Next Article in Journal
Deconstructive and Divergent Synthesis of Bioactive Natural Products
Next Article in Special Issue
Synthesis of Spirocyclopropane-Containing 4H-Pyrazolo[1,5-a]indoles via Alkylative Dearomatization and Intramolecular N-Imination of an Indole–O-(Methylsulfonyl)oxime
Previous Article in Journal
Synthesis, Reactivity and Coordination Chemistry of Group 9 PBP Boryl Pincer Complexes: [(PBP)M(PMe3)n] (M = Co, Rh, Ir; n = 1, 2)
Previous Article in Special Issue
The Last Decade of Optically Active α-Aminophosphonates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tinker, Tailor, Soldier, Spy: The Diverse Roles That Fluorine Can Play within Amino Acid Side Chains

School of Chemistry, The University of New South Wales (UNSW), Sydney 2052, Australia
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(17), 6192; https://doi.org/10.3390/molecules28176192
Submission received: 28 July 2023 / Revised: 17 August 2023 / Accepted: 17 August 2023 / Published: 22 August 2023
(This article belongs to the Special Issue Feature Papers in Organic Chemistry (Volume II))

Abstract

:
Side chain-fluorinated amino acids are useful tools in medicinal chemistry and protein science. In this review, we outline some general strategies for incorporating fluorine atom(s) into amino acid side chains and for elaborating such building blocks into more complex fluorinated peptides and proteins. We then describe the diverse benefits that fluorine can offer when located within amino acid side chains, including enabling 19F NMR and 18F PET imaging applications, enhancing pharmacokinetic properties, controlling molecular conformation, and optimizing target-binding.

1. Introduction

Fluorination has proven to be an exceptionally useful strategy in the development of small-molecule drugs and agrochemicals [1,2,3,4,5,6,7,8,9,10]. The presence of fluorine can confer a variety of advantages including enhanced resistance to metabolism, higher membrane permeability, more potent target-binding, and greater target selectivity.
While the value of fluorine is now well-established in the context of small-molecule drugs, a major current trend in the pharmaceutical industry is towards “beyond rule of 5” compounds, i.e., bioactive compounds that lie outside of the physicochemical parameters that are commonly accepted to correlate with oral bioavailability [11,12,13,14,15,16,17,18]. Particularly notable amongst this new generation of pharmaceutical agents are peptides and proteins.
Given the track record of fluorination in the context of small-molecule drugs, it seems likely that fluorination could offer significant benefits in the optimization of peptide- and protein-based drugs too.
The structure of a peptide offers several possible sites for fluorination, which can be broadly categorized as on the backbone or on the side chain. In terms of the peptide backbone, fluorine can be found within non-hydrolyzable amide isosteres (e.g., CF=CH; C(CF3)=CH; C(CF3)–NH), or partway along a backbone-extended amino acid (e.g., fluorostatines such as H2N–CH(iBu)–CH(OH)–CF2–CO2H). We have recently reviewed some of these aspects of backbone fluorination [19].
In the present review, we focus on side chain fluorination (Figure 1). We will briefly discuss the various strategies for synthesizing side chain-fluorinated amino acids (Section 2), and then we will delve into the varied roles that fluorine can play within amino acid side chains, including enabling NMR and PET imaging applications (Section 3); remediating problematic pharmacokinetic properties (Section 4); controlling conformation on scales ranging from individual amino acids all the way up to protein quaternary structure (Section 5); and, finally, enhancing target-binding interactions (Section 6). There have been several excellent reviews of some of these topics [20,21,22,23,24,25,26,27,28] but we feel that it is worthwhile to provide an updated and broad account of the field.

2. Synthetic Aspects

Mirroring the structural diversity of side chain-fluorinated amino acids is a diversity of methods available for their chemical synthesis. Several reviews have already been published on this topic [29,30,31,32,33,34,35,36,37,38], so we will provide only a very brief overview here (Section 2.1). Additionally, we will provide a concise discussion of methods for elaborating side chain-fluorinated amino acids into peptides and proteins (Section 2.2).

2.1. Strategies for the Synthesis of Side Chain-Fluorinated Amino Acids

One way to efficiently obtain a side chain-fluorinated amino acid is to commence with a simple commercially available fluorinated building block [39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57], and then append the required amino and carboxylate groups using one of several established methods. The Schöllkopf approach is a typical example (Scheme 1). By this means, the simple building block 2,3,4-6-tetrafluorobenzyl bromide (1) can be converted, stereoselectively, into 2,3,4,6-tetrafluorophenylalanine (3) [34,58,59].
Biochemical methods can also be utilized for transforming simple fluorinated building blocks into side chain-fluorinated amino acids [60,61,62]. For example, the directed evolution of tryptophan synthase β-subunit (TrpB) from Pyrococcus furiosus generated a mutant enzyme that could efficiently convert 5-fluoroindole (5) into the fluorinated tryptophan analog 6 [60].
A range of methods are available for forming the C–F bond at a later stage of the amino acid synthesis. Such fluorination methods can be broadly categorized according to the mechanism involved, one of which is nucleophilic fluorination [63,64,65,66,67,68,69,70,71,72]. A nucleophilic source of fluoride such as diethylaminosulfur trifluoride (DAST), morpholinosulfur trifluoride (morph-DAST), or silver fluoride (AgF) can be employed to displace a leaving group and deliver a side chain-fluorinated amino acid product. A typical example is shown in Scheme 1: treatment of the secondary alcohol 7 with Deoxo-Fluor (8) affected substitution with inversion to deliver the fluorinated β-amino ester 10 [67]. However, the modest yield of this particular transformation highlights a limitation of the nucleophilic fluorination approach more generally, which is that it can sometimes be outcompeted by side reactions such as elimination or rearrangement.
Fluorine can alternatively be incorporated into amino acid side chains via electrophilic fluorination. Electrophilic sources of fluorine such as N-fluorobenzensulfonimide (NFSI) or Selectfluor (14) can be employed to react with electron-rich amino acid side chains and deliver fluorinated targets [73,74,75,76,77,78]. For example, the electron-rich tyrosine side chain (11) can undergo an SEAr reaction with acetyl hypofluorite to deliver the fluorinated tyrosine derivative 12 in a reasonable yield (Scheme 1) [79]. This transformation is notable for its chemoselectivity: no reaction at the less electron-rich phenylalanine side chain of 11 is observed.
Another method for the synthesis of side chain-fluorinated amino acids is metal-catalyzed fluorination [80,81]. For example, palladium catalysis has been applied for the site-selective fluorination of a non-functionalized sp3 carbon to produce the fluorinated α-amino acid derivative 16 (Scheme 1) [82]. This reaction was aided by the directing group PIP (2-(pyridin-2-yl)isopropyl amine) that facilitated metal coordination.
Yet another general strategy for synthesizing side chain-fluorinated amino acids is photocatalytic/radical fluorination [83,84]. Using catalytic dibenzosuberenone (18) and Selectfluor (14) as the fluorine source, a visible-light-promoted fluorination of an sp3-hybridized C-H bond was realized (Scheme 1). The method was optimized for benzylic protons on phenylalanine-like residues and was shown to be efficient even for functionalizing short peptides such as 17 [84].

2.2. Elaboration of Side Chain-Fluorinated Amino Acids into Peptides and Proteins

Solid-phase peptide synthesis (SPPS) is one of the most commonly used methods to obtain synthetic peptides containing a side chain-fluorinated residue [85,86,87,88,89,90,91,92,93]. This technique relies on the use of a solid support or resin onto which the desired peptide is assembled. The synthesis is typically performed by initially attaching the C-terminal amino acid to the resin. The succeeding amino acids from the target sequence are then individually attached to this anchored residue in a stepwise manner using an appropriate coupling reagent. For instance, SPPS using a leucinol (Lol)-substituted trityl-chloride resin was utilized in introducing (4-fluorophenyl)alanine into an analog of the lipopeptaibiotic trichogin GA IV (21, Scheme 2) in order to understand its conformation through 19F-NMR studies [88].
In addition to chemical approaches, several biosynthetic pathways for amino acid incorporation into peptides and proteins are available. Precursor-directed biosynthesis operates by administering the fluorinated amino acid to a culture of the organism that produces the desired peptide or protein [94,95]. For example, iturins and fengycins are lipopeptides naturally produced by Bacillus sp. CS93, which has been shown to exhibit antifungal properties. Feeding the bacterial cultures with fluorinated tyrosine (22) resulted in the production of novel fluorinated counterparts of the lipopeptides (e.g., 23, Scheme 2) [95].
In some cases, biosynthetic incorporation of a fluorinated amino acid into the proteome is actually the mechanism of action of a therapeutic agent. This approach has been studied as a tool for inhibiting the growth of certain bacterial species [96,97,98]. In in vitro cases, fluorinated amino acids fed into bacterial cultures were found to have become misincorporated into the bacterial proteome, ultimately inducing toxicity via inhibition of cell growth [97,98]. A study on the toxicity of p-fluorophenylalanine (p-FPA) on Escherichia coli 15T revealed that when administered to the culture simultaneously with thymine, p-FPA induced thymine starvation and led to decreased RNA/DNA synthesis, leading to cell cycle arrest [99].
We return now to the situation where a high synthetic yield of a particular fluorinated protein is desired. The rate of incorporation of the fluorinated amino acid can be enhanced by making use of auxotrophic strains that are unable to synthesize a specific amino acid. Typically, the growth medium for the auxotrophic strain is deprived of that amino acid and replaced by a surrogate fluorinated analog, resulting in the incorporation of the latter into synthesized proteins in lieu of its natural amino acid counterpart [100,101,102]. This approach was used in the synthesis of a mutant basic leucine zipper (bzip) peptide [102]. Using an auxotrophic E. coli strain BL21(DE3), 5,5,5-trifluorolecine and 4,4,4-trifluorovaline were successfully incorporated as isoleucine surrogates.
The amber stop codon (UAG) can be exploited for a similar purpose. This nonsense codon can be included at a specific position of the mRNA where a desired fluorinated amino acid is to be introduced and can therefore be used for site-specific fluorination of a peptide [103,104,105,106]. Using a suppressor tRNA/aminoacyl-tRNA synthetase pair (tRNAPylCUA/MmFAcKRS1) derived from Methanosarcina mazei, Nε-fluoroacetyllysine (FAcK) was successfully incorporated into the Zspa Affibody (Afb) protein at the position of the amber codon [106].
Finally, display technologies have been explored for the elaboration of side chain-fluorinated amino acids into peptides. One such technique is mRNA display, a tool for directed evolution wherein peptides with a desired trait are generated through iterative cycles of diversification and selection (Scheme 2) [107,108,109,110]. The cycle is initiated by transcription of a DNA library to the corresponding mRNA, followed by ligation to puromycin at the 3′ end. Translation of the puromycin-linked mRNA produces mRNA-tagged peptides, which undergo cyclization and reverse transcription into corresponding peptide-mRNA-cDNA fusions. The desired fusions are then selected by binding to a bead-immobilized target. The cDNA from selected fusions subsequently undergoes error-prone PCR to promote amplification and generate the library for the next selection cycle [111,112]. It is possible for the DNA library to be expanded to accommodate unnatural amino acids such as side chain-fluorinated amino acids [111,113]. This approach led to the discovery of cyclic peptide 24 (Scheme 2), containing three side chain-fluorinated amino acid residues for inhibition of proprotein convertase subtilisin-like/kexin type 9 (PCSK 9), which is a valuable target for the treatment of coronary heart disease [107].

3. Fluorine, the “Spy”: Transmitting Intelligence on the Properties of Amino Acids, Peptides, and Proteins

Sometimes, when fluorine is introduced into an amino acid side chain, it does not dramatically alter the molecular properties. In such cases, the fluorine might be viewed as an “innocent bystander” or perhaps as a “spy”: it offers the opportunity to gather intelligence about the molecule’s properties through analytical techniques such as 19F NMR spectrometry (Section 3.1) or positron emission tomography (Section 3.2).

3.1. 19F-Containing Amino Acids as NMR Tags

The introduction of fluorine within a protein provides the opportunity to interrogate the properties and functions of the biomacromolecule through 19F NMR spectrometry. This is a longstanding concept that has been the subject of several recent reviews [114,115,116,117,118]. The process begins with the synthesis of a non-natural analog of the protein in which one or more residues are replaced with a side chain-fluorinated amino acid. Next, the 19F NMR spectrum of this non-natural protein is recorded as a point of reference, assuming that the presence of fluorine does not dramatically alter the structure compared to the native protein. Starting from this reference point, any subsequent changes in the 19F NMR spectrum of the protein can be used to detect, e.g., a conformational change in the protein, a ligand binding event, or some other supramolecular interaction of the protein.
There are several reasons why fluorine is especially advantageous as an NMR tag. The 19F nucleus has 100% isotopic abundance and high NMR sensitivity (83% compared to 1H). Since fluorine is not naturally present in any biomacromolecules, there are no background signals even if the protein of interest is present within a complex biological milieu. Finally, 19F NMR signals can appear over a very wide chemical shift range (>500 ppm) and they are extremely sensitive to their environment. Taken together, these two features mean that signal overlap is rare even if multiple copies of the same fluorinated amino acid residue are present at different positions in the protein sequence. They also mean that even very subtle changes to a protein’s structure can cause easily detectable perturbations to the 19F NMR spectrum.
One way to incorporate fluorine as an NMR tag into a protein is to employ a “prosthetic group” approach [119,120,121,122]. Reagents such as p-fluorobenzenesulfonyl chloride or 2,2,2,-trifluoroethanethiol can undergo reaction with solvent-exposed lysine or cysteine side chains, respectively, within an intact protein, delivering a fluorine-labelled structure (e.g., 2526, Figure 2a). In a more elaborate example, the enzyme transglutaminase was recently shown to accept 2,2,2,-trifluoroethylamine as a substrate, which allowed it to be used as a bioconjugation reagent for the labelling of a glutamate residue on the surface of a model protein (27) [123].
A complementary set of NMR tags include the monofluorinated amino acids 2829 and 22 (Figure 2a). Such structures can be considered to be more sophisticated than the prosthetic group examples in the sense that they more closely mimic natural amino acid side chains [88,92,124,125,126,127,128,129,130,131,132,133] and can be installed anywhere in the protein sequence (see Section 2.2), not just at solvent-exposed regions. However, a limitation of 2829 and 22 is that they contain just one fluorine atom, which limits the NMR sensitivity. More recently, polyfluorinated amino acids such as 3032 (Figure 2a) have attracted interest because their intense 19F NMR signals potentially allow lower concentrations of the protein analyte to be employed [134].
Fluorinated prolines (e.g., 3336, Figure 2a) are an interesting subset of NMR tags because they illustrate the importance of identifying a conformational match with the native amino acid [89,116,135,136]. 4-Fluoroprolines (e.g., 33) and 3-fluoroprolines (e.g., 34) adopt different puckers of the 5-membered ring (see also Section 5.1) and different cis/trans ratios of the peptide bond to the N-terminal side of the proline residue. In these cases, the fluorine is no longer an “innocent bystander”. However, if two fluorine substituents are installed at the 3- and 4-positions, with appropriate stereochemistry (e.g., 35), then the conformational influences of the two fluorines offset one another, resulting in a fluorinated analog that has very similar conformational characteristics to natural proline [137]. Another fluorinated proline derivative that has been found to closely mimic the conformational characteristics of proline itself, while providing an intense 19F NMR signal, is the trifluoromethyl-containing analog 37 (Figure 2) [89].
Having seen some examples of amino acids that bear 19F NMR labels (Figure 2a), let us turn our attention to the application of these building blocks and the study of protein structure and function.
The conformational dynamics of proteins, including their folding and unfolding processes, are fundamental aspects of biology [118,130]. A medicinally relevant case is the misfolding of proteins into amyloid fibrils, which is the basis of diseases such as Alzheimer’s disease and Creuzfeldt–Jakob disease. Mammalian prion protein (PrP) is a predominantly α-helical protein that is associated with neuronal cell membranes. Under certain circumstances, PrP can misfold into a β-sheet rich structure. The misfolded structure catalyzes the misfolding of further molecules of PrP, and this autocatalytic process leads to the accumulation of aggregates called amyloid fibrils, which can cause neuronal cell death and disease pathology. To study the pathway of amyloid formation, a protein-observed 19F NMR study was undertaken [138]. Three 3-fluorophenylalanine (29) residues were introduced into the protein to replace the Phe141, Phe175, and Phe198 of PrP (38, Figure 2b). The 19F NMR spectrum of the fluorinated protein (38) revealed the presence of several oligomeric species, with the predominant constituent being assigned as an octamer. Variable-temperature 19F NMR experiments then allowed the equilibrium distributions to be perturbed and the thermodynamic driving forces of aggregation to be elucidated. The protein-observed 19F NMR approach was also able to explain how certain mutations make the PrP protein more prone to aggregation through the stabilization of the octamer [138].
Another application of 19F-labelled proteins is to discover and optimize small molecule ligands. An example of this approach is seen with the protein known as BPTF, or “bromodomain and plant homeodomain-containing transcription factor”. This protein is involved in the regulation of chromatin accessibility, and its overexpression is associated with lung cancer. A 5-fluorotryptophan residue (28) was incorporated at the binding surface of BPTF (39, Figure 2b), which enabled a medium-throughput screen of ~200 potential ligands to be performed, using 19F NMR spectrometry as the detection technique [117]. This screen resulted in a promising hit molecule, the subsequent structure-activity optimization of which was also facilitated by protein-observed 19F NMR.
Protein-observed 19F NMR can be used to study the interactions of proteins with other biomacromolecular structures such as membranes. When investigating the membrane interactions of helical antimicrobial peptides by solid-state NMR, it is helpful to have a fluorine tag that is held in a fixed orientation relative to the helical axis [85,139,140,141,142,143]. The unusual amino acid 37 (Figure 2a), which contains a rigid bicyclo[1.1.1]pentane moiety, meets this requirement [144,145].
19F NMR spectrometry can be applied not only to interrogate the properties of a fluorinated molecule as described above, but alternatively to visualize where a fluorinated molecule travels within the body (i.e., 19F magnetic resonance imaging, or 19F-MRI). For example, 6-fluoro-DOPA, a ring-fluorinated analog of dihydroxyphenylalanine, has been employed as a brain imaging agent in a rat model of Parkinson’s disease [146]. A multi-fluorinated DOPA analog has also been developed in order to achieve a stronger signal for 19F-MRI applications [147]. However, the use of 19F NMR spectrometry as an imaging modality remains quite niche [148], particularly in comparison with PET, which is discussed in the next section.

3.2. 18F-Labelled Amino Acids and Peptides as PET Tracers

18F-Radiolabelled amino acids can be valuable agents for the diagnosis and visualization of cancer. Tumour cells have high biosynthetic demand [149], and one way that they can secure an increased supply of biosynthetic building blocks is by upregulating amino acid transporter proteins [150,151]. Thus, 18F-radiolabelled amino acids often selectively accumulate in tumours, where their presence can be detected by positron emission tomography.
One of the most operationally straightforward ways to incorporate an 18F radiolabel onto the side chain of an amino acid is to install a prosthetic group that contains 18F (Figure 3a). For example, the tyrosine side chain contains a phenol moiety, which can be alkylated with [18F]fluoroethyltosylate to provide 2-[18F]fluoroethyltyrosine (40). This important radiotracer is particularly valuable for the imaging of brain cancers as the amino acid transporter proteins mentioned above enable this tracer to efficiently cross the blood–brain barrier [150,152,153,154]. Prosthetic groups for the radiolabelling of several other amino acids besides tyrosine [155,156,157] have also been developed, taking advantage of the reactivity of the side chains of serine (41) [158], threonine (42) [158], tryptophan (43) [159], cysteine (44) [160,161], and ornithine (45) [151,162] (Figure 3a). Certain more elaborate prosthetic groups afford the alternative possibility of late-stage fluorination through 18F/19F isotopic exchange (e.g., 4647) [163,164].
A potential disadvantage of the prosthetic group approach is that the structure of the labelled amino acid has become rather different from the natural amino acid, so the in vivo distribution might also differ. To overcome this issue, it is sometimes desirable to attach the 18F atom directly to the amino acid side chain as a replacement for a C–H hydrogen (Figure 3a). Most commonly, a synthetic precursor bearing a leaving group is required to enable an SNAr [65,165,166,167,168,169,170] or SN2 reaction [171,172,173,174,175,176,177,178,179,180,181] to take place to install the 18F substituent (e.g., 4849). Some other reaction manifolds have also been exploited for radiofluorination of amino acid side chains or precursors thereof, including direct C–H fluorination (50) [182,183,184] and organocatalytic electrophilic fluorination (51) [76].
More complex structures have also been created in which the 18F-radiolabel is attached to an amino acid side chain within a peptide or protein architecture. This opens up the potential to visualize a variety of different disease states, depending on the biomacromolecule that is radiolabelled. There are broadly two strategies for synthesizing a peptide or protein in which one amino acid side chain bears an 18F-radiolabel: early-stage vs. late-stage fluorination.
The early-stage radiofluorination strategy commences with an 18F-labelled amino acid and then elaborates it into a peptide or protein (Figure 3b). Examples of this strategy are quite rare due to the challenge of synthesizing an entire peptide or protein on a short timescale, but the challenge can be met by leveraging biosynthetic machinery for the elaboration task. For example, 2-[18F]fluoroethyl tyrosine (40) was elaborated via a cell-free translation system into a small protein (or “affibody”) that binds to the human epidermal growth factor receptor (52, Figure 3b) [185]. This biosynthesis afforded a 6.5% overall radiochemical yield.
The second strategy for 18F-radiolabelling a peptide or protein is to perform a late-stage derivatization of one amino acid side chain within the biomacromolecule (Figure 3b). Lysine and cysteine are the most commonly targeted amino acids for this purpose [86,90,91,186,187,188,189,190,191]. For example, a lysine side chain within the 34-residue parathyroid hormone (53) was derivatized as the p-[18F]fluorobenzoyl amide, generating a macromolecular radiotracer (55) suitable for the study of osteoporosis [189]. In another example, a cysteine side chain within the 36 kDa protein annexin V was derivatized with 18F via a maleimide adduct, generating a macromolecular radiotracer capable of detecting apoptotic cells [190].

4. Fluorine, the “Tinker”: Improving the Pharmacokinetic Properties of Amino Acids, Peptides, and Proteins

The sub-optimal pharmacokinetic properties of peptides are one of the key obstacles to their development into viable drugs [15,17]. There is some evidence that fluorination of amino acid side chains can help to improve the hydrophobicity, permeability, and/or stability of the metabolism of certain amino acids and peptides [192]. Selected examples are presented below.

4.1. Hydrophobicity and Permeability

The bicyclic amino acid 56 (Figure 4) is a potent agonist of the metabotropic glutamate receptor, and it shows promise for the treatment of a variety of central nervous system disorders including schizophrenia. Compound 56 suffers from poor oral bioavailability, but this limitation is impressively overcome in the fluorinated analog 57, which replicates the agonist activity of 56 in vitro while being far more efficacious in vivo (Figure 4) [193]. It is unclear whether the improved oral bioavailability of 57 is attributable to increased hydrophobicity, to greater resistance to metabolism, or to some other effect.
Mephalan (57, Figure 4) is a DNA-targeted anticancer drug featuring a nitrogen mustard moiety located on the side chain of phenylalanine. A major limitation of compound 58 is its inability to traverse cell membranes. This limitation can be overcome through a prodrug approach, in which mephalan is temporarily masked as a di- or a tripeptide (e.g., 5960, Figure 4) [194,195]. The presence of the fluorinated amino acid in 5960 enhances the drugs’ membrane permeability, and thereby substantially boosts the anticancer potency in each case.

4.2. Stability towards Proteolysis

When it comes to the issue of proteolytic stability, much work has been done in terms of fluorination of the peptide backbone (e.g., peptidomimetics containing fluoroalkenes or –C(CF3)=CH– groups as non-hydrolyzable isosteres of the amide bond). By contrast, fluorinating peptide side chains is a less intuitive strategy for imparting proteolytic stability because in such structures the potentially hydrolyzable amide functional group is retained; essentially the hope is that a fluorinated side chain would be incompatible with the corresponding binding pocket within a protease enzyme’s active site, preventing its hydrolytic action. There is some evidence that this strategy can indeed impart proteolytic stability to peptides, but only in particular cases, not as a general trend [196,197,198,199].
The α-helical peptide magainin (61, Figure 5) exhibits antimicrobial activity through its ability to assemble into toroidal pores within the bacterial cell membrane. However, 61 contains several trypsin cleavage sites, which leads to a short half-life in vivo and limits the usefulness of 61 as a pharmaceutical agent. The fluorinated analogs 62 and 63 (Figure 5), which contain two or five hexafluoroleucine residues positioned along the hydrophobic face of the helix, respectively, show either a modest or a dramatic increase in proteolytic stability, due to steric incompatibility of the fluorinated side chains with the protease active site. However, the increased proteolytic stability of 63 comes at a price: this fluorinated peptide has lower antimicrobial activity than 61, due to its propensity to self-assemble into helical bundles in aqueous solution rather than toroidal pores within the bacterial membrane (the aggregation behavior of other highly fluorinated peptide helices is discussed in Section 5.4).
A related approach that can deliver increased proteolytic stability is to incorporate a fluorinated side chain as an additional structural feature. α,α-Disubstituted amino acids, in which a trifluoromethyl side chain is present in addition to a canonical side chain, can endow peptides with greater proteolytic stability if the location and stereochemistry of the trifluoromethyl substituent causes a clash within the protease active site [200].
In contrast with the aliphatic fluorinated side chains such as those seen in 6263 (Figure 5), the presence of aromatic fluorinated side chains seldom leads to greater proteolytic stability. For example, p-fluorophenylalanine has been incorporated into a variety of short peptides and globular proteins as a replacement for natural phenylalanine, but this usually leads to greater susceptibility, not resistance, to protease digestion [131,199,201,202,203,204,205,206]. This can be attributed to the likely ability of the p-fluorophenylalanine side chain to also bind efficiently within protease binding pockets that have evolved to accommodate natural phenylalanine. Indeed, the design of the prodrugs 5960 (Figure 4) highlights cases in which efficient hydrolysis of an amide bond adjacent to p-fluorophenylalanine is a desirable event as part of the prodrug strategy.

4.3. Resistance to P450 Oxidation

Compound 64 (Figure 6) is a potent inhibitor of the protease enzyme cathepsin K. As such, it is a promising lead compound for the treatment of osteoporosis. Compound 64 suffers from rapid metabolism in the body, due to the action of a cytochrome P450 enzyme (CYP3A), which catalyzes the hydroxylation of the leucine side chain of 64. This metabolic process is prevented in the fluorinated next-generation analog 65 (Figure 6), leading to dramatically enhanced bioavailability [207].

5. Fluorine, the “Tailor”: Folding Amino Acids, Peptides, and Proteins into Precise 3D Shapes

In the world of amino acids, peptides, and proteins, conformation is inextricably linked to function. Therefore, methods for controlling conformation can have a variety of valuable applications. Conformational control can be considered across a range of scales, from the individual amino acid level (Section 5.1) [73], to the peptide secondary structure (Section 5.2), to the protein tertiary structure (Section 5.3), and even to the quaternary structure (Section 5.4). Fluorination can have significant impacts on all of these scales.

5.1. Conformational Control at the Individual Amino Acid Level

Fluorine offers a unique ability to control the conformations of individual amino acid side chains. The polar C–F bond tends to align in predictable ways with neighboring functional groups [208,209]. For example, molecules containing a N+–C–C–F moiety preferentially adopt conformations in which the N+ and Fδ− atoms are gauche, due to electrostatic attraction. In another example, α-fluoroamides (i.e., molecules containing a F–C–C(=O)–NH moiety) preferentially adopt a conformation in which the CF and CO bonds are anti-periplanar, which can be rationalized in terms of dipolar forces. In yet another example, molecules containing a F–C–C–H moiety preferentially adopt a conformation in which the CF and CH bonds are anti-periplanar, due to σCH→σ*CF hyperconjugation (note that this latter effect can reinforce the conformational preference described above for N+–C–C–F compounds).
Such conformational effects can be exploited to control the conformations of amino acid side chains, and this has been demonstrated most notably for proline [27,210,211,212,213,214,215,216]. For example, fluorination at the 4-position of the proline side chain (i.e., 33 and 66, Figure 7) can stabilize either the C4-endo or C4-exo pucker depending on the configuration of the fluorinated stereocentre, which is attributable to σCH→σ*CF hyperconjugation in each case. This ability to control the pucker of the proline ring has been exploited for a range of applications, including enhancing the enantioselectivity of organocatalytic reactions (e.g., 7073, Figure 7) [217,218,219]. Fluorination at the 3-position of the proline side chain can influence the pucker in a similar way [220,221].
Fluorine can also influence the conformation of the ring-expanded proline analog, pipecolic acid (69, Figure 7) [222,223]. The six-membered ring of 69 preferentially adopts a chair conformation in which the carboxyl group is equatorial. This conformation is maintained in the difluorinated analog 67 (Figure 7), with the N+–C–C–F and F–C–C–F moieties both adopting favorable gauche alignments. In contrast, the diastereoisomeric analog 68 adopts a ring-flipped conformation in which the carboxyl group is forced into the axial position.
The ability of fluorine to control the conformations of proline analogs will be further examined in Section 5.2, Section 5.3 and Section 6.2.
Another context in which fluorine-derived conformational control of amino acid side chains can be valuable is in the elucidation of the binding conformation of certain receptor ligands. For example, the N-methyl-d-aspartate (NMDA) receptor is a target of interest for the treatment of several disorders of the nervous system. An X-ray crystal structure of this receptor bound to its natural ligand (NMDA, 74) reveals a “bent” ligand conformation in which the carboxylate substituents of 74 are gauche to one another (Figure 7) [224]. This information is supported by the relative activity of two fluorinated NMDA analogs, 75 and 76 (Figure 7) [225]. For analog 75, binding leads to a favorable gauche N+–C–C–F alignment, and as a result, this ligand exhibits strong agonism almost equal to that of the native ligand, 74. In contrast, for analog 76, binding would require an unfavorable anti-N+–C–C–F angle, and as a result, this ligand is virtually inactive.

5.2. Peptide Secondary Structure

We have seen that fluorine can influence the conformations of individual amino acid side chains (Section 5.1). Now, if a fluorinated amino acid is elaborated into a peptide, the fluorine-derived conformational control can sometimes extend beyond the individual amino acid and can also start to influence the preferred rotameric species along the peptide backbone.
Consider again the example of fluorinated proline. It has been established that in peptides containing trans-4-fluoroproline (e.g., 77, Figure 8a), the amide bond preceding the fluoroproline residue strongly favors the trans-conformation [93,226,227,228]. In contrast, peptides containing cis-4-fluoroproline (e.g., 78, Figure 8a) favor a cis-amide conformation adjacent to the fluoroproline residue. This contrast can be exploited to alter the target-binding properties of a bioactive peptide. For example, the peptide sequence in 7778 is derived from the gastrin hormone G17. This hormone binds to a G-protein coupled receptor called cholecystokinin-2 (CCK-2R), which is overexpressed in a range of cancers. The key binding motif of G17 is thought to adopt a compact, hairpin-like structure; this conformation is better replicated by the cis-4-fluoroproline-containing peptide 78 (Figure 8a), endowing this peptide with higher CCK-2R binding affinity than the trans-4-fluoroproline-containing peptide 77 [93].
The idea of fluorine-derived conformational control holding small peptides into desired shapes for binding to targets is further examined in Section 6.3.
The ability of side chain-fluorinated amino acids to influence peptide secondary structure is not just limited to hairpin turns: α-helices and β-sheets can be affected too [229]. For example, peptide 79 (Figure 8b) is an engineered structure that is capable of adopting either an α-helix or a β-sheet conformation. When fluorine atoms are successively introduced into one side chain, the helix is destabilized to a greater and greater extent [230]. This can be attributed to the increased hydrophobicity of the highly fluorinated side chains, which are not favorably accommodated at the water-exposed edge of the helix. Intriguingly, in this peptide scaffold, there is an inverse relationship between α-helical and β-sheet propensity. This knowledge about the effects of side chain hydrophobicity on the kinetics and thermodynamics of β-sheet formation could be relevant in the future for designing amyloid-based materials, or perhaps even in understanding the progression of amyloid-based diseases.
It should be noted that the outcome can be different if not one, but multiple fluorinated residues are incorporated along the same edge of an α-helix [231]. In some cases, the helix can be stabilized because it contains a “fluorous edge”, which can engage in favorable supramolecular aggregation phenomena. This concept was mentioned in Section 4.2 and is explored further in Section 5.4 of this review.

5.3. Protein Tertiary Structure

Side chain-fluorinated amino acids can act as protein “superfolders”. For example, collagen, which consists of many tripeptide repeats of typical structure 83 (Figure 9), must adopt an all-trans conformation in order to assemble into its final triple-helical structure. Fluorine can accelerate this folding process. The fluorine substituent in the non-natural collagen analog 84 exerts an inductive pull that lowers the amide C–N bond order, reducing the kinetic barrier to cis/trans isomerization, and thereby helping the peptide chain to find its way to the required all-trans conformation [232]. The fluorine also provides thermodynamic stabilization of the final structure by favoring the C4-exo pucker, which matches the pucker found in natural collagen [233].
Fluoroprolines have been incorporated into several other proteins besides collagen [211,212,214,215,234,235,236,237]. For example, fluoroproline incorporation has led to a “superfolding” analog of green fluorescent protein (GFP). The crystal structure of GFP reveals that the majority of its proline residues (9 out of 10) adopt the C4-endo pucker [238]. An analog of GFP in which all 10 proline residues are replaced with cis-4-fluoroproline (85, Figure 9) displays enhanced folding characteristics, attributable to (i) lowering of the kinetic barrier to cis/trans peptide bond isomerization, as described above for collagen, and (ii) thermodynamic stabilization of the required C4-endo pucker [239]. In contrast, the corresponding GFP analog containing trans-4-fluoroprolines does not fold efficiently and lacks fluorescent properties.
Fluorination can modulate protein tertiary structures in other ways too. For example, fluorination can be exploited to enhance or disrupt a specific interaction within a protein core, in order to assess the contribution of that interaction to the protein’s structure and/or function [58,139,240,241]. This idea has been applied to study both aromatic–aromatic and sulfur–π interactions within the dopamine D2 receptor (86, Figure 9) [242]. Analogs of 86 were created in which fluorine atoms were successively incorporated, separately, into the side chains of Phe197 and Trp342. Increasing levels of fluorination were found to correlate with reduced protein function in both cases, and this was taken as evidence that Phe197 and Trp342 probably participate as electron-rich components within the native protein in aromatic–aromatic and sulfur–π interactions, respectively. Cation–π interactions within other proteins have also been interrogated in a similar fashion [242,243,244,245].
Having seen examples of individual, targeted interactions mediated by fluorine within a protein core, let us now consider a scenario where there are multiple, non-specific interactions. Proteins have been engineered to have multiple, highly fluorinated amino acids buried within the hydrophobic interior (e.g., 87, Figure 9) [246,247,248,249,250,251]. For example, protein 87 contains 12 hexafluoroleucine residues (i.e., 72 fluorine atoms) whose side chains form a highly fluorinated spine within the protein core. Such “fluorous core” proteins typically display markedly greater structural stability compared with their non-fluorinated counterparts (e.g., ΔG°fold = −27.6 kcal/mol for 87, compared with ΔG°fold = −18.0 kcal/mol for the non-fluorinated parent protein) [251]. The increased structural stability of “fluorous core” proteins such as 87 can be rationalized by comparing the relative energies of the unfolded and folded states [87,137,251,252,253]. In the unfolded state, the perfluorinated side chains of 87 are solvent-exposed; this is unfavorable because the fluorinated moieties cannot interact attractively with the water solvent, and also because the fluorinated moieties occlude the backbone NH and CO groups from hydrogen bonding with water. These unfavorable phenomena are avoided in the folded state.

5.4. Protein Quaternary Structure

We consider again the structure of the “fluorous-core” protein 87 (Figure 9). It comprises a bundle of four α-helices, which are connected from one to the next via peptide loops. It is intriguing to consider a hypothetical scenario in which the loops were removed so that 87 was no longer a single protein but rather four separate α-helical peptides. Would the fluorous packing effect be strong enough to induce separate “fluorous-edged” peptide helices to come together and form higher-order structures?
The answer is yes [254]. Several peptide systems have been engineered that can self-assemble, zipper-like, into aggregates via a fluorous interface (e.g., 88, Figure 10) [21,197,198,255,256,257,258,259,260,261,262,263,264]. A notable feature of this fluorine-directed self-assembly phenomenon is that it can be engineered to occur either within hydrophilic peptide helices (which can dimerize in aqueous solution) or within hydrophobic peptide helices (which can dimerize while embedded within a lipid membrane) [265,266]. This versatility comes from the amphipathic character of the “fluorous edge”.
Self-association phenomena have been investigated with other fluorinated peptide systems too. Certain peptides have a propensity to aggregate into supramolecular architectures like fibrils or hydrogels. Such aggregation-prone peptides can have a wide range of sizes, but a common feature is the presence of phenylalanine residues within the sequence. In several cases, the replacement of phenylalanine residues with fluorophenylalanine has been found to modify the propensity for self-assembly and/or the mechanical properties of the resulting supramolecular architecture [267,268,269,270].
All of the quaternary structures that we have examined so far have been of homomeric species. We will now conclude this section with an example of heteromeric peptide aggregation that is driven by fluorine-based interactions [271]. The first peptide of interest is a designed 32-residue scaffold known as α2D (89, Figure 10). In the absence of any other peptides, α2D folds and aggregates into a homodimer in which four phenylalanine side chains (two from each monomer) form stacking interactions at the dimer interface. This homodimeric folding pattern is also seen with a fluorinated analog of α2D (90, Figure 10) in which the phenylalanine residues are replaced with pentafluorophenylalanine. Remarkably, however, if the two homodimers 8990 are mixed together, they re-assemble selectively into heterodimers (91) as depicted in Figure 10. This preference for heterodimerization can be explained by the quadrupolar attraction of the phenyl group of 89 with the pentafluorophenyl group of 90.

6. Fluorine, the “Soldier”: Guiding Amino Acids, Peptides, and Proteins to Hit Their Biological Targets

Our attention now turns from “within” a biomolecule to its surroundings, i.e., how an amino acid or peptide interacts with its broader environment. Fluorination can modulate this in useful ways, for example, in the design of mechanism-based enzyme inhibitors (Section 6.1), in the enhancement of the intermolecular forces between an amino acid side chain and its biological target (Section 6.2), or through conformational pre-organization of a bioactive peptide (Section 6.3).

6.1. Mechanism-Based Enzyme Inhibitors

Side chain-fluorinated amino acids have proven to be useful as mechanism-based inhibitors of a range of pyridoxal phosphate (PLP) dependent enzymes, including amino acid racemases, decarboxylases, and transaminases [272,273,274]. Several such enzymes are of medicinal importance as targets for the treatment of African sleeping sickness, hirsutism, epilepsy, and cancer [275]. A fluorine atom strategically located at the β-position of a substrate mimic (e.g., 92, Scheme 3) provides a leaving group that can ultimately lead to irreversible alkylation, and hence inactivation, of the enzyme.

6.2. Enhancing Intermolecular Forces

Fluorination can also enhance the non-covalent binding of amino acids and peptides towards their biological targets. This is possible because fluorination can modulate intermolecular forces.
The first type that we will consider is hydrophobic (or van der Vaals) forces [276]. In some instances, the presence of a fluorinated substituent like -CF3 on the side chain of a peptide ligand (e.g., 93, Figure 11) [277] can provide good size- and shape-complementarity within a binding pocket of a biomacromolecule; the fluorines contribute to the binding interaction simply by presenting an appropriately shaped hydrophobic volume.
In other instances, fluorination can deliver more targeted interactions within a binding site. For example, a common structural motif within peptide-based drugs is a fluorophenylalanine residue (e.g., 94, Figure 11) [37,278,279,280,281]. The presence of a polar C–F bond within the phenylalanine side chain of 94 offers the opportunity for adventitious dipolar interactions to be achieved within a hydrophobic binding pocket [282], leading to enhanced affinity. This is separate from the other benefits that we have already seen in terms of the pharmacokinetic properties of fluorophenylalanines (see Section 4). Other aromatic amino acids (e.g., tryptophan) have also been fluorinated as a means to alter their dipolar character and enhance their bioactivity [283].
Another type of intermolecular force that aromatic amino acid side chains can participate in is cation–π interactions. As discussed in Section 5.3, multiply-fluorinated aryl side chains can be incorporated into proteins in order to disrupt, and thereby measure the importance of, cation–π interactions to the protein’s tertiary structural integrity [242,243,244,245]. Those examples are technically intramolecular in nature. It is also possible to employ fluorination to study intermolecular cation–π interactions [284,285,286]. For example, pentafluorophenylalanine residues (i.e., 95, Figure 11) have been introduced into ion channel proteins in order to interrogate the contribution of cation–π interactions in the binding of ion channel blockers such as tetrodotoxin [287,288].
Fluorine substituents can also alter the hydrogen-bonding character of amino acid side chains. For example, fluorine has been introduced into tyrosine side chains (e.g., 96, Figure 11) as a means of modulating the hydrogen-bonding ability of the adjacent phenolic group through inductive effects [289,290]. This strategy has been applied to optimize the binding of fluorinated small-molecule ligands to their cognate receptors (e.g., 3-fluorotyrosine binding to an amino acid transporter protein). The conceptual inverse, in which a protein is fluorinated in order to alter its interactions with non-fluorinated small-molecule binders, is also possible. For example, the enzyme glutathione S-transferase contains a key tyrosine residue within its active site. Replacement of this key residue with 3-fluorotyrosine (96) resulted in an altered hydrogen-bonding ability within the active site and correspondingly altered reaction kinetics, which is information that helped to reveal the native enzyme’s catalytic mechanism [291,292].
The difluoromethyl (-CF2H) substituent (e.g., 97, Figure 11) is another example of a fluorinated hydrogen-bonding moiety that has been successfully incorporated into amino acid side chains to enhance target-binding affinity [293,294,295].
Finally, fluorine can be useful for the design of bioisosteres of post-translationally modified amino acids [296]. A common type of post-translational modification is the phosphorylation of tyrosine residues (i.e., to provide Ar-O-PO32−). Aryl phosphonates (i.e., residues containing Ar-CH2-PO32−) are non-hydrolyzable isosteres of tyrosine phosphates, and they can serve as inhibitors of phosphatase enzymes. However, the inhibitory potency is dramatically enhanced with the more advanced isosteres, difluorophosphonates (i.e., Ar-CF2-PO32−, e.g., 98, Figure 11) [46,297]. The fluorine substituents provide a closer mimicry of the phosphate group in terms of the hydrogen bond acceptor ability, but also in terms of the pKa2 and even the Ar-X-P bite angle [209].

6.3. Conformational Pre-Organization

Aside from the modulation of intermolecular forces (Figure 11), another way that fluorine can enhance the binding affinity of a peptide ligand towards its target is through conformational control [221]. If a ligand is flexible, then an entropic penalty must be paid upon target binding. However, if the ligand can be pre-organized into the target-binding conformation, then the entropic cost is pre-paid, and this can translate into higher binding affinity.
We have already seen an example of conformational pre-organization delivering higher binding affinity (Section 5.2). Another example of the concept of conformational pre-organization is seen with thrombin inhibitor 99 (Figure 12). This molecule is disordered in solution, with the proline side chain interconverting between the C4-endo and C4-exo puckers; however, only the latter conformation is suitable for target binding [210]. In the fluorinated analog 100, the required C4-exo pucker is pre-organized, and this leads to stronger target binding. Conversely, in the fluorinated analog 101, the “wrong” C4-endo pucker is pre-organized, and this leads to weaker target binding.

7. Conclusions

Side chain-fluorinated amino acids offer fascinating and fruitful research opportunities.
Synthetic chemistry underpins the field. A wide variety of methods for synthesizing side chain-fluorinated amino acids have now been established (outlined in Section 2.1 of this review), encompassing nucleophilic, electrophilic, metal-catalyzed, and photochemical fluorination methods. These synthetic advances have enabled the creation of structurally diverse amino acid targets, ranging from selectively fluorinated structures bearing one or a small number of fluorines on the side chain, all the way up to perfluorinated structures bearing a large number of fluorines. A range of methods for elaborating such fluorinated amino acids into peptides and proteins are also available (Section 2.2), including both chemical and biochemical strategies.
Fluorine is able to perform a variety of roles within amino acid side chains. The first role (described in Section 3 of this review) may be likened to that of a “spy”: fluorine enables the collection of information about the properties of biologically relevant molecules, through 19F-NMR or 18F-PET analysis. Fluorine’s second role (Section 4) is akin to that of a “tinker”: fluorine can repair some of the well-known pharmacokinetic problems associated with peptide-based drugs, including their permeability and their susceptibility to metabolism. Fluorine’s third role (Section 5) invites comparison with that of a “tailor”: fluorine can control the ways in which amino acids, peptides, and proteins fold. Fluorine’s fourth and final role (Section 6) may be likened to that of a “soldier”: fluorine can guide molecules to better hit their biological targets through both covalent and non-covalent means.
In the future, it seems likely that research into side chain-fluorinated amino acids will continue to yield valuable outcomes. Two areas merit special mention. First, the concept of site-selective, late-stage protein fluorination [298] is an exciting development that could enable a greater variety of fluorinated proteins to be created for diverse applications. Second, it is noteworthy that although many side chain-fluorinated amino acids have been examined in this review, the examples that are stereoselectively fluorinated are relatively scarce and mostly limited to proline analogs; there seems to be an opportunity to further explore a wider variety of stereoselectively fluorinated amino acids for applications such as conformational control and enhancement of target-binding.

Author Contributions

All authors performed literature review and contributed to the writing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shah, P.; Westwell, A.D. The role of fluorine in medicinal chemistry. J. Enzyme Inhib. Med. Chem. 2007, 22, 527–540. [Google Scholar] [CrossRef]
  2. Kirk, K.L. Fluorination in medicinal chemistry: Methods, strategies, and recent developments. Org. Process Res. Dev. 2008, 12, 305–321. [Google Scholar] [CrossRef]
  3. Zhou, Y.; Wang, J.; Gu, Z.; Wang, S.; Zhu, W.; Acena, J.L.; Soloshonok, V.A.; Izawa, K.; Liu, H. Next generation of fluorine-containing pharmaceuticals, compounds currently in phase II–III clinical trials of major pharmaceutical companies: New structural trends and therapeutic areas. Chem. Rev. 2016, 116, 422–518. [Google Scholar] [CrossRef]
  4. Meanwell, N.A. Fluorine and fluorinated motifs in the design and application of bioisosteres for drug design. J. Med. Chem. 2018, 61, 5822–5880. [Google Scholar] [CrossRef]
  5. Inoue, M.; Sumii, Y.; Shibata, N. Contribution of Organofluorine Compounds to Pharmaceuticals. ACS Omega 2020, 5, 10633–10640. [Google Scholar] [CrossRef]
  6. Haufe, G.; Leroux, F. Fluorine in Life Sciences: Pharmacueticals, Medicinal Diagnostics, and Agrochemicals; Academic Press: London, UK, 2018. [Google Scholar]
  7. Johnson, B.M.; Shu, Y.Z.; Zhuo, X.; Meanwell, N.A. Metabolic and Pharmaceutical Aspects of Fluorinated Compounds. J. Med. Chem. 2020, 63, 6315–6386. [Google Scholar] [CrossRef]
  8. Gillis, E.P.; Eastman, K.J.; Hill, M.D.; Donnelly, D.J.; Meanwell, N.A. Applications of Fluorine in Medicinal Chemistry. J. Med. Chem. 2015, 58, 8315–8359. [Google Scholar] [CrossRef]
  9. Bohm, H.J.; Banner, D.; Bendels, S.; Kansy, M.; Kuhn, B.; Muller, K.; Obst-Sander, U.; Stahl, M. Fluorine in medicinal chemistry. ChemBioChem 2004, 5, 637–643. [Google Scholar] [CrossRef]
  10. Ogawa, Y.; Tokunaga, E.; Kobayashi, O.; Hirai, K.; Shibata, N. Current Contributions of Organofluorine Compounds to the Agrochemical Industry. iScience 2020, 23, 101467. [Google Scholar] [CrossRef]
  11. Leader, B.; Baca, Q.J.; Golan, D.E. Protein therapeutics: A summary and pharmacological classification. Nat. Rev. Drug Discov. 2008, 7, 21–39. [Google Scholar] [CrossRef]
  12. Doak, B.C.; Zheng, J.; Dobritzsch, D.; Kihlberg, J. How Beyond Rule of 5 Drugs and Clinical Candidates Bind to Their Targets. J. Med. Chem. 2016, 59, 2312–2327. [Google Scholar] [CrossRef]
  13. Lagasse, H.A.; Alexaki, A.; Simhadri, V.L.; Katagiri, N.H.; Jankowski, W.; Sauna, Z.E.; Kimchi-Sarfaty, C. Recent advances in (therapeutic protein) drug development. F1000Research 2017, 6, 113. [Google Scholar] [CrossRef] [PubMed]
  14. DeGoey, D.A.; Chen, H.J.; Cox, P.B.; Wendt, M.D. Beyond the Rule of 5: Lessons learned from AbbVie’s drugs and compound collection. J. Med. Chem. 2018, 61, 2636–2651. [Google Scholar] [CrossRef] [PubMed]
  15. Henninot, A.; Collins, J.C.; Nuss, J.M. The current state of peptide drug discovery: Back to the future? J. Med. Chem. 2018, 61, 1382–1414. [Google Scholar] [CrossRef] [PubMed]
  16. Lee, A.C.-L.; Harris, J.L.; Khanna, K.K.; Hong, J.-H. A Comprehensive Review on Current Advances in Peptide Drug Development and Design. Int. J. Mol. Sci. 2019, 20, 2383. [Google Scholar] [CrossRef] [PubMed]
  17. Craik, D.J.; Kan, M.W. How can we improve peptide drug discovery? Learning from the past. Expert Opin. Drug Discov. 2021, 16, 1399–1402. [Google Scholar] [CrossRef] [PubMed]
  18. Muttenthaler, M.; King, G.F.; Adams, D.J.; Alewood, P.F. Trends in peptide drug discovery. Nat. Rev. Drug Discov. 2021, 20, 309–325. [Google Scholar] [CrossRef]
  19. Mansour, F.; Hunter, L. Synthesis and applications of backbone-fluorinated amino acids. In Fluorine in Life Sciences: Pharmaceuticals, Medicinal Diagnostics, and Agrochemicals; Academic Press: Cambridge, MA, USA, 2019; pp. 325–347. [Google Scholar] [CrossRef]
  20. Yoder, N.C.; Kumar, K. Fluorinated amino acids in protein design and engineering. Chem. Soc. Rev. 2002, 31, 335–341. [Google Scholar] [CrossRef]
  21. Jäckel, C.; Koksch, B. Fluorine in Peptide Design and Protein Engineering. Eur. J. Org. Chem. 2005, 21, 4483–4503. [Google Scholar] [CrossRef]
  22. Salwiczek, M.; Nyakatura, E.K.; Gerling, U.I.; Ye, S.; Koksch, B. Fluorinated amino acids: Compatibility with native protein structures and effects on protein-protein interactions. Chem. Soc. Rev. 2012, 41, 2135–2171. [Google Scholar] [CrossRef]
  23. Buer, B.C.; Marsh, E.N.G. Design, synthesis, and study of fluorinated proteins. Methods Mol. Biol. 2014, 1216, 89–116. [Google Scholar] [CrossRef] [PubMed]
  24. Odar, C.; Winkler, M.; Wiltschi, B. Fluoro amino acids: A rarity in nature, yet a prospect for protein engineering. Biotechnol. J. 2015, 10, 427–446. [Google Scholar] [CrossRef] [PubMed]
  25. Berger, A.A.; Voller, J.S.; Budisa, N.; Koksch, B. Deciphering the Fluorine Code—The Many Hats Fluorine Wears in a Protein Environment. Acc. Chem. Res. 2017, 50, 2093–2103. [Google Scholar] [CrossRef]
  26. Mei, H.; Han, J.; White, S.; Graham, D.J.; Izawa, K.; Sato, T.; Fustero, S.; Meanwell, N.A.; Soloshonok, V.A. Tailor-Made Amino Acids and Fluorinated Motifs as Prominent Traits in Modern Pharmaceuticals. Chemistry 2020, 26, 11349–11390. [Google Scholar] [CrossRef] [PubMed]
  27. Kubyshkin, V.; Davis, R.; Budisa, N. Biochemistry of fluoroprolines: The prospect of making fluorine a bioelement. Beilstein J. Org. Chem. 2021, 17, 439–460. [Google Scholar] [CrossRef]
  28. Akcay, G.; Kumar, K. A New Paradigm for Protein Design and Biological Self-Assembly. J. Fluor. Chem. 2009, 130, 1178–1182. [Google Scholar] [CrossRef]
  29. Kukhar, V.P. Fluorine-containing amino acids. J. Fluor. Chem. 1994, 69, 199–205. [Google Scholar] [CrossRef]
  30. Taguchi, T.; Okada, M. Fluorinated cyclopropanes. J. Fluor. Chem. 2000, 105, 279–283. [Google Scholar] [CrossRef]
  31. Smits, R.; Cadicamo, C.D.; Burger, K.; Koksch, B. Synthetic strategies to α-trifluoromethyl and α-difluoromethyl substituted α-amino acids. Chem. Soc. Rev. 2008, 37, 1727–1739. [Google Scholar] [CrossRef]
  32. Sorochinsky, A.E.; Soloshonok, V.A. Asymmetric synthesis of fluorine-containing amines, amino alcohols, α- and β-amino acids mediated by chiral sulfinyl group. J. Fluor. Chem. 2010, 131, 127–139. [Google Scholar] [CrossRef]
  33. Qiu, X.L.; Qing, F.L. Recent Advances in the Synthesis of Fluorinated Amino Acids. Eur. J. Org. Chem. 2011, 2011, 3261–3278. [Google Scholar] [CrossRef]
  34. Moschner, J.; Stulberg, V.; Fernandes, R.; Huhmann, S.; Leppkes, J.; Koksch, B. Approaches to Obtaining Fluorinated α-Amino Acids. Chem. Rev. 2019, 119, 10718–10801. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, X.X.; Gao, Y.; Hu, X.S.; Ji, C.B.; Liu, Y.L.; Yu, J.S. Recent Advances in Catalytic Enantioselective Synthesis of Fluorinated α- and β-Amino Acids. Adv. Synth. Catal. 2020, 362, 4763–4793. [Google Scholar] [CrossRef]
  36. Mykhailiuk, P.K. Fluorine-containing prolines: Synthetic strategies, applications, and opportunities. J. Org. Chem. 2022, 87, 6961–7005. [Google Scholar] [CrossRef]
  37. Mei, H.; Han, J.; Klika, K.D.; Izawa, K.; Sato, T.; Meanwell, N.A.; Soloshonok, V.A. Applications of fluorine-containing amino acids for drug design. Eur. J. Med. Chem. 2020, 186, 111826. [Google Scholar] [CrossRef]
  38. Mikami, K.; Fustero, S.; Sánchez-Roselló, M.; Aceña, J.; Soloshonok, V.; Sorochinsky, A. Synthesis of Fluorinated β-Amino Acids. Synthesis 2011, 19, 3045–3079. [Google Scholar] [CrossRef]
  39. Soloshonok, V.A.; Kukhar, V.P.; Galushko, S.V.; Svistunova, N.Y.; Avilov, D.V.; Kuz’mina, N.A.; Raevski, N.I.; Struchkov, Y.T.; Pysarevsky, A.P.; Belokon, Y.N. General method for the synthesis of enantiomerically pure β-hydroxy-α-amino acids, containing fluorine atoms in the side chains. Case of stereochemical distinction between methyl and trifluoromethyl groups. X-ray crystal and molecular structure of the nickel(II) complex of (2S,3S)-2(trifluoromethyl)threonine. J. Chem. Soc. Perkin Trans. 1 1993, 24, 3143–3155. [Google Scholar] [CrossRef]
  40. Damhaut, P.; Lemaire, C.; Plenevaux, A.; Brihaye, C.; Christiaens, L.; Comar, D. No-carrier-added aymmetric synthesis of α-methyl-α-amino acids labelled with fluorine-18. Tetrahedron 1997, 53, 5785–5796. [Google Scholar] [CrossRef]
  41. Fustero, S.; Navarro, A.; Pina, B.; García Soler, J.; Bartolomé, A.; Asensio, A.; Simón, A.; Bravo, P.; Fronza, G.; Volonterio, A.; et al. Enantioselective synthesis of fluorinated α-amino acids and derivative in combination with ring-closing metathesis: Intramolecular π-stacking interactions as a source of stereocontrol. Org. Lett. 2001, 3, 2621–2624. [Google Scholar] [CrossRef]
  42. Herbert, B.; Kim, I.H.; Kirk, K.L. Synthesis of 2-fluoro- and 6-fluoro-(2S,3R)-(3,4-dihydroxyphenyl)serine as potential in vivo precursors of fluorinated norepinephrines. J. Org. Chem. 2001, 66, 4892–4897. [Google Scholar] [CrossRef]
  43. Fustero, S.; Bartolomé, A.; Sanz-Cervera, J.F.; Sánchez-Roselló, M.; García Soler, J.; de Arellano, C.R.; Fuentes, A.S. Diastereoselective synthesis of fluorinated, seven-membered β-amino acids derivatives via ring-closing metathesis. Org. Lett. 2003, 5, 2523–2526. [Google Scholar] [CrossRef] [PubMed]
  44. Jiang, Z.-X.; Qin, Y.; Qing, F. Asymmetric synthesis of both enantiomers of anti-4,4,4-trifluorothreonine and 2-amino-4,4,4-trifluorobutanoic acid. J. Org. Chem. 2003, 68, 7544–7547. [Google Scholar] [CrossRef] [PubMed]
  45. Berkowitz, D.B.; de la Salud-Bea, R.; Jahng, W. Synthesis of quaternary amino acids bearing a (2’Z)-fluorovinyl α-branch: Potential PLP enzyme inactivators. Org. Lett. 2004, 6, 1821–1824. [Google Scholar] [CrossRef] [PubMed]
  46. Otaka, A.; Mitsuyama, E.; Watanabe, J.; Watanabe, H.; Fujii, N. Synthesis of fluorine-containing bioisosteres corresponding to phosphoamino acids and dipeptide units. Biopolymers 2004, 76, 140–149. [Google Scholar] [CrossRef] [PubMed]
  47. Pang, W.; Zhu, S.; Jiang, H.; Zhu, S. Transition metal-catalyzed formation of CF3-substituted α,β-unsaturated alkene and the synthesis of α-trifluoromethyl substituted β-amino ester. Tetrahedron 2006, 62, 11760–11765. [Google Scholar] [CrossRef]
  48. Yajima, T.; Nagano, H. Photoinduced diastereoselective addition of perfluoroalkyl iodides to acrylic acid derivatives for the synthesis of fluorinated amino acids. Org. Lett. 2007, 9, 2513–2515. [Google Scholar] [CrossRef]
  49. Pigza, J.A.; Quach, T.; Molinski, T.F. Oxazoline-oxazinone oxidative rearrangement. divergent syntheses of (2S,3S)-4,4,4-trifluorovaline and (2S,4S)-5,5,5-Trifluoroleucine. J. Org. Chem. 2009, 74, 5510–5515. [Google Scholar] [CrossRef]
  50. Drège, E.; Guillaume, A.; Boukhedimi, N.; Marrot, J.; Troufflard, C.; Tran Huu-Dau, M.-E.; Joseph, D.; Delarue-Cochin, S. A facile and stereocontrolled synthesis of γ-substituted γ-fluoroglutamates by conjugate addition: Conflicting effect between fluorinated enaminoester and hinderered Michael acceptor. J. Org. Chem. 2010, 75, 7596–7604. [Google Scholar] [CrossRef]
  51. Aceña, J.L.; Sorochinsky, A.E.; Moriwaki, H.; Sato, T.; Soloshonok, V.A. Synthesis of fluorine-containing α-amino acids in enantiomerically pure form via homologation of Ni(II) complexes of glycine and alanine Schiff bases. J. Fluor. Chem. 2013, 155, 21–38. [Google Scholar] [CrossRef]
  52. Erdbrink, H.; Nyakatura, E.K.; Huhmann, S.; Gerling, U.I.M.; Lentz, D.; Koksch, B.; Czekelius, C. Synthesis of enantiomerically pure (2S,3S)-5,5,5-trifluoroisoleucine and (2R,3S)-5,5,5-trifluoro-allo-isoleucine. Beilstein J. Org. Chem. 2013, 9, 2009–2014. [Google Scholar] [CrossRef]
  53. Ivashkin, P.; Lemonnier, G.; Tora, A.S.; Pin, J.P.; Goudet, C.; Jubault, P.; Pannecoucke, X. Synthesis and studies on the mGluR agonist activity of FAP4 stereoisomers. Bioorg. Med. Chem. Lett. 2015, 25, 2523–2526. [Google Scholar] [CrossRef] [PubMed]
  54. Grigolato, L.; Brittain, W.D.G.; Hudson, A.S.; Czyzewska, M.M.; Cobb, S.L. Synthesis of pentafluorosulfanyl (SF5) containing aromatic amino acids. J. Fluor. Chem. 2018, 212, 166–170. [Google Scholar] [CrossRef] [PubMed]
  55. Zhao, J.-B.; Ren, X.; Zheng, B.-Q.; Ji, J.; Qiu, Z.-B.; Li, Y. A diastereoselective Mannich reaction of α-fluoroketones with ketimines: Construction of β-fluoroamine motifs with vicinal tetrasubstituted stereocenters. Tetrahedron Lett. 2018, 59, 2091–2094. [Google Scholar] [CrossRef]
  56. Ouchakour, L.; Ábrahámi, R.A.; Forró, E.; Haukka, M.; Fülöp, F.; Kiss, L. Stereocontrolled synthesis of fluorine-containing piperidine γ-amino acid derivatives. Eur. J. Org. Chem. 2019, 2019, 2202–2211. [Google Scholar] [CrossRef]
  57. Kirk, K.L.; Herbert, B.; Lu, S.F.; Jayachandran, B.; Padgett, W.L.; Olufunke, O.; Daly, J.W.; Haufe, G.; Laue, K.W. Chemical and biochemical approaches to the enantiomers of chiral fluorinated catecholamines and amino acids. In Asymmetric Fluoroorganic Chemistry; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 1999; Volume 746, pp. 194–209. [Google Scholar]
  58. Zheng, H.; Comeforo, K.; Gao, J. Expanding the fluorous arsenal: Tetrafluorinated phenylalanines for protein design. J. Am. Chem. Soc. 2009, 131, 18–19. [Google Scholar] [CrossRef]
  59. Pace, C.J.; Zheng, H.; Mylvaganam, R.; Kim, D.; Gao, J. Stacked fluoroaromatics as supramolecular synthons for programming protein dimerization specificity. Angew. Chem. Int. Ed. 2012, 51, 103–107. [Google Scholar] [CrossRef]
  60. Buller, A.R.; Brinkmann-Chen, S.; Romney, D.K.; Herger, M.; Murciano-Calles, J.; Arnold, F.H. Directed evolution of the tryptophan synthase β-subunit for stand-alone function recapitulates allosteric activation. Proc. Natl. Acad. Sci. USA 2015, 112, 14599–14604. [Google Scholar] [CrossRef]
  61. Islam, M.N.; Hitchings, R.; Kumar, S.; Fontes, F.L.; Lott, J.S.; Kruh-Garcia, N.A.; Crick, D.C. Mechanism of Fluorinated Anthranilate-Induced Growth Inhibition in Mycobacterium tuberculosis. ACS Infect. Dis. 2019, 5, 55–62. [Google Scholar] [CrossRef]
  62. Dong, C.; Huang, F.; Deng, H.; Schaffrath, C.; Spencer, J.B.; O’Hagan, D.; Naismith, J.H. Crystal structure and mechanism of fluorinating enzyme. Nature 2004, 427, 561–565. [Google Scholar] [CrossRef]
  63. Doi, M.; Nishi, Y.; Kiritoshi, N.; Iwata, T.; Mago, M.; Nakano, H.; Uchiyama, S.; Nakazawa, T.; Wakamiya, T.; Kobayashi, Y. Simple and efficient syntheses if Boc- and Fmoc-protected 4(R)- and 4(S)-fluoroproline solely from 4(R)-hydroxyproline. Tetrahedron 2002, 58, 8453–8459. [Google Scholar] [CrossRef]
  64. Hajduch, J.; Cramer, J.C.; Kirk, K.L. An Enantioselective Synthesis of (S)-4-Fluorohistidine. J. Fluor. Chem. 2008, 129, 807–810. [Google Scholar] [CrossRef] [PubMed]
  65. Shen, B.; Ehrlichmann, W.; Uebele, M.; Machulla, H.J.; Reischl, G. Automated synthesis of n.c.a. [18F]FDOPA via nucleophilic aromatic substitution with [18F]fluoride. Appl. Radiat. Isot. 2009, 67, 1650–1653. [Google Scholar] [CrossRef]
  66. Kiss, L.; Forro, E.; Fustero, S.; Fulop, F. Regio- and diastereoselective fluorination of alicyclic β-amino acids. Org. Biomol. Chem. 2011, 9, 6528–6534. [Google Scholar] [CrossRef] [PubMed]
  67. Kiss, L.; Nonn, M.; Sillanpaa, R.; Fustero, S.; Fulop, F. Efficient regio- and stereoselective access to novel fluorinated β-aminocyclohexanecarboxylates. Beilstein J. Org. Chem. 2013, 9, 1164–1169. [Google Scholar] [CrossRef]
  68. Remete, A.; Nonn, M.; Fustero, S.; Fülöp, F.; Kiss, L. A Stereocontrolled Protocol to Highly Functionalized Fluorinated Scaffolds through a Fluoride Opening of Oxiranes. Molecules 2016, 21, 1493. [Google Scholar] [CrossRef] [PubMed]
  69. Alluri, S.R.; Riss, P.J. Stereospecific radiosynthesis of 3-fluoro amino acids: Access to enantiomerically pure radioligands for positron emission tomography. Org. Biomol. Chem. 2018, 16, 2219–2224. [Google Scholar] [CrossRef]
  70. Malashchuk, A.; Chernykh, A.V.; Dobrydnev, A.V.; Grygorenko, O.O. Fluorine-labelled spiro[3.3]heptane-derived building blocks: Is single fluorine the best? Eur. J. Org. Chem. 2021, 2021, 4897–4910. [Google Scholar] [CrossRef]
  71. Padmakshan, D.; Bennett, S.; Otting, G.; Easton, C. Stereocontrolled Synthesis of (S)-γ-Fluoroleucine. Synlett 2007, 2007, 1083–1084. [Google Scholar] [CrossRef]
  72. Molnar, I.G.; Tanzer, E.M.; Daniliuc, C.; Gilmour, R. Enantioselective aziridination of cyclic enals facilitated by the fluorine-iminium ion gauche effect. Chemistry 2014, 20, 794–800. [Google Scholar] [CrossRef]
  73. Tsushima, T.; Kawada, K.; Nishikawa, J.; Sato, T.; Tori, K.; Tsuji, T.; Misaki, S. Fluorine-containing amino acids and their derivatives. 3. Stereoselective synthesis and unusual conformational features of threo- and erythro-3-fluorophenylalanine. J. Org. Chem. 1984, 49, 1163–1169. [Google Scholar] [CrossRef]
  74. Labroo, V.M.; Hebel, D.; Kirk, K.L.; Cohen, L.A.; Lemieux, C.; Schiller, P.W. Direct electrophilic fluorination of tyrosine in dermorphin analogues and its effect on biological activity, receptor affinity and selectivity. Int. J. Peptide Protein Res. 1991, 37, 430–439. [Google Scholar] [CrossRef]
  75. Pedregal, C.; Prowse, W. Stereoselective synthesis of 2-amino-3-fluoro bicyclo[3.1.0]hexane-2,6-dicarboxylic acid. Bioorg. Med. Chem. 2002, 10, 433–436. [Google Scholar] [CrossRef] [PubMed]
  76. Buckingham, F.; Kirjavainen, A.K.; Forsback, S.; Krzyczmonik, A.; Keller, T.; Newington, I.M.; Glaser, M.; Luthra, S.K.; Solin, O.; Gouverneur, V. Organomediated enantioselective 18F fluorination for PET applications. Angew. Chem. Int. Ed. 2015, 54, 13366–13369. [Google Scholar] [CrossRef] [PubMed]
  77. Wei, W.; Khangarot, R.K.; Stahl, L.; Veresmortean, C.; Pradhan, P.; Yang, L.; Zajc, B. Generating stereodiversity: Diastereoselective fluorination and highly diastereoselective epimerization of α-amino acid building blocks. Org. Lett. 2018, 20, 3574–3578. [Google Scholar] [CrossRef] [PubMed]
  78. Bandak, D.; Babii, O.; Vasiuta, R.; Komarov, I.V.; Mykhailiuk, P.K. Design and synthesis of novel 19F-amino acid: A promising 19F NMR label for peptide studies. Org. Lett. 2015, 17, 226–229. [Google Scholar] [CrossRef] [PubMed]
  79. Hebel, D.; Kirk, K.L.; Cohen, L.A.; Labroo, V.M. First direct fluorination of tyrosine-containing biologically active peptides. Tetrahedron Lett. 1990, 31, 619–622. [Google Scholar] [CrossRef]
  80. Lu, D.F.; Liu, G.S.; Zhu, C.L.; Yuan, B.; Xu, H. Iron(II)-catalyzed intramolecular olefin aminofluorination. Org. Lett. 2014, 16, 2912–2915. [Google Scholar] [CrossRef]
  81. Zhu, R.Y.; Tanaka, K.; Li, G.C.; He, J.; Fu, H.Y.; Li, S.H.; Yu, J.Q. Ligand-enabled stereoselective β-C(sp3)-H fluorination: Synthesis of unnatural enantiopure anti-β-fluoro-α-amino acids. J. Am. Chem. Soc. 2015, 137, 7067–7070. [Google Scholar] [CrossRef]
  82. Miao, J.; Yang, K.; Kurek, M.; Ge, H. Palladium-catalyzed site-selective fluorination of unactivated C(sp3)-H bonds. Org. Lett. 2015, 17, 3738–3741. [Google Scholar] [CrossRef]
  83. Lévine-Pinto, H.; Bouabdallah, B.; Morgat, J.L.; Gourdji, D.; Fromageot, P. Specific and direct fluorination of an histidine-containing peptide: Thyroliberin. Biochem. Biophys. Res. Commun. 1981, 103, 1121–1130. [Google Scholar] [CrossRef]
  84. Bume, D.D.; Pitts, C.R.; Jokhai, R.T.; Lectka, T. Direct, visible light-sensitized benzylic C-H fluorination of peptides using dibenzosuberenone: Selectivity for phenylalanine-like residues. Tetrahedron 2016, 72, 6031–6036. [Google Scholar] [CrossRef]
  85. Glaser, R.W.; Sachse, C.; Durr, U.H.; Wadhwani, P.; Ulrich, A.S. Orientation of the antimicrobial peptide PGLa in lipid membranes determined from 19F-NMR dipolar couplings of 4-CF3-phenylglycine labels. J. Magn. Reson. 2004, 168, 153–163. [Google Scholar] [CrossRef]
  86. Schottelius, M.; Berger, S.; Poethko, T.; Schwaiger, M.; Wester, H.-J. Development of novel 68Ga- and 18F-labeled GnRH-I analogues with high GnRHR-targeting efficiency. Bioconjug. Chem. 2008, 19, 1256–1268. [Google Scholar] [CrossRef]
  87. Clark, G.A.; Baleja, J.D.; Kumar, K. Cross-strand interactions of fluorinated amino acids in β-hairpin constructs. J. Am. Chem. Soc. 2012, 134, 17912–17921. [Google Scholar] [CrossRef] [PubMed]
  88. Peggion, C.; Biondi, B.; Battistella, C.; De Zotti, M.; Oancea, S.; Formaggio, F.; Toniolo, C. Spectroscopically labeled peptaibiotics. Synthesis and properties of selected trichogin GA IV analogs bearing a side-chain-monofluorinated aromatic amino acid for 19F-NMR analysis. Chem. Biodivers. 2013, 10, 904–919. [Google Scholar] [CrossRef] [PubMed]
  89. Kubyshkin, V.; Afonin, S.; Kara, S.; Budisa, N.; Mykhailiuk, P.K.; Ulrich, A.S. γ-(S)-Trifluoromethyl proline: Evaluation as a structural substitute of proline for solid state 19F-NMR peptide studies. Org. Biomol. Chem. 2015, 13, 3171–3181. [Google Scholar] [CrossRef] [PubMed]
  90. Wodtke, R.; Ruiz-Gomez, G.; Kuchar, M.; Pisabarro, M.T.; Novotna, P.; Urbanova, M.; Steinbach, J.; Pietzsch, J.; Loser, R. Cyclopeptides containing the DEKS motif as conformationally restricted collagen telopeptide analogues: Synthesis and conformational analysis. Org. Biomol. Chem. 2015, 13, 1878–1896. [Google Scholar] [CrossRef]
  91. Kuchar, M.; Neuber, C.; Belter, B.; Bergmann, R.; Lenk, J.; Wodtke, R.; Kniess, T.; Steinbach, J.; Pietzsch, J.; Loser, R. Evaluation of fluorine-18-labeled α1(I)-N-telopeptide analogs as substrate-based radiotracers for PET imaging of melanoma-associated lysyl oxidase. Front. Chem. 2018, 6, 121. [Google Scholar] [CrossRef]
  92. Arias, M.; Aramini, J.M.; Riopel, N.D.; Vogel, H.J. Fluorine-19 NMR spectroscopy of fluorinated analogs of tritrpticin highlights a distinct role for Tyr residues in antimicrobial peptides. Biochim. Biophys. Acta (BBA) Biomembr. 2020, 1862, 183260. [Google Scholar] [CrossRef]
  93. Corlett, A.; Sani, M.A.; Van Zuylekom, J.; Ang, C.S.; von Guggenberg, E.; Cullinane, C.; Blyth, B.; Hicks, R.J.; Roselt, P.D.; Thompson, P.E.; et al. A new turn in peptide-based imaging agents: Foldamers afford improved theranostics targeting cholecystokinin-2 receptor-positive cancer. J. Med. Chem. 2021, 64, 4841–4856. [Google Scholar] [CrossRef]
  94. O’Connor, N.K.; Rai, D.K.; Clark, B.R.; Murphy, C.D. Production of the novel lipopeptide antibiotic trifluorosurfactin via precursor-directed biosynthesis. J. Fluor. Chem. 2012, 143, 210–215. [Google Scholar] [CrossRef]
  95. O’Connor, N.K.; Hudson, A.S.; Cobb, S.L.; O’Neil, D.; Robertson, J.; Duncan, V.; Murphy, C.D. Novel fluorinated lipopeptides from Bacillus sp. CS93 via precursor-directed biosynthesis. Amino Acids 2014, 46, 2745–2752. [Google Scholar] [CrossRef] [PubMed]
  96. Leick, V. Effect of actinomycin D and DL-p-fluorophenylalanine on ribosome formation in Tetrahmena pyriformis. Eur. J. Biochem. 1969, 8, 215–220. [Google Scholar] [CrossRef]
  97. Takeuchi, T.; Prockop, D.J. Biosynthesis of abnormal collagens with amino acid analogues. 1. Incorporation of l-azetidine-2-carboxylic acid and cis-4-fluoro-l-proline into protocollagen and collagen. Biochim. Biophys. Acta (BBA) Protein Struct. 1969, 175, 142–155. [Google Scholar] [CrossRef] [PubMed]
  98. Barker, C.; Lewis, D. Impaired regulation of aromatic amino acid synthesis in a mutant resistant to p-fluorophenylalanine. J. Gen. Microbiol. 1974, 82, 337–343. [Google Scholar] [CrossRef]
  99. Hardy, C.; Binkley, S.B. The effect of p-fluorophenylalanine on nucleic acid biosynthesis and cell divisionin Escherichia coli. Biochemistry 1967, 6, 1892–1898. [Google Scholar] [CrossRef]
  100. Wang, P.; Tang, Y.; Tirrell, D.A. Incorporation of trifluoroisoleucine into proteins in vivo. J. Am. Chem. Soc. 2003, 125, 6900–6906. [Google Scholar] [CrossRef]
  101. Wang, P.; Fichera, A.; Kumar, K.; Tirrell, D.A. Alternative translations of a single RNA message: An identity switch of (2S,3R)-4,4,4-trifluorovaline between valine and isoleucine codons. Angew. Chem. Int. Ed. 2004, 43, 3664–3666. [Google Scholar] [CrossRef]
  102. Son, S.; Tanrikulu, I.C.; Tirrell, D.A. Stabilization of bzip peptides through incorporation of fluorinated aliphatic residues. ChemBioChem 2006, 7, 1251–1257. [Google Scholar] [CrossRef]
  103. 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, 11124–11127. [Google Scholar] [CrossRef]
  104. Rodriguez, E.A.; Lester, H.A.; Dougherty, D.A. Improved amber and opal suppressor tRNAs for incorporation of unnatural amino acids in vivo. Part 1: Minimizing misacylation. RNA 2007, 13, 1703–1714. [Google Scholar] [CrossRef]
  105. Ye, S.; Ann Berger, A.; Petzold, D.; Reimann, O.; Matt, B.; Koksch, B. Chemical aminoacylation of tRNAs with fluorinated amino acids for in vitro protein mutagenesis. Beilstein J. Org. Chem. 2010, 6, 40. [Google Scholar] [CrossRef]
  106. Kobayashi, T.; Hoppmann, C.; Yang, B.; Wang, L. Using Protein-Confined Proximity to Determine Chemical Reactivity. J. Am. Chem. Soc. 2016, 138, 14832–14835. [Google Scholar] [CrossRef]
  107. Alleyne, C.; Amin, R.P.; Bhatt, B.; Bianchi, E.; Blain, J.C.; Boyer, N.; Branca, D.; Embrey, M.W.; Ha, S.N.; Jette, K.; et al. Series of Novel and Highly Potent Cyclic Peptide PCSK9 Inhibitors Derived from an mRNA Display Screen and Optimized via Structure-Based Design. J. Med. Chem. 2020, 63, 13796–13824. [Google Scholar] [CrossRef]
  108. Ford, D.J.; Duggan, N.M.; Fry, S.E.; Ripoll-Rozada, J.; Agten, S.M.; Liu, W.; Corcilius, L.; Hackeng, T.M.; van Oerle, R.; Spronk, H.M.H.; et al. Potent Cyclic Peptide Inhibitors of FXIIa Discovered by mRNA Display with Genetic Code Reprogramming. J. Med. Chem. 2021, 64, 7853–7876. [Google Scholar] [CrossRef]
  109. Iskandar, S.E.; Bowers, A.A. mRNA Display Reaches for the Clinic with New PCSK9 Inhibitor. ACS Med. Chem. Lett. 2022, 13, 1379–1383. [Google Scholar] [CrossRef]
  110. Tucker, T.J.; Embrey, M.W.; Alleyne, C.; Amin, R.P.; Bass, A.; Bhatt, B.; Bianchi, E.; Branca, D.; Bueters, T.; Buist, N.; et al. A Series of Novel, Highly Potent, and Orally Bioavailable Next-Generation Tricyclic Peptide PCSK9 Inhibitors. J. Med. Chem. 2021, 64, 16770–16800. [Google Scholar] [CrossRef]
  111. Josephson, K.; Ricardo, A.; Szostak, J.W. mRNA display: From basic principles to macrocycle drug discovery. Drug Discov. Today 2014, 19, 388–399. [Google Scholar] [CrossRef]
  112. Newton, M.S.; Cabezas-Perusse, Y.; Tong, C.L.; Seelig, B. In vitro selection of peptides and proteins—Advantages of mRNA display. ACS Synth. Biol. 2020, 9, 181–190. [Google Scholar] [CrossRef]
  113. Nagumo, Y.; Fujiwara, K.; Horisawa, K.; Yanagawa, H.; Doi, N. PURE mRNA display for in vitro selection of single-chain antibodies. J. Biochem. 2016, 159, 519–526. [Google Scholar] [CrossRef]
  114. Sharaf, N.G.; Gronenborn, A.M. 19F-modified proteins and 19F-containing ligands as tools in solution NMR studies of protein interactions. Methods Enzymol. 2015, 565, 67–95. [Google Scholar] [CrossRef]
  115. Arntson, K.E.; Pomerantz, W.C. Protein-observed fluorine NMR: A bioorthogonal approach for small molecule discovery. J. Med. Chem. 2016, 59, 5158–5171. [Google Scholar] [CrossRef]
  116. Verhoork, S.J.M.; Killoran, P.M.; Coxon, C.R. Fluorinated Prolines as Conformational Tools and Reporters for Peptide and Protein Chemistry. Biochemistry 2018, 57, 6132–6143. [Google Scholar] [CrossRef]
  117. Divakaran, A.; Kirberger, S.E.; Pomerantz, W.C.K. SAR by (Protein-Observed) 19F NMR. Acc. Chem. Res. 2019, 52, 3407–3418. [Google Scholar] [CrossRef]
  118. Gimenez, D.; Phelan, A.; Murphy, C.D.; Cobb, S.L. 19F NMR as a tool in chemical biology. Beilstein J. Org. Chem. 2021, 17, 293–318. [Google Scholar] [CrossRef]
  119. Tsetlin, V.I.; Arseniev, A.S.; Utkin, Y.N.; Gurevich, A.Z.; Senyavina, L.B.; Bystrov, V.F.; Ivanov, V.T.; Ovchinnikov, Y.A. Conformational studies of Neurotoxin II from Naja naja oxiana. Eur. J. Biochem. 1979, 94, 337–346. [Google Scholar] [CrossRef]
  120. Liao, T.; Berlin, K.D. The use of p-fluorobenzenesulfonyl chloride as a reagent for studies of proteins by fluorine nuclear magnetic resonance. Anal. Biochem. 1985, 148, 365–375. [Google Scholar] [CrossRef]
  121. 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.; et al. Probing Ligand Binding Sites on Large Proteins by Nuclear Magnetic Resonance Spectroscopy of Genetically Encoded Non-Canonical Amino Acids. J. Med. Chem. 2023, 66, 5289–5304. [Google Scholar] [CrossRef]
  122. Huang, Y.; Reddy, K.D.; Bracken, C.; Qiu, B.; Zhan, W.; Eliezer, D.; Boudker, O. Environmentally ultrasensitive fluorine probe to resolve protein conformational ensembles by 19F NMR and cryo-EM. J. Am. Chem. Soc. 2023, 145, 8583–8592. [Google Scholar] [CrossRef]
  123. Hattori, Y.; Heidenreich, D.; Ono, Y.; Sugiki, T.; Yokoyama, K.I.; Suzuki, E.I.; Fujiwara, T.; Kojima, C. Protein 19F-labeling using transglutaminase for the NMR study of intermolecular interactions. J. Biomol. NMR 2017, 68, 271–279. [Google Scholar] [CrossRef]
  124. Lian, C.; Le, H.; Montez, B.; Patterson, J.; Harrell, S.; Laws, D.; Matsumura, I.; Pearson, J.; Oldfield, E. Fluorine-19 nuclear magnetic resonance spectroscopic study of fluorophenylalanine- and fluorotyrptophan-labeled avian egg white lysozymes. Biochemistry 1994, 33, 5238–5245. [Google Scholar] [CrossRef]
  125. 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, 13167–13176. [Google Scholar] [CrossRef] [PubMed]
  126. Dupureur, C.M.; Dominguez, M.A. The PD...(D/E)XK motif in restriction enzymes: A link between function and conformation. Biochemistry 2001, 40, 387–394. [Google Scholar] [CrossRef] [PubMed]
  127. Frieden, C. The kinetics of side chain stabilization during protein folding. Biochemistry 2003, 42, 12439–12446. [Google Scholar] [CrossRef] [PubMed]
  128. Wang, X.; Mercier, P.; Letourneau, P.J.; Sykes, B.D. Effects of Phe-to-Trp mutation and fluorotryptophan incorporation on the solution structure of cardiac troponin C, and analysis of its suitability as a potential probe for in situ NMR studies. Protein Sci. 2005, 14, 2447–2460. [Google Scholar] [CrossRef]
  129. Evanics, F.; Kitevski, J.L.; Bezsonova, I.; Forman-Kay, J.; Prosser, R.S. 19F NMR studies of solvent exposure and peptide binding to an SH3 domain. Biochim. Biophys. Acta (BBA) Gen. Subj. 2007, 1770, 221–230. [Google Scholar] [CrossRef]
  130. Li, C.; Lutz, E.A.; Slade, K.M.; Ruf, R.A.; Wang, G.F.; Pielak, G.J. 19F NMR studies of α-synuclein conformation and fibrillation. Biochemistry 2009, 48, 8578–8584. [Google Scholar] [CrossRef]
  131. Voloshchuk, N.; Zhu, A.Y.; Snydacker, D.; Montclare, J.K. Positional effects of monofluorinated phenylalanines on histone acetyltransferase stability and activity. Bioorg. Med. Chem. Lett. 2009, 19, 5449–5451. [Google Scholar] [CrossRef]
  132. Pomerantz, W.C.; Wang, N.; Lipinski, A.K.; Wang, R.; Cierpicki, T.; Mapp, A.K. Profiling the dynamic interfaces of fluorinated transcription complexes for ligand discovery and characterization. ACS Chem. Biol. 2012, 7, 1345–1350. [Google Scholar] [CrossRef]
  133. Mishra, N.K.; Urick, A.K.; Ember, S.W.; Schonbrunn, E.; Pomerantz, W.C. Fluorinated aromatic amino acids are sensitive 19F NMR probes for bromodomain-ligand interactions. ACS Chem. Biol. 2014, 9, 2755–2760. [Google Scholar] [CrossRef]
  134. Suzuki, Y.; Brender, J.R.; Soper, M.T.; Krishnamoorthy, J.; Zhou, Y.; Ruotolo, B.T.; Kotov, N.A.; Ramamoorthy, A.; Marsh, E.N.G. Resolution of oligomeric species during the aggregation of Aβ1–40 using 19F NMR. Biochemistry 2013, 52, 1903–1912. [Google Scholar] [CrossRef]
  135. Tressler, C.M.; Zondlo, N.J. (2S,4R)- and (2S,4S)-perfluoro-tert-butyl 4-hydroxyproline: Two conformationally distinct proline amino acids for sensitive application in 19F NMR. J. Org. Chem. 2014, 79, 5880–5886. [Google Scholar] [CrossRef]
  136. Tressler, C.M.; Zondlo, N.J. Perfluoro-tert-butyl hydroxyprolines as sensitive, conformationally responsive molecular probes: Detection of protein kinase activity by 19F NMR. ACS Chem. Biol. 2020, 15, 1096–1103. [Google Scholar] [CrossRef] [PubMed]
  137. Hoffmann, W.; Langenhan, J.; Huhmann, S.; Moschner, J.; Chang, R.; Accorsi, M.; Seo, J.; Rademann, J.; Meijer, G.; Koksch, B.; et al. An Intrinsic Hydrophobicity Scale for Amino Acids and Its Application to Fluorinated Compounds. Angew. Chem. Int. Ed. 2019, 58, 8216–8220. [Google Scholar] [CrossRef] [PubMed]
  138. Larda, S.T.; Simonetti, K.; Al-Abdul-Wahid, M.S.; Sharpe, S.; Prosser, R.S. Dynamic equilibria between monomeric and oligomeric misfolded states of the mammalian prion protein measured by 19F NMR. J. Am. Chem. Soc. 2013, 135, 10533–10541. [Google Scholar] [CrossRef] [PubMed]
  139. Woll, M.G.; Hadley, E.B.; Mecozzi, S.; Gellman, S.H. Stabilizing and destabilizing effects of phenylalanine → F5-phenylalanine mutations on the folding of a small protein. J. Am. Chem. Soc. 2006, 128, 15932–15933. [Google Scholar] [CrossRef]
  140. Esteban-Martin, S.; Strandberg, E.; Salgado, J.; Ulrich, A.S. Solid state NMR analysis of peptides in membranes: Influence of dynamics and labeling scheme. Biochim. Biophys. Acta (BBA) Biomembr. 2010, 1798, 252–257. [Google Scholar] [CrossRef]
  141. Grage, S.L.; Xu, X.; Schmitt, M.; Wadhwani, P.; Ulrich, A.S. 19F-Labeling of peptides revealing long-range NMR distances in fluid membranes. J. Phys. Chem. Lett. 2014, 5, 4256–4259. [Google Scholar] [CrossRef]
  142. Kokhan, S.O.; Tymtsunik, A.V.; Grage, S.L.; Afonin, S.; Babii, O.; Berditsch, M.; Strizhak, A.V.; Bandak, D.; Platonov, M.O.; Komarov, I.V.; et al. Design, synthesis, and application of an optimized monofluorinated aliphatic label for peptide studies by solid-state 19F NMR spectroscopy. Angew. Chem. Int. Ed. 2016, 55, 14788–14792. [Google Scholar] [CrossRef]
  143. Grage, S.L.; Sani, M.A.; Cheneval, O.; Henriques, S.T.; Schalck, C.; Heinzmann, R.; Mylne, J.S.; Mykhailiuk, P.K.; Afonin, S.; Komarov, I.V.; et al. Orientation and Location of the Cyclotide Kalata B1 in Lipid Bilayers Revealed by Solid-State NMR. Biophys. J. 2017, 112, 630–642. [Google Scholar] [CrossRef]
  144. Mikhailiuk, P.K.; Afonin, S.; Chernega, A.N.; Rusanov, E.B.; Platonov, M.O.; Dubinina, G.G.; Berditsch, M.; Ulrich, A.S.; Komarov, I.V. Conformationally rigid trifluoromethyl-substituted α-amino acid designed for peptide structure analysis by solid-state 19F NMR spectroscopy. Angew. Chem. Int. Ed. 2006, 45, 5659–5661. [Google Scholar] [CrossRef]
  145. Afonin, S.; Mikhailiuk, P.K.; Komarov, I.V.; Ulrich, A.S. Evaluating the amino acid CF3-bicyclopentylglycine as a new label for solid-state 19F-NMR structure analysis of membrane-bound peptides. J. Pept. Sci. 2007, 13, 614–623. [Google Scholar] [CrossRef] [PubMed]
  146. Yanagisawa, D.; Oda, K.; Inden, M.; Morikawa, S.; Inubushi, T.; Taniguchi, T.; Hijioka, M.; Kitamura, Y.; Tooyama, I. Fluorodopa is a Promising Fluorine-19 MRI Probe for Evaluating Striatal Dopaminergic Function in a Rat Model of Parkinson’s Disease. J. Neurosci. Res. 2017, 95, 1485–1494. [Google Scholar] [CrossRef] [PubMed]
  147. Orlandi, S.; Cavazzini, M.; Capuani, S.; Ciardello, A.; Pozzi, G. Synthesis and 19F NMR parameters of a perfluoro-tert-butoxy tagged l-DOPA analogue. J. Fluor. Chem. 2020, 237, 109596. [Google Scholar] [CrossRef]
  148. Tirotta, I.; Dichiarante, V.; Pigliacelli, C.; Cavallo, G.; Terraneo, G.; Bombelli, F.B.; Metrangolo, P.; Resnati, G. 19F magnetic resonance imaging (MRI): From design of materials to clinical applications. Chem. Rev. 2015, 115, 1106–1129. [Google Scholar] [CrossRef]
  149. Ishiwata, K.; Kubota, K.; Murakami, M.; Kubota, R.; Sasaki, T.; Ishii, S.; Seda, M. Re-evaluation of amino acid PET studies: Can the protein synthesis rates in brain and tumor tissues be easured in vivo? J. Nucl. Med. 1993, 34, 1936–1943. [Google Scholar]
  150. Makrides, V.; Bauer, R.; Weber, W.; Wester, H.J.; Fischer, S.; Hinz, R.; Huggel, K.; Opfermann, T.; Herzau, M.; Ganapathy, V.; et al. Preferred transport of O-(2-[18F]fluoroethyl)-d-tyrosine (D-FET) into the porcine brain. Brain Res. 2007, 1147, 25–33. [Google Scholar] [CrossRef]
  151. Kim, Y.; Lee, S.J.; Yook, C.M.; Oh, S.J.; Ryu, J.S.; Lee, J.J. Biological evaluation of new [18F]F-labeled synthetic amino acid derivatives as oncologic radiotracers. J. Label. Compd. Radiopharm. 2016, 59, 404–410. [Google Scholar] [CrossRef]
  152. Langen, K.J.; Hamacher, K.; Weckesser, M.; Floeth, F.; Stoffels, G.; Bauer, D.; Coenen, H.H.; Pauleit, D. O-(2-[18F]fluoroethyl)-l-tyrosine: Uptake mechanisms and clinical applications. Nucl. Med. Biol. 2006, 33, 287–294. [Google Scholar] [CrossRef]
  153. Fedorova, O.S.; Kuznetsova, O.F.; Shatik, S.V.; Stepanova, M.A.; Belokon, Y.N.; Maleev, V.I.; Krasikova, R.N. 18F-labeled tyrosine derivatives: Synthesis and experimental studies on accumulation in tumors and abscesses. Russ. J. Bioorg. Chem. 2009, 35, 306–314. [Google Scholar] [CrossRef]
  154. Stegmayr, C.; Stoffels, G.; Filss, C.; Heinzel, A.; Lohmann, P.; Willuweit, A.; Ermert, J.; Coenen, H.H.; Mottaghy, F.M.; Galldiks, N.; et al. Current trends in the use of O-(2-[18F]fluoroethyl)-l-tyrosine ([18F]FET) in neurooncology. Nucl. Med. Biol. 2021, 92, 78–84. [Google Scholar] [CrossRef] [PubMed]
  155. Franck, D.; Kniess, T.; Steinbach, J.; Zitzmann-Kolbe, S.; Friebe, M.; Dinkelborg, L.M.; Graham, K. Investigations into the synthesis, radiofluorination and conjugation of a new [18F]fluorocyclobutyl prosthetic group and its in vitro stability using a tyrosine model system. Bioorg. Med. Chem. 2013, 21, 643–652. [Google Scholar] [CrossRef] [PubMed]
  156. Betts, H.M.; Milicevic Sephton, S.; Tong, C.; Awais, R.O.; Hill, P.J.; Perkins, A.C.; Aigbirhio, F.I. Synthesis, in vitro evaluation, and radiolabeling of fluorinated puromycin analogues: Potential candidates for PET imaging of protein synthesis. J. Med. Chem. 2016, 59, 9422–9430. [Google Scholar] [CrossRef] [PubMed]
  157. Chiotellis, A.; Muller, A.; Weyermann, K.; Leutwiler, D.S.; Schibli, R.; Ametamey, S.M.; Kramer, S.D.; Mu, L. Synthesis and preliminary biological evaluation of O-2((2-[18F]fluoroethyl)methylamino)ethyltyrosine ([18F]FEMAET) as a potential cationic amino acid PET tracer for tumor imaging. Amino Acids 2014, 46, 1947–1959. [Google Scholar] [CrossRef]
  158. Maschauer, S.; Pischetsrieder, M.; Kuwert, T.; Prante, O. Utility of 1,3,4,6-tetra-O-acetyl-2-deoxy-2-[18F]fluoro-glucopyranoside for no-carrier-added 18F-glycosylation of amino acids. J. Label. Compd. Radiopharm. 2005, 48, 701–719. [Google Scholar] [CrossRef]
  159. Chiotellis, A.; Muller, A.; Mu, L.; Keller, C.; Schibli, R.; Kramer, S.D.; Ametamey, S.M. Synthesis and biological evaluation of 18F-labeled Fluoroethoxy tryptophan analogues as potential PET tumor imaging agents. Mol. Pharm. 2014, 11, 3839–3851. [Google Scholar] [CrossRef]
  160. Kiesewetter, D.O.; Gao, H.; Ma, Y.; Niu, G.; Quan, Q.; Guo, N.; Chen, X. 18F-radiolabeled analogs of exendin-4 for PET imaging of GLP-1 in insulinoma. Eur. J. Nucl. Med. Mol. Imaging 2012, 39, 463–473. [Google Scholar] [CrossRef]
  161. Liu, S.; Ma, H.; Zhang, Z.; Lin, L.; Yuan, G.; Tang, X.; Nie, D.; Jiang, S.; Yang, G.; Tang, G. Synthesis of enantiopure 18F-trifluoromethyl cysteine as a structure-mimetic amino acid tracer for glioma imaging. Theranostics 2019, 9, 1144–1153. [Google Scholar] [CrossRef]
  162. Baguet, T.; Bouton, J.; Janssens, J.; Pauwelyn, G.; Verhoeven, J.; Descamps, B.; Van Calenbergh, S.; Vanhove, C.; De Vos, F. Radiosynthesis, in vitro and preliminary biological evaluation of [18F]2-amino-4-((2-((3-fluorobenzyl)oxy)benzyl)(2-((3-(fluoromethyl)benzyl)oxy)benzyl)amino)butanoic acid, a novel alanine serine cysteine transporter 2 inhibitor-based positron emission tomography tracer. J. Label. Compd. Radiopharm. 2020, 63, 442–455. [Google Scholar] [CrossRef]
  163. Iovkova, L.; Könning, D.; Wängler, B.; Schirrmacher, R.; Schoof, S.; Arndt, H.D.; Jurkschat, K. SiFA-Modified phenylalanine: A key compound for the efficient synthesis of 18F-labelled peptides. Eur. J. Inorg. Chem. 2011, 2011, 2238–2246. [Google Scholar] [CrossRef]
  164. Bernard, J.; Malacea-Kabbara, R.; Clemente, G.S.; Burke, B.P.; Eymin, M.J.; Archibald, S.J.; Juge, S. o-Boronato- and o-trifluoroborato-phosphonium salts supported by l-α-amino acid side chain. J. Org. Chem. 2015, 80, 4289–4298. [Google Scholar] [CrossRef]
  165. Fuchtner, F.; Steinbach, J. Efficient synthesis of the 18F-labelled 3-O-methyl-6-[18F]fluoro-l-DOPA. Appl. Radiat. Isot. 2003, 58, 575–578. [Google Scholar] [CrossRef] [PubMed]
  166. Krasikova, R.N.; Zaitsev, V.V.; Ametamey, S.M.; Kuznetsova, O.F.; Fedorova, O.S.; Mosevich, I.K.; Belokon, Y.N.; Vyskocil, S.; Shatik, S.V.; Nader, M.; et al. Catalytic enantioselective synthesis of 18F-fluorinated α-amino acids under phase-transfer conditions using (S)-NOBIN. Nucl. Med. Biol. 2004, 31, 597–603. [Google Scholar] [CrossRef] [PubMed]
  167. Haase, C.; Bergmann, R.; Fuechtner, F.; Hoepping, A.; Pietzsch, J. L-type amino acid transporters LAT1 and LAT4 in cancer: Uptake of 3-O-methyl-6-18F-fluoro-l-DOPA in human adenocarcinoma and squamous cell carcinoma in vitro and in vivo. J. Nucl. Med. 2007, 48, 2063–2071. [Google Scholar] [CrossRef]
  168. Chondrogiannis, S.; Grassetto, G.; Marzola, M.C.; Rampin, L.; Massaro, A.; Bellan, E.; Ferretti, A.; Mazza, A.; Al-Nahhas, A.; Rubello, D. 18F-DOPA PET/CT biodistribution consideration in 107 consecutive patients with neuroendocrine tumours. Nucl. Med. Commun. 2012, 33, 179–184. [Google Scholar] [CrossRef] [PubMed]
  169. Li, C.T.; Palotti, M.; Holden, J.E.; Oh, J.; Okonkwo, O.; Christian, B.T.; Bendlin, B.B.; Buyan-Dent, L.; Harding, S.J.; Stone, C.K.; et al. A dual-tracer study of extrastriatal 6-[18F]fluoro-m-tyrosine and 6-[18F]-fluoro-l-DOPA uptake in Parkinson’s disease. Synapse 2014, 68, 325–331. [Google Scholar] [CrossRef] [PubMed]
  170. Yamaguchi, A.; Hanaoka, H.; Higuchi, T.; Tsushima, Y. Selective synthesis of l-2-[18F]fluoro-alpha-methylphenylalanine via copper-mediated 18F-fluorination of (mesityl)(aryl)iodonium salt. J. Label. Compd. Radiopharm. 2020, 63, 368–375. [Google Scholar] [CrossRef]
  171. Wester, H.; Herz, M.; Senekowitsch-Schmidtke, R.; Schwaiger, M.; Stöcklin, G.; Hamacher, K. Preclinical evaluation of 4-[18F]fluoroprolines: Diastereomeric effect on metabolism and uptake in mice. Nucl. Med. Biol. 1999, 26, 259–265. [Google Scholar] [CrossRef]
  172. Yu, W.; Williams, L.; Camp, V.M.; Malveaux, E.; Olson, J.J.; Goodman, M.M. Stereoselective synthesis and biological evaluation of syn-1-amino-3-[18F]fluorocyclobutyl-1-carboxylic acid as a potential positron emission tomography brain tumor imaging agent. Bioorg. Med. Chem. 2009, 17, 1982–1990. [Google Scholar] [CrossRef]
  173. Koglin, N.; Mueller, A.; Berndt, M.; Schmitt-Willich, H.; Toschi, L.; Stephens, A.W.; Gekeler, V.; Friebe, M.; Dinkelborg, L.M. Specific PET imaging of xC transporter activity using a 18F-labeled glutamate derivative reveals a dominant pathway in tumor metabolism. Clin. Cancer Res. 2011, 17, 6000–6011. [Google Scholar] [CrossRef]
  174. Schuster, D.M.; Nanni, C.; Fanti, S.; Oka, S.; Okudaira, H.; Inoue, Y.; Sörensen, J.; Owenius, R.; Choyke, P.; Turkbey, B.; et al. Anti-1-amino-3-18F-fluorocyclobutane-1-carboxylic acid: Physiologic uptake patterns, incidental findings, and variants that may simulate disease. J. Nucl. Med. 2014, 55, 1986–1992. [Google Scholar] [CrossRef]
  175. Bouhlel, A.; Zhou, D.; Li, A.; Yuan, L.; Rich, K.M.; McConathy, J. Synthesis, radiolabeling, and biological evaluation of (R)- and (S)-2-amino-5-[18F]fluoro-2-methylpentanoic acid ((R)-, (S)-[18F]FAMPe) as potential positron emission tomography tracers for brain tumors. J. Med. Chem. 2015, 58, 3817–3829. [Google Scholar] [CrossRef]
  176. Dunphy, M.P.S.; Harding, J.J.; Venneti, S.; Zhang, H.; Burnazi, E.M.; Bromberg, J.; Omuro, A.M.; Hsieh, J.J.; Mellinghoff, I.K.; Staton, K.; et al. in vivo PET Assay of tumor glutamine flux and metabolism: In-human trial of 18F-(2S,4R)-4-fluoroglutamine. Radiology 2018, 287, 667–675. [Google Scholar] [CrossRef] [PubMed]
  177. Čolović, M.; Yang, H.; Merkens, H.; Colpo, N.; Bénard, F.; Schaffer, P. The effect of chirality on the application of 5-[18F]fluoro-aminosuberic acid ([18F]FASu) for oxidative stress imaging. Mol. Imaging Biol. 2019, 22, 873–882. [Google Scholar] [CrossRef] [PubMed]
  178. Wu, R.; Liu, S.; Liu, Y.; Sun, Y.; Cheng, X.; Huang, Y.; Yang, Z.; Wu, Z. Synthesis and biological evaluation of [18F](2S,4S)4-(3-fluoropropyl) arginine as a tumor imaging agent. Eur. J. Med. Chem. 2019, 183, 111730. [Google Scholar] [CrossRef]
  179. Liu, S.; Wu, R.; Sun, Y.; Ploessl, K.; Zhang, Y.; Liu, Y.; Wu, Z.; Zhu, L.; Kung, H.F. Design, synthesis and evaluation of a novel glutamine derivative (2S,4R)-2-amino-4-cyano-4-[18F]fluorobutanoic acid. New. J. Chem. 2020, 44, 9109–9117. [Google Scholar] [CrossRef]
  180. Pickel, T.C.; Voll, R.J.; Yu, W.; Wang, Z.; Nye, J.A.; Bacsa, J.; Olson, J.J.; Liebeskind, L.S.; Goodman, M.M. Synthesis, radiolabeling, and biological evaluation of the cis stereoisomers of 1-amino-3-fluoro-4-(fluoro-18F)cyclopentane-1-carboxylic Acid as PET imaging agents. J. Med. Chem. 2020, 63, 12008–12022. [Google Scholar] [CrossRef] [PubMed]
  181. Pickel, T.C.; Pashikanti, G.; Voll, R.J.; Yu, W.; Zhang, Z.; Nye, J.A.; Bacsa, J.; Olson, J.J.; Liebeskind, L.S.; Goodman, M.M. Synthesis, radiolabeling, and biological evaluation of the trans-stereoisomers of 1-amino-3-(fluoro-18F)-4-fluorocyclopentane-1-carboxylic acid as PET imaging agents. ACS Pharmacol. Transl. Sci. 2021, 4, 1195–1203. [Google Scholar] [CrossRef]
  182. Lacan, G.; Satyamurthy, N.; Barrio, J.R. Synthesis of stereo (R and S) and geometric (E and Z) isomers of [18F]fluoro-β-fluoromethylene-m-tyrosine derivatives: In vivo Probes of central dopaminergic function. Nucl. Med. Biol. 1999, 26, 359–363. [Google Scholar] [CrossRef]
  183. Hanaoka, H.; Ohshima, Y.; Yamaguchi, A.; Suzuki, H.; Ishioka, N.S.; Higuchi, T.; Arano, Y.; Tsushima, Y. Novel 18F-labeled α-methyl-phenylalanine derivative with high tumor accumulation and ideal pharmacokinetics for tumor-specific imaging. Mol. Pharm. 2019, 16, 3609–3616. [Google Scholar] [CrossRef]
  184. Nodwell, M.B.; Yang, H.; Merkens, H.; Malik, N.; Colovic, M.; Bjorn, W.; Martin, R.E.; Benard, F.; Schaffer, P.; Britton, R. 18F-Branched-chain amino acids: Structure-activity relationships and PET imaging potential. J. Nucl. Med. 2019, 60, 1003–1009. [Google Scholar] [CrossRef] [PubMed]
  185. Yanai, A.; Harada, R.; Iwata, R.; Yoshikawa, T.; Ishikawa, Y.; Furumoto, S.; Ishida, T.; Yanai, K. Site-specific labeling of F-18 proteins using a supplemented cell-free protein synthesis system and O-2-[18F]fluoroethyl-l-tyrosine: [18F]FET-HER2 affibody molecule. Mol. Imaging Biol. 2019, 21, 529–537. [Google Scholar] [CrossRef] [PubMed]
  186. Cheng, D.; Yin, D.; Li, G.; Wang, M.; Li, S.; Zheng, M.; Cai, H.; Wang, Y. Radiolabeling and in vitro and in vivo characterization of [18F]FB-[R8,15,21, L17]-VIP as a PET imaging agent for tumor overexpressed VIP receptors. Chem. Biol. Drug Des. 2006, 68, 319–325. [Google Scholar] [CrossRef] [PubMed]
  187. Ren, G.; Liu, Z.; Miao, Z.; Liu, H.; Subbarayan, M.; Chin, F.T.; Zhang, L.; Gambhir, S.S.; Cheng, Z. PET of malignant melanoma using 18F-labeled metallopeptides. J. Nucl. Med. 2009, 50, 1865–1872. [Google Scholar] [CrossRef]
  188. Rojas, S.; Nolis, P.; Gispert, J.D.; Spengler, J.; Albericio, F.; Herance, J.R.; Abad, S. Efficient cysteine labelling of peptides with N-succinimidyl 4-[18F]fluorobenzoate: Stability study and in vivo biodistribution in rats by positron emission tomography (PET). RSC Adv. 2013, 3, 8028–8036. [Google Scholar] [CrossRef]
  189. Yang, Y.; Richter, S.; Wuest, F.; Doschak, M.R. Synthesis and structural identification of fluorine-18 labeled parathyroid hormone. J. Label. Compd. Radiopharm. 2015, 58, 453–457. [Google Scholar] [CrossRef]
  190. Perreault, A.; Knight, J.C.; Wang, M.; Way, J.; Wuest, F. 18F-Labeled wild-type annexin V: Comparison of random and site-selective radiolabeling methods. Amino Acids 2016, 48, 65–74. [Google Scholar] [CrossRef]
  191. Yang, X.; Mease, R.C.; Pullambhatla, M.; Lisok, A.; Chen, Y.; Foss, C.A.; Wang, Y.; Shallal, H.; Edelman, H.; Hoye, A.T.; et al. [18F]Fluorobenzoyllysinepentanedioic acid carbamates: New scaffolds for positron emission tomography (PET) imaging of prostate-specific membrane antigen (PSMA). J. Med. Chem. 2016, 59, 206–218. [Google Scholar] [CrossRef]
  192. Gillman, K.W.; Starrett, J.E., Jr.; Parker, M.F.; Xie, K.; Bronson, J.J.; Marcin, L.R.; McElhone, K.E.; Bergstrom, C.P.; Mate, R.A.; Williams, R.; et al. Discovery and evaluation of BMS-708163, a potent, selective and orally bioavailable γ-secretase inhibitor. ACS Med. Chem. Lett. 2010, 1, 120–124. [Google Scholar] [CrossRef]
  193. Nakazato, A.; Kumagai, T.; Sakagami, K.; Yoshikawa, R.; Suzuki, Y.; Chaki, S.; Ito, H.; Taguchi, T.; Nakanishi, S.; Okuyama, S. Synthesis, SARs and pharmacological characterization of 2-amino-3 or 6-fluorobicyclo[3.1.0]hexane-2,6-dicarboxylic acid derivatives as potent, selective, and orally active group II metabotropic glutamate receptor agonists. J. Med. Chem. 2000, 43, 4893–4909. [Google Scholar] [CrossRef]
  194. Larsson, R.; Dhar, S.; Ehrsson, H.; Nygren, P.; Lewensohn, R. Comparison of the cytotoxic activity of melphalan with l-prolyl-m-l-sarcolysyl-l-p-fluorophenylalanine in human tumour cell lines and primary cultures of tumour cells from patients. Br. J. Cancer 1998, 78, 328–335. [Google Scholar] [CrossRef]
  195. Gullbo, J.; Dhar, S.; Luthman, K.; Ehrsson, H.; Lewensohn, R.; Nygren, P.; Larsson, R. Antitumor activity of the alkylating oligopeptides J1 (l-melphalanyl-p-l-fluorophenylalanine ethyl ester) and P2 (l-prolyl-m-l-sarcolysyl-p-l-fluorophenylalanine ethyl ester): Comparison with melphalan. Anti-Cancer Drugs 2003, 14, 617–624. [Google Scholar] [CrossRef] [PubMed]
  196. Vine, W.H.; Hsieh, K.; Marshall, G.R. Synthesis of fluorine-containing peptides. Analogues of angiotensin II containing hexafluorovaline. J. Med. Chem. 1981, 24, 1043–1047. [Google Scholar] [CrossRef] [PubMed]
  197. Gottler, L.M.; Lee, H.Y.; Shelburne, C.E.; Ramamoorthy, A.; Marsh, E.N.G. Using fluorous amino acids to modulate the biological activity of an antimicrobial peptide. ChemBioChem 2008, 9, 370–373. [Google Scholar] [CrossRef] [PubMed]
  198. Asante, V.; Mortier, J.; Schluter, H.; Koksch, B. Impact of fluorination on proteolytic stability of peptides in human blood plasma. Bioorg. Med. Chem. 2013, 21, 3542–3546. [Google Scholar] [CrossRef] [PubMed]
  199. Huhmann, S.; Koksch, B. Fine-tuning the proteolytic stability of peptides with fluorinated amino acids. Eur. J. Org. Chem. 2018, 2018, 3667–3679. [Google Scholar] [CrossRef]
  200. Koksch, B.; Sewald, N.; Hofmann, H.; Burger, K.; Jakubke, H. Proteolytically stable peptides by incorporation of α-Tfm amino acids. J. Pept. Sci. 1997, 3, 157–167. [Google Scholar] [CrossRef]
  201. Krowarsch, D.; Cierpicki, T.; Jelen, F.; Otlewski, J. Canonical protein inhibitors of serine proteases. Cell. Mol. Life Sci. 2003, 60, 2427–2444. [Google Scholar] [CrossRef]
  202. Budisa, N.; Wenger, W.; Wiltschi, B. Residue-specific global fluorination of Candida antarctica lipase B in Pichia pastoris. Mol. Biosyst. 2010, 6, 1630–1639. [Google Scholar] [CrossRef]
  203. Piekielna, J.; Perlikowska, R.; do-Rego, J.C.; do-Rego, J.L.; Cerlesi, M.C.; Calo, G.; Kluczyk, A.; Lapinski, K.; Tomboly, C.; Janecka, A. Synthesis of mixed opioid affinity cyclic endomorphin-2 analogues with fluorinated phenylalanines. ACS Med. Chem. Lett. 2015, 6, 579–583. [Google Scholar] [CrossRef]
  204. Huhmann, S.; Stegemann, A.K.; Folmert, K.; Klemczak, D.; Moschner, J.; Kube, M.; Koksch, B. Position-dependent impact of hexafluoroleucine and trifluoroisoleucine on protease digestion. Beilstein J. Org. Chem. 2017, 13, 2869–2882. [Google Scholar] [CrossRef]
  205. Middendorp, S.J.; Wilbs, J.; Quarroz, C.; Calzavarini, S.; Angelillo-Scherrer, A.; Heinis, C. Peptide Macrocycle Inhibitor of Coagulation Factor XII with Subnanomolar Affinity and High Target Selectivity. J. Med. Chem. 2017, 60, 1151–1158. [Google Scholar] [CrossRef]
  206. Laskowski, M.; Kato, I. Protein inhibitors of proteinases. Ann. Rev. Biochem. 1980, 49, 593–626. [Google Scholar] [CrossRef]
  207. Gauthier, J.Y.; Chauret, N.; Cromlish, W.; Desmarais, S.; Duong, L.T.; Falgueyret, J.P.; Kimmel, D.B.; Lamontagne, S.; Leger, S.; LeRiche, T.; et al. The discovery of odanacatib (MK-0822), a selective inhibitor of cathepsin K. Bioorg. Med. Chem. Lett. 2008, 18, 923–928. [Google Scholar] [CrossRef]
  208. Hunter, L. The C-F bond as a conformational tool in organic and biological chemistry. Beilstein J. Org. Chem. 2010, 6, 38. [Google Scholar] [CrossRef]
  209. O’Hagan, D. Understanding organofluorine chemistry. An introduction to the C-F bond. Chem. Soc. Rev. 2008, 37, 308–319. [Google Scholar] [CrossRef]
  210. Staas, D.D.; Savage, K.L.; Sherman, V.L.; Shimp, H.L.; Lyle, T.A.; Tran, L.O.; Wiscount, C.M.; McMasters, D.R.; Sanderson, P.E.; Williams, P.D.; et al. Discovery of potent, selective 4-fluoroproline-based thrombin inhibitors with improved metabolic stability. Bioorg. Med. Chem. 2006, 14, 6900–6916. [Google Scholar] [CrossRef]
  211. 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, 2924–2931. [Google Scholar] [CrossRef]
  212. Borgogno, A.; Ruzza, P. The impact of either 4-R-hydroxyproline or 4-R-fluoroproline on the conformation and SH3m-cort binding of HPK1 proline-rich peptide. Amino Acids 2013, 44, 607–614. [Google Scholar] [CrossRef]
  213. Catherine, C.; Oh, S.J.; Lee, K.-H.; Min, S.-E.; Won, J.-I.; Yun, H.; Kim, D.-M. Engineering thermal properties of elastin-like polypeptides by incorporation of unnatural amino acids in a cell-free protein synthesis system. Biotechnol. Bioprocess. Eng. 2015, 20, 417–422. [Google Scholar] [CrossRef]
  214. Dietz, D.; Kubyshkin, V.; Budisa, N. Applying γ-substituted prolines in the foldon peptide: Polarity contradicts preorganization. ChemBioChem 2015, 16, 403–406. [Google Scholar] [CrossRef]
  215. Rienzo, M.; Rocchi, A.R.; Threatt, S.D.; Dougherty, D.A.; Lummis, S.C. Perturbation of Critical Prolines in Gloeobacter violaceus Ligand-gated Ion Channel (GLIC) Supports Conserved Gating Motions among Cys-loop Receptors. J. Biol. Chem. 2016, 291, 6272–6280. [Google Scholar] [CrossRef]
  216. Patrick, D.A.; Gillespie, J.R.; McQueen, J.; Hulverson, M.A.; Ranade, R.M.; Creason, S.A.; Herbst, Z.M.; Gelb, M.H.; Buckner, F.S.; Tidwell, R.R. Urea derivatives of 2-aryl-benzothiazol-5-amines: A new class of potential drugs for human African trypanosomiasis. J. Med. Chem. 2017, 60, 957–971. [Google Scholar] [CrossRef]
  217. Chandler, C.L.; List, B. Catalytic, asymmetric transannular aldolizations: Total synthesis of (+)-hirsutene. J. Am. Chem. Soc. 2008, 130, 6737–6739. [Google Scholar] [CrossRef]
  218. Díaz, J.; Goodman, J.M. Proline-catalyzed aldol reactions of cyclic diketones: Fluorine modifies pathways as well as transition states. Tetrahedron 2010, 66, 8021–8028. [Google Scholar] [CrossRef]
  219. Yap, D.Q.J.; Cheerlavancha, R.; Lowe, R.; Wang, S.; Hunter, L. Investigation of cis- and trans-4-Fluoroprolines as Enantioselective Catalysts in a Variety of Organic Transformations. Aust. J. Chem. 2015, 68, 44–49. [Google Scholar] [CrossRef]
  220. Hofman, G.J.; Ottoy, E.; Light, M.E.; Kieffer, B.; Kuprov, I.; Martins, J.C.; Sinnaeve, D.; Linclau, B. Minimising conformational bias in fluoroprolines through vicinal difluorination. Chem. Commun. 2018, 54, 5118–5121. [Google Scholar] [CrossRef]
  221. Testa, A.; Lucas, X.; Castro, G.V.; Chan, K.H.; Wright, J.E.; Runcie, A.C.; Gadd, M.S.; Harrison, W.T.A.; Ko, E.J.; Fletcher, D.; et al. 3-Fluoro-4-hydroxyprolines: Synthesis, conformational analysis, and stereoselective recognition by the VHL E3 ubiquitin ligase for targeted protein degradation. J. Am. Chem. Soc. 2018, 140, 9299–9313. [Google Scholar] [CrossRef]
  222. Singh, S.; Martinez, C.-M.; Calvet-Vitale, S.; Prasad, A.K.; Prangé, T.; Dalko, P.I.; Dhimane, H. Synthesis and conformational analysis of fluorinated pipecolic acids. Synlett 2012, 23, 2421–2425. [Google Scholar] [CrossRef]
  223. Chen, S.; Ruan, Y.; Lu, J.L.; Hunter, L.; Hu, X.G. Diastereoselective synthesis and conformational analysis of 4,5-difluoropipecolic acids. Org. Biomol. Chem. 2020, 18, 8192–8198. [Google Scholar] [CrossRef]
  224. Vance, K.M.; Simorowski, N.; Traynelis, S.F.; Furukawa, H. Ligand-specific deactivation time course of GluN1/GluN2D NMDA receptors. Nat. Commun. 2011, 2, 294. [Google Scholar] [CrossRef]
  225. Chia, P.W.; Livesey, M.R.; Slawin, A.M.; van Mourik, T.; Wyllie, D.J.; O’Hagan, D. 3-Fluoro-N-methyl-D-aspartic acid (3F-NMDA) stereoisomers as conformational probes for exploring agonist binding at NMDA receptors. Chemistry 2012, 18, 8813–8819. [Google Scholar] [CrossRef]
  226. Mykhailiuk, P.K.; Kubyshkin, V.; Bach, T.; Budisa, N. Peptidyl-prolyl model study: How does the electronic effect influence the amide bond conformation? J. Org. Chem. 2017, 82, 8831–8841. [Google Scholar] [CrossRef]
  227. Horng, J.C.; Raines, R.T. Stereoelectronic effects on polyproline conformation. Protein Sci. 2006, 15, 74–83. [Google Scholar] [CrossRef]
  228. Newberry, R.W.; Raines, R.T. 4-Fluoroprolines: Conformational analysis and effects on the stability and folding of peptides and proteins. Top. Heterocycl. Chem. 2017, 48, 1–25. [Google Scholar] [CrossRef]
  229. Chiu, H.-P.; Suzuki, Y.; Gullickson, D.; Ahmad, R.; Kokona, B.; Fairman, R.; Cheng, R.P. Helix propensity of highly fluorinated amino acids. J. Am. Chem. Soc. 2006, 128, 15556–15557. [Google Scholar] [CrossRef]
  230. Gerling, U.I.M.; Salwiczek, M.; Cadicamo, C.D.; Erdbrink, H.; Czekelius, C.; Grage, S.L.; Wadhwani, P.; Ulrich, A.S.; Behrends, M.; Haufe, G.; et al. Fluorinated amino acids in amyloid formation: A symphony of size, hydrophobicity and α-helix propensity. Chem. Sci. 2014, 5, 819–830. [Google Scholar] [CrossRef]
  231. Lim, D.S.; Lin, J.-H.; Welch, J.T. The synthesis and characterization of a pentafluorosulfanylated peptide. Eur. J. Org. Chem. 2012, 21, 3946–3954. [Google Scholar] [CrossRef]
  232. Eberhardt, E.S.; Panasik, N.; Raines, R.T. Inductive effects of the energetics of prolyl peptide bond isomerization: Implications for collagen folding and stability. J. Am. Chem. Soc. 1996, 118, 12261–12266. [Google Scholar] [CrossRef]
  233. Holmgren, S.K.; Taylor, K.M.; Bretscher, L.E.; Raines, R.T. Code for collagen’s stability deciphered. Nature 1998, 392, 666–667. [Google Scholar] [CrossRef]
  234. Limapichat, W.; Lester, H.A.; Dougherty, D.A. Chemical scale studies of the Phe-Pro conserved motif in the cys loop of Cys loop receptors. J. Biol. Chem. 2010, 285, 8976–8984. [Google Scholar] [CrossRef] [PubMed]
  235. Rubini, M.; Scharer, M.A.; Capitani, G.; Glockshuber, R. (4R)- and (4S)-fluoroproline in the conserved cis-prolyl peptide bond of the thioredoxin fold: Tertiary structure context dictates ring puckering. ChemBioChem 2013, 14, 1053–1057. [Google Scholar] [CrossRef] [PubMed]
  236. Mosesso, R.; Dougherty, D.A.; Lummis, S.C.R. Probing Proline Residues in the Prokaryotic Ligand-Gated Ion Channel, ELIC. Biochemistry 2018, 57, 4036–4043. [Google Scholar] [CrossRef]
  237. Mosesso, R.; Dougherty, D.A.; Lummis, S.C.R. Proline residues in the transmembrane/extracellular domain interface loops have different behaviors in 5-HT3 and nACh receptors. ACS Chem. Neurosci. 2019, 10, 3327–3333. [Google Scholar] [CrossRef] [PubMed]
  238. Iwai, H.; Lingel, A.; Pluckthun, A. Cyclic green fluorescent protein produced in vivo using an artificially split PI-PfuI intein from Pyrococcus furiosus. J. Biol. Chem. 2001, 276, 16548–16554. [Google Scholar] [CrossRef]
  239. Moroder, L.; Budisa, N. Synthetic biology of protein folding. ChemPhysChem 2010, 11, 1181–1187. [Google Scholar] [CrossRef] [PubMed]
  240. Wynn, R.; Harkins, P.C.; Richars, F.M.; Fox, R.O. Comparison of straight chain and cyclic unnatural amino acids embedded in the core of staphylococcal nuclease. Protein Sci. 1997, 6, 1621–1626. [Google Scholar] [CrossRef]
  241. Cornilescu, G.; Hadley, E.B.; Woll, M.G.; Markley, J.L.; Gellman, S.H.; Cornilescu, C.C. Solution structure of a small protein containing a fluorinated side chain in the core. Protein Sci. 2006, 16, 14–19. [Google Scholar] [CrossRef]
  242. Daeffler, K.N.; Lester, H.A.; Dougherty, D.A. Functionally important aromatic-aromatic and sulfur-π interactions in the D2 dopamine receptor. J. Am. Chem. Soc. 2012, 134, 14890–14896. [Google Scholar] [CrossRef]
  243. Zhong, W.; Gallivan, J.P.; Zhang, Y.; Li, L.; Lester, H.A.; Dougherty, D.A. From ab initio quantum mechanics to molecular neurobiology: A cation-π binding site in the nicotinic receptor. Proc. Natl. Acad. Sci. USA 1998, 95, 12088–12093. [Google Scholar] [CrossRef]
  244. Beene, D.L.; Brandt, G.S.; Zhong, W.; Zacharias, N.M.; Lester, H.A.; Dougherty, D.A. Cation-π interactions in ligand recognition by serotonergic (5-HT3A) and nicotinic acetylcholine receptors: The anomalous binding properties of nicotine. Biochemistry 2002, 41, 10262–10269. [Google Scholar] [CrossRef]
  245. Lummis, S.C.; Beene, D.L.; Harrison, N.J.; Lester, H.A.; Dougherty, D.A. A cation-π binding interaction with a tyrosine in the binding site of the GABAC receptor. Chem. Biol. 2005, 12, 993–997. [Google Scholar] [CrossRef] [PubMed]
  246. Marsh, E.N.G. Towards the nonstick egg: Designing fluorous proteins. Chem. Biol. 2000, 7, R153–R157. [Google Scholar] [CrossRef]
  247. Lee, H.-Y.; Lee, K.-H.; Al-Hashimi, H.M.; Marsh, E.N.G. Modulating protein structure with fluorous amino acids: Increased stability and native-like structure conferred on a 4-helix bundle protein by hexafluoroleucine. J. Am. Chem. Soc. 2006, 128, 337–343. [Google Scholar] [CrossRef] [PubMed]
  248. Buer, B.C.; de la Salud-Bea, R.; Al Hashimi, H.M.; Marsh, E.N.G. Engineering protein stability and specificity using fluorous amino acids: The importance of packing effects. Biochemistry 2009, 48, 10810–10817. [Google Scholar] [CrossRef] [PubMed]
  249. Marsh, E.N.G. Fluorinated proteins: From design and synthesis to structure and stability. Acc. Chem. Res. 2014, 47, 2878–2886. [Google Scholar] [CrossRef]
  250. Gottler, L.M.; de la Salud-Bea, R.; Marsh, E.N.G. The fluorous effect in proteins: Properties of α4F6, a 4-α-helix bundle protein with a fluorocarbon core. Biochemistry 2008, 47, 4480–4490. [Google Scholar] [CrossRef]
  251. Buer, B.C.; Meagher, J.L.; Stuckey, J.A.; Marsh, E.N.G. Structural basis for the enhanced stability of highly fluorinated proteins. Proc. Natl. Acad. Sci. USA 2012, 109, 4810–4815. [Google Scholar] [CrossRef]
  252. Buer, B.C.; Meagher, J.L.; Stuckey, J.A.; Marsh, E.N.G. Comparison of the structures and stabilities of coiled-coil proteins containing hexafluoroleucine and t-butylalanine provides insight into the stabilizing effects of highly fluorinated amino acid side-chains. Protein Sci. 2012, 21, 1705–1715. [Google Scholar] [CrossRef]
  253. Robalo, J.R.; Huhmann, S.; Koksch, B.; Vila Verde, A. The Multiple Origins of the Hydrophobicity of Fluorinated Apolar Amino Acids. Chem 2017, 3, 881–897. [Google Scholar] [CrossRef]
  254. Lee, K.; Lee, H.; Slutsky, M.M.; Anderson, J.T.; Marsh, E.N.G. Fluorous effect in proteins: De novo design and characterization of a four-α-helix bundle protein containing hexafluoroleucine. Biochemistry 2004, 43, 16277–16284. [Google Scholar] [CrossRef] [PubMed]
  255. Bilgiçer, B.; Fichera, A.; Kumar, K. A coiled coil with a fluorous core. J. Am. Chem. Soc. 2001, 123, 4393–4399. [Google Scholar] [CrossRef]
  256. Bilgiçer, B.; Xing, X.; Kumar, K. Programmed self-sorting of coiled coils with leucine and hexafluoroleucine cores. J. Am. Chem. Soc. 2001, 123, 11815–11816. [Google Scholar] [CrossRef] [PubMed]
  257. 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, 923–925. [Google Scholar] [CrossRef]
  258. 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, 1494–1496. [Google Scholar] [CrossRef]
  259. Tang, Y.; Ghirlanda, G.; Vaidehi, N.; Kua, J.; Mainz, D.T.; Goddard, W.A.; DeGrado, W.F.; Tirrell, D.A. Stabilization of coiled-coil peptide omains by introduction of trifluoroleucine. Biochemistry 2001, 40, 2790–2796. [Google Scholar] [CrossRef]
  260. 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, 11089–11090. [Google Scholar] [CrossRef]
  261. Jäckel, C.; Seufert, W.; Thust, S.; Koksch, B. Evaluation of the molecular interactions of fluorinated amino acids with native polypeptides. ChemBioChem 2004, 5, 717–720. [Google Scholar] [CrossRef]
  262. Salwiczek, M.; Koksch, B. Effects of fluorination on the folding kinetics of a heterodimeric coiled coil. ChemBioChem 2009, 10, 2867–2870. [Google Scholar] [CrossRef]
  263. Salwiczek, M.; Samsonov, S.; Vagt, T.; Nyakatura, E.; Fleige, E.; Numata, J.; Colfen, H.; Pisabarro, M.T.; Koksch, B. Position-dependent effects of fluorinated amino acids on the hydrophobic core formation of a heterodimeric coiled coil. Chemistry 2009, 15, 7628–7636. [Google Scholar] [CrossRef]
  264. Gottler, L.M.; de la Salud-Bea, R.; Shelburne, C.E.; Ramamoorthy, A.; Marsh, E.N.G. Using fluorous amino acids to probe the effects of changing hydrophobicity on the physical and biological properties of the β-hairpin antimicrobial peptide protegrin-1. Biochemistry 2008, 47, 9243–9250. [Google Scholar] [CrossRef] [PubMed]
  265. Niemz, A.; Tirrell, D.A. Self-association and mebrane-binding behaviour of melittins containing trifluoroleucine. J. Am. Chem. Soc. 2001, 123, 7407–7413. [Google Scholar] [CrossRef] [PubMed]
  266. Bilgiçer, B.; Kumar, K. De novo design of defined helical bundles in membrane environments. Proc. Natl. Acad. Sci. USA 2004, 101, 15324–15329. [Google Scholar] [CrossRef] [PubMed]
  267. Cejas, M.A.; Kinney, W.A.; Chen, C.; Vinter, J.G.; Almond, H.R.; Balss, K.M.; Maryanoff, C.A.; Schmidt, U.; Breslav, M.; Mahan, A.; et al. Thrombogenic collagen-mimetic peptides: Self-assembly of triple helix-based fibrils drive by hydrophobic interactions. Proc. Natl. Acad. Sci. USA 2008, 105, 8513–8518. [Google Scholar] [CrossRef] [PubMed]
  268. Yuvienco, C.; More, H.T.; Haghpanah, J.S.; Tu, R.S.; Montclare, J.K. Modulating Supramolecular Assemblies and Mechanical Properties of Engineered Protein Materials by Fluorinated Amino Acids. Biomacromolecules 2012, 13, 2273–2278. [Google Scholar] [CrossRef]
  269. Kralj, S.; Bellotto, O.; Parisi, E.; Garcia, A.M.; Iglesias, D.; Semeraro, S.; Deganutti, C.; D’Andrea, P.; Vargiu, A.V.; Geremia, S.; et al. Heterochirality and Halogenation Control Phe-Phe Hierarchical Assembly. ACS Nano 2020, 14, 16951–16961. [Google Scholar] [CrossRef]
  270. Aviv, M.; Cohen-Gerassi, D.; Orr, A.A.; Misra, R.; Arnon, Z.A.; Shimon, L.J.W.; Shacham-Diamand, Y.; Tamamis, P.; Adler-Abramovich, L. Modification of a Single Atom Affects the Physical Properties of Double Fluorinated Fmoc-Phe Derivatives. Int. J. Mol. Sci. 2021, 22, 9634. [Google Scholar] [CrossRef]
  271. Zheng, H.; Gao, J. Highly specific heterodimerization mediated by quadrupole interactions. Angew. Chem. Int. Ed. 2010, 49, 8635–8639. [Google Scholar] [CrossRef]
  272. Silverman, R.B.; Abeles, R.H. Inactivation of pyridoxal phosphate dependent enzymes by mono-and polyhaloalanines. Biochemistry 1976, 15, 4718–4723. [Google Scholar] [CrossRef]
  273. Metcalf, B.W.; Bey, P.; Danzin, C.; Jung, M.J.; Casara, P.; Vevert, J.P. Catalytic irreversible inhibition of mammalian ornithine decarboxylase (E.C.4.1.1.17) by substrate and product analogs. J. Am. Chem. Soc. 1977, 100, 2551–2553. [Google Scholar] [CrossRef]
  274. Pan, Y.; Qiu, J.; Silverman, R.B. Design, synthesis, and biological activity of a difluoro-substituted, conformationally rigid vigabatrin analogue as a potent γ-aminobutyric acid aminotransferase inhibitor. J. Med. Chem. Lett 2003, 46, 5292–5293. [Google Scholar] [CrossRef] [PubMed]
  275. LoGiudice, N.; Le, L.; Abuan, I.; Leizorek, Y.; Roberts, S.C. Alpha-difluoromethylornithine, an irreversible inhibitor of polyamine biosynthesis, as a therapeutic strategy against hyperproliferative and infectious diseases. Med. Sci. 2018, 6, 12. [Google Scholar] [CrossRef] [PubMed]
  276. Ojemalm, K.; Higuchi, T.; Lara, P.; Lindahl, E.; Suga, H.; von Heijne, G. Energetics of side-chain snorkeling in transmembrane helices probed by nonproteinogenic amino acids. Proc. Natl. Acad. Sci. USA 2016, 113, 10559–10564. [Google Scholar] [CrossRef] [PubMed]
  277. Leppkes, J.; Dimos, N.; Loll, B.; Hohmann, T.; Dyrks, M.; Wieseke, A.; Keller, B.G.; Koksch, B. Fluorine-induced polarity increases inhibitory activity of BPTI towards chymotrypsin. RSC Chem. Biol. 2022, 3, 773–782. [Google Scholar] [CrossRef] [PubMed]
  278. Ippolito, J.A.; Christianson, D.W. The contribution of halogen atoms to protein-ligand interactions. Int. J. Biol. Macromol. 1992, 14, 193–197. [Google Scholar] [CrossRef]
  279. Kalindjian, S.B.; Buck, I.M.; Davies, J.M.R.; Dunstone, D.J.; Hudson, M.L.; Low, C.M.R.; McDonald, I.M.; Pether, M.J.; Steel, K.I.M.; Tozer, M.J.; et al. Non-peptide cholecystokinin-B/gastrin receptor antagonists based on bicyclic, heteroaromatic skeletons. J. Med. Chem. 1996, 39, 1806–1815. [Google Scholar] [CrossRef]
  280. Hoyt, S.B.; London, C.; Gorin, D.; Wyvratt, M.J.; Fisher, M.H.; Abbadie, C.; Felix, J.P.; Garcia, M.L.; Li, X.; Lyons, K.A.; et al. Discovery of a novel class of benzazepinone NaV1.7 blockers: Potential treatments for neuropathic pain. Bioorg. Med. Chem. Lett. 2007, 17, 4630–4634. [Google Scholar] [CrossRef]
  281. Piepenbrink, K.H.; Borbulevych, O.Y.; Sommese, R.F.; Clemens, J.; Armstrong, K.M.; Desmond, C.; Do, P.; Baker, B.M. Fluorine substitutions in an antigenic peptide selectively modulate T-cell receptor binding in a minimally perturbing manner. Biochem. J. 2009, 423, 353–361. [Google Scholar] [CrossRef]
  282. Müller, K.; Faeh, C.; Diederich, F. Fluorine in pharmaceuticals: Looking beyond intuition. Science 2007, 317, 1881–1886. [Google Scholar] [CrossRef]
  283. Andersen, O.S.; Greathouse, D.V.; Providence, L.L.; Becker, M.D.; Koeppe, R.E. Importance of tryptophan dipoles for protein function: 5-Fluorination of tryptophans in gramicidin A channels. J. Am. Chem. Soc. 1998, 120, 5142–5146. [Google Scholar] [CrossRef]
  284. 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 unnaturalamino acids into the catalytic sites of squalene cyclase. J. Am. Chem. Soc. 2006, 128, 13184–13194. [Google Scholar] [CrossRef]
  285. Pless, S.A.; Millen, K.S.; Hanek, A.P.; Lynch, J.W.; Lester, H.A.; Lummis, S.C.; Dougherty, D.A. A cation-π interaction in the binding site of the glycine receptor is mediated by a phenylalanine residue. J. Neurosci. 2008, 28, 10937–10942. [Google Scholar] [CrossRef]
  286. He, T.; Gershenson, A.; Eyles, S.J.; Lee, Y.-J.; Liu, W.R.; Wang, J.; Gao, J.; Roberts, M.F. Fluorinated aromatic amino acids distinguish cation-π interactions from membrane insertion. J. Biol. Chem. 2015, 290, 19334–19342. [Google Scholar] [CrossRef]
  287. Ahern, C.A.; Eastwood, A.L.; Lester, H.A.; Dougherty, D.A.; Horn, R. A cation-π interaction between extracellular TEA and an aromatic residue in potassium channels. J. Gen. Physiol. 2006, 128, 649–657. [Google Scholar] [CrossRef] [PubMed]
  288. Santarelli, V.P.; Eastwood, A.L.; Dougherty, D.A.; Horn, R.; Ahern, C.A. A cation-π interaction discriminates among sodium channels that are either sensitive or resistant to tetrodotoxin block. J. Biol. Chem. 2007, 282, 8044–8051. [Google Scholar] [CrossRef] [PubMed]
  289. Granados, A.; Olmo, A.D.; Peccati, F.; Billard, T.; Sodupe, M.; Vallribera, A. Fluorous l-carbidopa precursors: Highly enantioselective synthesis and computational prediction of bioactivity. J. Org. Chem. 2018, 83, 303–313. [Google Scholar] [CrossRef]
  290. Jin, C.; Wei, L.; Ohgaki, R.; Tominaga, H.; Xu, M.; Okuda, S.; Okanishi, H.; Kawamoto, Y.; He, X.; Nagamori, S.; et al. Interaction of Halogenated Tyrosine/Phenylalanine Derivatives with Organic Anion Transporter 1 in the Renal Handling of Tumor Imaging Probes. J. Pharmacol. Exp. Ther. 2020, 375, 451–462. [Google Scholar] [CrossRef] [PubMed]
  291. Parsons, J.F.; Armstrong, R.N. Proton configuration in the ground state and transition state of a glutathione transferase-catalyzed reaction inferred from the properties of tetradeca(3-fluorotyrosyl)glutathione transferase. J. Am. Chem. Soc. 1996, 118, 2295–2296. [Google Scholar] [CrossRef]
  292. Xiao, G.; Parsons, J.F.; Armstrong, R.N.; Gilliland, G.L. Crystal structure of tetradeca-(3-fluorotyrosyl)-glutathione transferase. J. Am. Chem. Soc. 1997, 119, 9325–9326. [Google Scholar] [CrossRef]
  293. Narjes, F.; Koehler, K.F.; Koch, U.; Gerlach, B.; Colarusso, S.; Steinkühler, C.; Brunetti, M.; Altamura, S.; De Francesco, R.; Matassa, V.G. A designed P1 cysteine mimetic for covalent and non-covalent inhibitors of HCV NS3 protease. Bioorg. Med. Chem. Lett. 2002, 12, 701–704. [Google Scholar] [CrossRef]
  294. Han, W.; Hu, Z.; Jiang, X.; Wasserman, Z.R.; Decicco, C.P. Glycine α-ketoamides as HCV NS3 protease inhibitors. Bioorg. Med. Chem. Lett. 2003, 13, 1111–1114. [Google Scholar] [CrossRef]
  295. Zheng, B.; D’Andrea, S.V.; Sun, L.-Q.; Wang, A.X.; Chen, Y.; Hrnciar, P.; Friborg, J.; Falk, P.; Hernandez, D.; Yu, F.; et al. Potent inhibitors of hepatitis C virus NS3 protease: Employment of a difluoromethyl group as a hydrogen-bond donor. ACS Med. Chem. Lett. 2018, 9, 143–148. [Google Scholar] [CrossRef]
  296. Lemonnier, G.; Lion, C.; Quirion, J.C.; Pin, J.P.; Goudet, C.; Jubault, P. α-Amino-β-fluorocyclopropanecarboxylic acids as a new tool for drug development: Synthesis of glutamic acid analogs and agonist activity towards metabotropic glutamate receptor 4. Bioorg. Med. Chem. 2012, 20, 4716–4726. [Google Scholar] [CrossRef] [PubMed]
  297. Chen, L.; Wu, L.; Otaka, A.; Smyth, M.S.; Roller, P.P.; Burke, T.R.; den Hertog, J.; Zhang, Z. Why is phosphonodifluoromethyl phenylalanine a more potent inhibitory moiety than phosphonomethyl phenylalanine towards protein-tyrosine phosphatases. Biochem. Biophys. Res. Commun. 1995, 216, 976–984. [Google Scholar] [CrossRef] [PubMed]
  298. Isenegger, P.G.; Josephson, B.; Gaunt, B.; Davy, M.J.; Gouverneur, V.; Baldwin, A.J.; Davis, B.G. Posttranslational, site-directed photochemical fluorine editing of protein sidechains to probe residue oxidation state via 19F-nuclear magnetic resonance. Nat. Protoc. 2023, 18, 1543–1562. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Fluorine can play diverse roles when located within amino acid side chains.
Figure 1. Fluorine can play diverse roles when located within amino acid side chains.
Molecules 28 06192 g001
Scheme 1. General strategies for synthesizing side chain-fluorinated amino acids.
Scheme 1. General strategies for synthesizing side chain-fluorinated amino acids.
Molecules 28 06192 sch001
Scheme 2. Approaches for elaborating fluorinated amino acids into peptides and proteins.
Scheme 2. Approaches for elaborating fluorinated amino acids into peptides and proteins.
Molecules 28 06192 sch002
Figure 2. (a) Selected examples of fluorinated NMR tags; (b) 19F NMR spectrometry can be used to interrogate the structure, the conformational dynamics, or the binding events of proteins (e.g., mammalian prion protein, 38; bromodomain and plant homeodomain-containing transcription factor, 39).
Figure 2. (a) Selected examples of fluorinated NMR tags; (b) 19F NMR spectrometry can be used to interrogate the structure, the conformational dynamics, or the binding events of proteins (e.g., mammalian prion protein, 38; bromodomain and plant homeodomain-containing transcription factor, 39).
Molecules 28 06192 g002
Figure 3. (a) Amino acids with 18F-radiolabelled side chains; (b) early- vs. late-stage radiofluorination of proteins (human epidermal growth factor 2, 52; parathyroid hormone, 55).
Figure 3. (a) Amino acids with 18F-radiolabelled side chains; (b) early- vs. late-stage radiofluorination of proteins (human epidermal growth factor 2, 52; parathyroid hormone, 55).
Molecules 28 06192 g003
Figure 4. The oral bioavailability of amino acid-based medicinal agents 56 and 58 is enhanced in the side chain-fluorinated analogs 57, 59, and 60.
Figure 4. The oral bioavailability of amino acid-based medicinal agents 56 and 58 is enhanced in the side chain-fluorinated analogs 57, 59, and 60.
Molecules 28 06192 g004
Figure 5. Magainin (61) is highly susceptible to proteolysis. The fluorinated analogs 6263 (containing two and five hexafluoroleucine residues, respectively) are progressively more resistant to proteolysis.
Figure 5. Magainin (61) is highly susceptible to proteolysis. The fluorinated analogs 6263 (containing two and five hexafluoroleucine residues, respectively) are progressively more resistant to proteolysis.
Molecules 28 06192 g005
Figure 6. Fluorination of an amino acid side chain confers resistance to P450 metabolism.
Figure 6. Fluorination of an amino acid side chain confers resistance to P450 metabolism.
Molecules 28 06192 g006
Figure 7. Controlling the conformations of individual amino acid side chains.
Figure 7. Controlling the conformations of individual amino acid side chains.
Molecules 28 06192 g007
Figure 8. Influencing peptide secondary structure through side chain fluorination: (a) hairpin turn formation; (b) helix disruption.
Figure 8. Influencing peptide secondary structure through side chain fluorination: (a) hairpin turn formation; (b) helix disruption.
Molecules 28 06192 g008
Figure 9. Influencing protein tertiary structure through side chain fluorination.
Figure 9. Influencing protein tertiary structure through side chain fluorination.
Molecules 28 06192 g009
Figure 10. Controlling protein aggregation through side chain fluorination.
Figure 10. Controlling protein aggregation through side chain fluorination.
Molecules 28 06192 g010
Scheme 3. β,β,β-Trifluoroalanine (92) is a mechanism-based inhibitor of several pyridoxal phosphate (PLP) dependent enzymes.
Scheme 3. β,β,β-Trifluoroalanine (92) is a mechanism-based inhibitor of several pyridoxal phosphate (PLP) dependent enzymes.
Molecules 28 06192 sch003
Figure 11. Fluorine can modulate the intermolecular forces that amino acid side chains engage in.
Figure 11. Fluorine can modulate the intermolecular forces that amino acid side chains engage in.
Molecules 28 06192 g011
Figure 12. Conformational pre-organization can enhance the binding of a peptide ligand to its target (c.f. 100 vs. 101).
Figure 12. Conformational pre-organization can enhance the binding of a peptide ligand to its target (c.f. 100 vs. 101).
Molecules 28 06192 g012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Miles, S.A.; Nillama, J.A.; Hunter, L. Tinker, Tailor, Soldier, Spy: The Diverse Roles That Fluorine Can Play within Amino Acid Side Chains. Molecules 2023, 28, 6192. https://doi.org/10.3390/molecules28176192

AMA Style

Miles SA, Nillama JA, Hunter L. Tinker, Tailor, Soldier, Spy: The Diverse Roles That Fluorine Can Play within Amino Acid Side Chains. Molecules. 2023; 28(17):6192. https://doi.org/10.3390/molecules28176192

Chicago/Turabian Style

Miles, Samantha A., Joshua Andrew Nillama, and Luke Hunter. 2023. "Tinker, Tailor, Soldier, Spy: The Diverse Roles That Fluorine Can Play within Amino Acid Side Chains" Molecules 28, no. 17: 6192. https://doi.org/10.3390/molecules28176192

Article Metrics

Back to TopTop