design with RoseTTAFold All-Atom. Science 2024,  DOI: 10.1126/science.adl2528
    528. 528
      Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A. Highly Accurate Protein Structure Prediction with AlphaFold. Nature 2021, 596 (7873), 583589,  DOI: 10.1038/s41586-021-03819-2
    529. 529
      Baumann, T.; Hauf, M.; Richter, F.; Albers, S.; Möglich, A.; Ignatova, Z.; Budisa, N. Computational Aminoacyl-tRNA Synthetase Library Design for Photocaged Tyrosine. Int. J. Mol. Sci. 2019, 20 (9), 2343,  DOI: 10.3390/ijms20092343
    530. 530
      Cervettini, D.; Tang, S.; Fried, S. D.; Willis, J. C. W.; Funke, L. F. H.; Colwell, L. J.; Chin, J. W. Rapid Discovery and Evolution of Orthogonal Aminoacyl-tRNA Synthetase-tRNA Pairs. Nat. Biotechnol. 2020, 38 (8), 989999,  DOI: 10.1038/s41587-020-0479-2
    531. 531
      Taylor, C. J.; Hardy, F. J.; Burke, A. J.; Bednar, R. M.; Mehl, R. A.; Green, A. P.; Lovelock, S. L. Engineering Mutually Orthogonal PylRS/tRNA Pairs for Dual Encoding of Functional Histidine Analogues. Protein Sci. 2023, 32 (5), e4640  DOI: 10.1002/pro.4640
    532. 532
      Fredens, J.; Wang, K.; de la Torre, D.; Funke, L. F. H.; Robertson, W. E.; Christova, Y.; Chia, T.; Schmied, W. H.; Dunkelmann, D. L.; Beránek, V. Total Synthesis of Escherichia coli with a Recoded Genome. Nature 2019, 569 (7757), 514518,  DOI: 10.1038/s41586-019-1192-5
    533. 533
      Lajoie, M. J.; Rovner, A. J.; Goodman, D. B.; Aerni, H.-R.; Haimovich, A. D.; Kuznetsov, G.; Mercer, J. A.; Wang, H. H.; Carr, P. A.; Mosberg, J. A. Genomically Recoded Organisms Expand Biological Functions. Science 2013, 342 (6156), 357360,  DOI: 10.1126/science.1241459
    534. 534
      Mukai, T.; Hoshi, H.; Ohtake, K.; Takahashi, M.; Yamaguchi, A.; Hayashi, A.; Yokoyama, S.; Sakamoto, K. Highly Reproductive Escherichia coli Cells with No Specific Assignment to the UAG Codon. Sci. Rep. 2015, 5 (1), 9699,  DOI: 10.1038/srep09699
    535. 535
      Wang, T.; Liang, C.; An, Y.; Xiao, S.; Xu, H.; Zheng, M.; Liu, L.; Wang, G.; Nie, L. Engineering the Translational Machinery for Biotechnology Applications. Mol. Biotechnol. 2020, 62 (4), 219227,  DOI: 10.1007/s12033-020-00246-y
    536. 536
      Zackin, M. T.; Stieglitz, J. T.; Van Deventer, J. A. Genome-Wide Screen for Enhanced Noncanonical Amino Acid Incorporation in Yeast. ACS Synth. Biol. 2022, 11 (11), 36693680,  DOI: 10.1021/acssynbio.2c00267
    537. 537
      Exner, M. P.; Kuenzl, T.; To, T. M. T.; Ouyang, Z.; Schwagerus, S.; Hoesl, M. G.; Hackenberger, C. P. R.; Lensen, M. C.; Panke, S.; Budisa, N. Design of S-Allylcysteine in Situ Production and Incorporation Based on a Novel Pyrrolysyl-tRNA Synthetase Variant. ChemBioChem 2017, 18 (1), 8590,  DOI: 10.1002/cbic.201600537
    538. 538
      Kim, S.; Sung, B. H.; Kim, S. C.; Lee, H. S. Genetic Incorporation of L-Dihydroxyphenylalanine (DOPA) Biosynthesized by a Tyrosine Phenol-Lyase. Chem. Commun. 2018, 54 (24), 30023005,  DOI: 10.1039/C8CC00281A
    539. 539
      Nojoumi, S.; Ma, Y.; Schwagerus, S.; Hackenberger, C. P. R.; Budisa, N. In-Cell Synthesis of Bioorthogonal Alkene Tag S-Allyl-Homocysteine and Its Coupling with Reprogrammed Translation. Int. J. Mol. Sci. 2019, 20 (9), 2299,  DOI: 10.3390/ijms20092299