Next Article in Journal
PPARδ Agonist GW501516 Suppresses the TGF-β-Induced Profibrotic Response of Human Bronchial Fibroblasts from Asthmatic Patients
Next Article in Special Issue
Oligomeric State and Holding Activity of Hsp60
Previous Article in Journal
Advances in the Anti-Tumor Activity of Biflavonoids in Selaginella
Previous Article in Special Issue
A Novel Ambroxol-Derived Tetrahydroquinazoline with a Potency against SARS-CoV-2 Proteins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

FDA-Approved Fluorinated Heterocyclic Drugs from 2016 to 2022

Department of Biological, Chemical and Pharmaceutical Sciences and Technologies (STEBICEF), University of Palermo, Viale delle Scienze, Ed. 17, 90128 Palermo, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
This review is dedicated to the loving memory of Prof. Nicolò Vivona (1939–2020), who introduced us to the chemistry of fluorinated heterocycles.
Int. J. Mol. Sci. 2023, 24(9), 7728; https://doi.org/10.3390/ijms24097728
Submission received: 31 March 2023 / Revised: 19 April 2023 / Accepted: 21 April 2023 / Published: 23 April 2023
(This article belongs to the Collection Feature Papers in Molecular Biophysics)

Abstract

:
The inclusion of fluorine atoms or heterocyclic moiety into drug structures represents a recurrent motif in medicinal chemistry. The combination of these two features is constantly appearing in new molecular entities with various biological activities. This is demonstrated by the increasing number of newly synthesized fluorinated heterocyclic compounds among the Food and Drug Administration FDA-approved drugs. In this review, the biological activity, as well as the synthetic aspects, of 33 recently FDA-approved fluorinated heterocyclic drugs from 2016 to 2022 are highlighted.

1. Introduction

The presence of many different heterocyclic rings in natural products, such as alkaloids, vitamins, antibiotics, peptides, etc., prompted the introduction of structural motifs into synthetic drugs [1,2].
Therefore, heterocycles are considered to be prominent scaffolds for the synthesis of biologically active compounds and prospective drugs [3,4]. In fact, it is estimated that heterocyclic moieties are present in around 85% of bioactive compounds [5]. On the other hand, in the second half of 20th century, another fundamental tool for drug design was introduced with the incorporation of fluorine atoms into drugs [6,7,8]. Since the introduction of the first fluorocorticosteroid, fludrocortisone, in 1954 [9], the fluorinated drugs market has exponentially evolved, with 20% of those on the market being fluorinated drugs and around 30% of fluorinated drugs being blockbuster pharmaceuticals, such as Lipitor, Fluoxetine, Linezolid or Fluticasone [8]. To date, more than 300 fluorinated pharmaceuticals have been approved for use as drugs [6]. The success of the introduction of fluorine atoms is linked to the peculiar physicochemical properties of the C-F bond [10], which are the high bond strength, polarity and minimal steric hindrance of fluorine [11], combined with a general metabolic stability that, nevertheless, is an issue that is currently under exploration [12].
The introduction of fluorine, for example, allows researchers to easily modulate the pKa of neighboring functionalities, improving the bioavailability and affinity to specific receptors [13,14].
In addition, the monofluorination or trifluoromethylation of alkyl groups decreases the drug lipophilicity due to the strong electron-withdrawing capabilities of fluorine. On the other hand, fluoro-arenes are more lipophilic due to the low polarizability of the C-F bond [15]. In addition, the presence of a fluorine atom can also enhance the membrane’s permeability [16].
The importance of fluorinated compounds is also linked to their use as diagnostic tools within imaging techniques such as 19F-MRI and 18F-PET [17,18].
The direct link between fluorinated moieties and heterocycles led to the formation of the sub-class of fluorinated heterocycles, which combine the strength of these two fundamental scaffolds in modern medicinal chemistry. This important class of fluorinated pharmaceuticals includes some of the selected examples of FDA-approved drugs reported in Figure 1. Among these compounds, there have been several game changers over the last decades, such as fluorouracil, the class of fluoroquinolones, sitagliptin and fluorodeoxyglucose, just to mention a few [19].
In this review, recent advances in the field of fluorinated heterocyclic drugs are presented, discussing FDA-approved molecules from 2016 to 2022. The molecules considered in this article are limited to those with a fluorinated group directly linked to the heterocyclic ring. The biological targets and the therapeutic indications are presented together with synthetic details. The fluorination strategy and influence of the fluorinated moiety on bioactivity are also discussed. The sections are reported in chronological order, starting with the most recently approved drugs; for each section, the compounds are presented in alphabetical order.

Introduction of Fluorine Atoms in Organic Molecules

Fluorinated starting materials used as precursors to obtain fluorinated, approved drugs can present mono-, di- and trifluoro alkyl groups; the last ones can be generally introduced via building blocks such as trifluoroacetate, trifluoroethylamine, trifluoro and ethyl triflate. Furthermore, several starting materials used for this purpose are formed by aromatic or heterocyclic rings bearing F atoms, such as (poly)fluorobenzoic acid, fluoro- or trifluoromethylpyridines, just to cite a few of them (see below).
Some of the main processes used for the introduction of F atoms are summarized in Figure 2. The introduction of C-F or CF2 groups is achieved through the nucleophilic fluorination of electrophiles; some examples involving dithiane or sulphonate formation are reported in Figure 2a,b [20,21].
Trifluoromethylated starting materials such as trifluoroacetic acid are industrially prepared in excellent yields by the electrochemical fluorination of acetyl chloride or acetic anhydride in anhydrous hydrogen fluoride, followed by the hydrolysis of the resulting trifluoroacetyl fluoride (Figure 2c) [22]. Fluorination methods of arenes include traditional nucleophilic substitution (Figure 2d) and transition-metal-catalyzed nucleophilic fluorination or deoxofluorination [21]. For the introduction of trifluoromethyl groups, substrates can be trifluoromethylated by employing electrophilic trifluoromethylating reagents [23], such as Togni reagents [24], and S-(trifluoromethyl)dibenzothiophenium salts [25], or lower cost reagents such as CF3I or CF3H, which are favored for industrial processes, while the use of the solid and bench-top-stable reagents such as NaSO2CF3 in radical trifluoromethylations for the trifluoromethylation of electron-rich arenes avoids perfluoroalkylations (Figure 2e) [21].

2. FDA-Approved Drugs in 2022

In 2022, the FDA approved 37 new therapeutic and diagnostic products [26]. Monoclonal antibodies (mAbs) continue to be one of the most widely licensed groups of biological therapies. Nevertheless, 22 of them are novel chemical entities (NCEs), 14 of which contain fluorine atoms and nitrogen heterocycles [27]. Lenacapavir (Scheme 1 and Scheme 2) and Oteseconazole (Scheme 3), two new approved drugs released in the past year, combine these two key characteristics.

2.1. Lenacapavir

Lenacapavir (SUNLENCA®) is a human immunodeficiency virus type 1 (HIV-1) capsid inhibitor developed by Gilead Science Inc., and it is administered in cases when conventional antiretroviral therapies are ineffective. The mechanism of action is totally different from that of other antivirals used for the treatment of HIV-1. Indeed, Lenacapavir establishes several hydrophobic and electrostatic interactions with capsid subunits (CA1 and CA2). For example, the difluorobenzyl group can be stabilized inside a hydrophobic pocket of the CA1 N-terminal domain. This entails interference with virus life cycle processes where CA is involved, such as reverse transcription, nuclear import and integration [28,29].
The synthetic method to obtain Lenacapavir is divided into three steps and is reported in Scheme 1 and Scheme 2 [30].
The Claisen condensation of 1 using a strong base, such as lithium bis(trimethylsilyl)amide and ethyl 2,2,2-trifluoroacetate, leads to enolate 2, presenting a CF3 group. At this point, pyrazole ring 3 formation occurs due to the addition of an ethyl hydrazinoacetate salt.
Intermediate 4 is obtained through N-hydroxyphtalimide-catalyzed selective oxidation, followed by saponification with NaOH. To obtain building block 6, the desulfurative fluorination of dithiolane 5 takes place, followed by supercritical fluid chromatography (SFC).
The synthesis proceeds with the construction of building block 9. To acquire substituted indazole core 8, hydrazine is combined with fluorobenzonitrile 7, and then a trifluoroethyl group is introduced via a substitution at position one.
A cross-coupling reaction between bis(pinacolato)diboron and 8 in the presence of a palladium/triphenylphosphine catalyst and potassium propionate creates 9. In order to produce Lenacapavir, deprotection preceded by the formation of an amide bond between carboxylic acid 6 and ammine 14 must occur.
Intermediate 14 is obtained via two palladium-catalyzed coupling reactions between fluorinated compound 10 and amine 13, followed by a protection with methanesulfonyl chloride on the amino group linked to the indazole ring.

2.2. Oteseconazole

Oteseconazole (VIVJOA™) is an antifungal agent that was released by Mycovia Pharmaceuticals and is administered to reduce the incidence of recurrent vulvovaginal candidiasis (RVVC). It affects the integrity of the cell membrane of pathogenic strains of candida by interacting with cytochrome P450 (CYP51) [31,32].
Oteseconazole’s selectivity for fungal metalloenzyme CYP51 is provided by the tetrazole moiety. In turn, the heterocyclic residue is connected through a metabolically resistant difluoro methyl linker with a substituted phenyl trifluoroethyl ether. The synthesis is presented in Scheme 3 [33].
Starting with pyridine 15, ethyl bromodifluoroacetate and 2,4-difluorobromobenzene are used to introduce the CF2 linker and 2,4-difluorobenzene group, respectively, producing 16. Through a diazomethane-mediated epoxidation reaction, intermediate 17 is obtained and reacts with 4-(trifluoromethoxy)phenylboronic acid 18 via a Pd catalyzed Suzuky–Miyaura coupling reaction, producing 19. In the next two steps, the introduction of the triazole ring occurs via nucleophilic attack, leading to the opening of the epoxy ring. Oteseconazole is finally achieved as a single enantiomer by chiral preparative HPLC.

3. FDA-Approved Drugs in 2021

In 2021, 50 new drugs were approved. Thirty-three small molecules with 10 fluorinated compounds and 28 heterocyclic compounds are included in this list, together with fluorinated heterocycles Atogepant, Piflufolastat, Sotorasib, Umbralisib, and Vericiguat (Scheme 4, Scheme 5, Scheme 6, Scheme 7 and Scheme 8), as discussed below [34].

3.1. Atogepant

Atogepant (Qulipta™) is a novel drug designed by Abb Vie for the preventive treatment of migraines in adults. Developed as a calcitonin gene-related peptide (CGRP) receptor antagonist, this neuropeptide and its receptors are located in the trigeminal nerves involved in pain sensations [35]. The antagonist’s high affinity to the receptor is increased due to a 2,3,6 fluoro substitution on the phenyl moiety. The 2,2,2-trifluoethyl group masking the piperidinone ring improved the pharmacokinetic and pharmacodynamic characteristics more than other gepant drugs do. In addition, fluorine atoms could also be associated with a lower hepatotoxicity [36]. The method for Atogepant synthesis is reported in Scheme 4 [37]. 1-(2,3,6-trifluorophenyl)propan-2-one 20 is alkylated to 21 via N-Boc-iodoserine-OMe. Piperidine intermediate 22 is a result of a reductive amination with 2,2,2-trifluoroethanamine and sodium triacetoxyborohydride as a reducing agent, followed by cyclization and optical resolution using normal-phase liquid chromatography (NPLC). In order to generate the first building block, 23, deprotection of the amine group via hydrochloric acid takes place. The second intermediate, 26, is achieved by performing a previously patented procedure on 24 [38]. Azospiro bispyridine 24 undergoes diazotation–iodination process via sodium nitrite in the presence of p-toluensulfonic acid and potassium iodide. Ester 25, obtained via palladium-catalyzed carbonylation, is subsequently saponified to obtain 26. Finally, a coupling reaction with aminopiperidinone 23 and carboxylic acid 26 is carried out to allow the formation of an amide bond leading to the final product, Atogepant.
Scheme 4. Synthesis of Apogent.
Scheme 4. Synthesis of Apogent.
Ijms 24 07728 sch004

3.2. Piflufolastat F 18

Piflufolastat F 18, commercially known as PYLARIFY by Progenics Pharmaceuticals Inc., is a diagnostic imaging agent radiolabeled with 18F isotope, which detects prostate-specific membrane antigen (PSMA) via positron emission tomography (PET). It was approved on May 2021 by the FDA as a radioactive diagnostic tool, thanks to which it is possible to obtain accurate and early information on prostate cancer metastases, even in those patients with low prostate-specific antigen (PSA) levels [39,40].
Piflufolastat synthesis is described in Scheme 5 [40].
Piflufolastat is the result of the deprotection of the carboxyl group with TFA, preceded by a nucleophilic substitution with urea derivate 27 on 6-[18F]Fluoro-nicotinic acid-2,3,5,6-tetrafluoro-phenyl ester 28, synthetized according to the previously reported procedure [41].
Scheme 5. Synthesis of Piflufolast 18F.
Scheme 5. Synthesis of Piflufolast 18F.
Ijms 24 07728 sch005

3.3. Sotorasib

The brand name for the new drug marketed by Amgen is LUMAKRASTMTM. The pharmacologically active agent, Sotorasib, it is an RAS small GTPase inhibitor used to treat colorectal cancer and non-small-cell lung cancer brought on by the KRASG12C oncogene [42]. To improve the pharmacokinetic properties such as oral bioavailability, azaquinazolinone is designed with a fluorine atom on C6 carbon, a fluorophenol residue on C7 and a nitrogen atom instead of C8 carbon [43]. The Sotorasib synthetic process is described in Scheme 6 [44].
Starting with 2,6-dichloro-5-fluoronicotinic acid 29, acyl chloride is obtained through the use of oxalyl chloride; this is then converted to the corresponding amide, 30. The formation of compound 31 is carried out through the reaction between nicotinamide 30 and 2-isopropyl-4-methylpyridin-3-amine. Compound 31 is then treated with potassium hexamethyldisilazane to drive cyclization and produce the duly substituted 2,4-dihydroxypyrido [2,3-d]pyrimidine ring 32. At this point, the chlorination reaction produces intermediate 33, which, in turn, can be combined with a Boc-protected methylpiperazine to produce a selective amination. The resulting compound 34 is reacted with organotrifluoroborate salt via a Suzuki−Miyaura cross-coupling reaction to attach fluorophenol moiety 35. Deprotection and amidation of nitrogen of the piperazine ring eventually produces Sotorasib. In 2022, the same Amgen group developed a commercial manufacturing process, in which they improved several synthetical steps, starting with compound 32 [45].
Scheme 6. Synthesis of Sotorasib.
Scheme 6. Synthesis of Sotorasib.
Ijms 24 07728 sch006

3.4. Umbralisib

Umbralisib is sold under the brand name Ukuoniq and was developed by TG Therapeutics. It was approved in February 2021 for the treatment of marginal zone lymphoma (MZL) and follicular lymphoma (FL) [46]. The mode of action of Umbralisib is related to the inhibition of kinase PI3K-delta and casein kinase CK1-epsilon. Umbralisib contains a 6-fluoro-chromen-4-one central heterocyclic core and two other fluorophenyl groups. The synthesis was disclosed in a patent in 2014 and is presented in Scheme 7 [47]. Fluorinated chromen-4-one ring 38 is constructed, starting with 3-fluorophenylacetic acid 36, which, after conversion into chloride and subsequent acylation of 4-fluoroanisole, yields compound 37. The treatment of phenol 37 with propionic anhydride produces 38 via acylation and subsequent cyclocondensation. The radical bromination of the methylene group with NBS yields 39. The following steps result in the obtainment of racemic alcohol 40 after the hydrolysis of 39, as well as the subsequent formation of two enantiomers, 42 and 44. S enantiomer 42 could be selectively obtained via the stereoselective reduction of ketone 41 with R Alpine borane, which is obtained by means of the Swern oxidation of racemic 40. R enantiomer 44 was obtained via a Mitsunobu reaction with 4-chlorobenzoic acid and DEAD, followed by the hydrolysis of ester 43. Alcohol 44 is coupled with pyrazolopyrimidine 49, again under Mitsunobu conditions, to acquire Umbralisib as a single enantiomer. Compound 49 is obtained via the Suzuki coupling of iodopyrazolopyrimidine 48 with aryl pinacolborane 47. Compound 47 is obtained in two steps using 4-bromo-2-fluoro-phenol 45.
Scheme 7. Synthesis of Umbralisib.
Scheme 7. Synthesis of Umbralisib.
Ijms 24 07728 sch007

3.5. Vericiguat

Vericiguat is sold under the brand name Verquvo and was developed by Bayer AG and Merck & Co. It was approved in January 2021 to reduce the risks of cardiovascular death and heart failure [48].
Vericiguat is a guanylate cyclase (sGC) stimulator with a 1H-pyrazolo [3,4-b]pyridine core bearing a fluorine atom at C-5. The presence of the fluorine atom increases the metabolic stability and induces lower clearance. The method for the synthesis of Vericiguat is reported in Scheme 8 [48].
Tetrafluoropropanol 50, the starting fluorinated building block, is converted in two steps into morpholino derivative 51, and then into morpholinium cation 52 after methylation with methyl methanesulfonate. Compound 53 is obtained after a treatment with NaOH, which induces the elimination of the first fluorine atom as HF.
Other two fluorine atoms from the difluoromethyl group are lost during hydrolysis into 2-fluoroacrolein derivative 54. α,β-unsaturated aldehyde 54 reacts with aminopyrazole 55 under acidic cyclization conditions, allowing the introduction of the 5-fluorine atom into the 1H-pyrazolo [3,4-b]pyridine core of derivative 56. Ester 56 is then converted in three steps into amidine 59, via amide 57 and nitrile 58. The C-3 pyrimidine ring is then constructed with a condensation between 59 and hydrazonomalonitrile 60. Using compound 61, the synthesis of Vericiguat is completed in two steps via the catalytic hydrogenation of the diazo moiety of 61 to triaminopyrimidine 62, and finally, via the formation of the carbamate group of the target compound after a treatment with methyl chloroformate.
Scheme 8. Synthesis of Vericiguat.
Scheme 8. Synthesis of Vericiguat.
Ijms 24 07728 sch008

4. FDA-Approved Drugs in 2020

In 2020, the FDA approved 53 new molecular entities, including 34 small molecules and 4 diagnostic agents [49]. Thirty-one out of thirty-four molecules contain at least one heterocyclic ring, and eleven out of thirty-four molecules contain at least one fluorine atom. In the following paragraph, four heterocyclic compounds bearing a fluorinated moiety directly linked to the ring are reported (Scheme 9, Scheme 10, Scheme 11 and Scheme 12). Additionally, the approved 18F-containing diagnostic agent, Tauvid, is presented (Scheme 13).

4.1. Berotralstat

Berotralstat is sold under the brand name Orladeyo and was developed by BioCryst Pharmaceuticals. It was approved in December 2020 to treat Hereditary Angioedema (HAE) attacks [50]. Berotralstat is a selective inhibitor of plasma kallikrein, bearing a trifluoromethylpyrazole moiety. Another fluorine is present on the central phenyl ring. The patented synthetic approach to acquiring this drug is reported below (Scheme 9) [51].
The trifluoromethylpyrazole portion of compound 65 is constructed through the condensation of trifluoro β-diketone 63 and arylhydrazine 64 in acetic acid. Cyanopyrazole 65 is then reduced into amine 66, using LiAlH4, which is in turn protected as N-Boc derivative 67 during the successive oxidation of the furan ring to yield acid 68. Coupling between amine 69 and amide 70 is performed using bromotris-pyrrolidino-phosphonium hexafluorophosphate (PyBrOP) as an activating agent. The formation of amine 71 occurs after the treatment of 70 with thionyl chloride, and then cyclopropanemethylamine. Berotralstat was finally obtained as a single enantiomer after acidic Boc-deprotection and chiral SFC resolution.
Scheme 9. Synthesis of Berotralstat.
Scheme 9. Synthesis of Berotralstat.
Ijms 24 07728 sch009

4.2. Cedazuridine

Cedazuridine, in combination with decitabine, is sold under the brand name Inqovi and was developed by Otsuka Pharma. It was approved in July 2020 for the treatment of myelodysplastic syndromes (MDS) and chronic myelomonocytic leukemia (CMML), reducing the risk of progression of secondary acute myeloid leukemia (sAML) [52].
Cedazuridine is a cytidine deaminase inhibitor that is able to improve the oral bioavailability of decitabine, avoiding its degradation in the gastrointestinal tract. The presence of two fluorine atoms at the ribose ring increase the level of metabolic stability under acidic conditions, improving the pharmacokinetic profile via unfluorinated analogs, retaining the same binding mode of unfluorinated tetrahydrouridines [53]. The synthesis of Cedazuridine is performed in two steps, starting with the analogue, Gemcitabine (Scheme 10). The Rh/C catalytic hydrogenation of Gemcitabine produces compound 76, which is reduced into a mixture of isomers containing Cedazurine and its epimer as major products using NaBH4. The difluorotetrahydrofuran ring of Gemcitabine is synthesized via a Reformatzky reaction of fluorinated bromoacetate 72 with D-glyceraldehyde acetonide 73 to furnish 74 as a mixture of anti/syn diasteroisomers at a 3:1 ratio. The hydrolysis/lactonization of 74 into 75 is performed with Dowex 50 resin. Lactone 75 is then converted into Gemcitabine in four steps [54].
Scheme 10. Synthesis of Cedazuridine.
Scheme 10. Synthesis of Cedazuridine.
Ijms 24 07728 sch010

4.3. Pralsetinib

Pralsetinib is sold under the brand name Gavreto and was developed by Blueprint Medicines [55]. It was approved in September 2020 for the treatment of metastatic fusion-positive non-small-cell lung cancer [56]. Pralsetinib is REarranged during transfection (RET) inhibitor and it is the first-in-class specific RET inhibitor with more selectivity than other kinases have. The presence of the 4-fluoropyrazolo group allows a different binding mode on the BP-II pocket, which is crucial for high-affinity binding and to avoid resistance from gatekeeper mutations [56].
4-Fluoropyrazole 77 gives nucleophilic displacement of bromopyridine 78 under basic conditions, yielding pyrazolylpyridine 79. The stereoselective reductive amination to hydrochloride 82 is accomplished by means of the condensation of the acyl group of 79 with chiral sulfinamide 80, followed by the reduction with L-Selectride and acidic hydrolysis of sulfinamide 81. The latter fluorinated building block is coupled with acid 83 (as mixture of diastereoisomers) using PyBop as activating agents. A final chiral SFC resolution produces Pralsetinib as a single enantiomer. The method for the patented synthesis of Pralsetinib is described in Scheme 11 [57].
Scheme 11. Synthesis of Pralsetinib.
Scheme 11. Synthesis of Pralsetinib.
Ijms 24 07728 sch011

4.4. Selumetinib

Selumetinib is sold under the brand name Koselugo and was developed by AstraZeneca. It was approved in April 2020 for the treatment of neurofibromatosis type 1 (NF1) [58]. Selumetinib, characterized by the presence of a 4-fluorobenzimidazole core, is a mitogen-activated protein kinase (MEK) inhibitor that is able to target the Raf-MEK-ERK signaling pathway. The synthesis of Selumetinib is described in Scheme 12 [59].
Trifluorobenzoic acid 84 was employed as starting fluorinated building block for the initial construction of the fluorobenzimidazole ring. The nitration of 84 into 85 and the nucleophilic aromatic displacement of fluoride in nitro-activated derivative 85 with ammonia led to the acquisition of compound 86. Treatment with trimethylsilyldiazomethane (TMS-CHN2) converts acid 86 into methyl ester 87, which, in turn, reacts with aniline in xylene at 125 °C, causing the nucleophilic displacement of a second fluorine atom into 88. The latter nitro derivative is reduced to o-diamminobenzene 89 via iron and ammonium chloride. The benzimidazole ring formation on 90 is performed with formamidine acetate (FAA) in EtOH at 80 °C. The halogenation of the anilino portion of 90 in two consecutive steps produces 91, with the NBS-mediated introduction of bromine in the para position, and then, 92, with ortho-chlorination performed with NCS. The methylation of 92 with methyliodide, employing K2CO3 in DMF, occurs at N-1, yielding regio-isomer 93. Basic hydrolysis into acid 94 is followed by coupling with O-(2-(vinyloxy)ethyl)hydroxylamine in the presence of EDC and HOBt. Hydroxamic acid derivative 95 is finally deprotected via the acidic hydrolysis of the vinyl ether portions, producing Selumetinib.
Scheme 12. Synthesis of Selumetinib.
Scheme 12. Synthesis of Selumetinib.
Ijms 24 07728 sch012

4.5. Tauvid

Tauvid, also named Flortaucipir F18, was developed by Eli Lilly and was approved in May 2020 as a positron emission tomography (PET) imaging probe for Alzheimer’s disease (AD) [60].
Tauvid is the first approved tracer able to bind tau protein [61]. The [18F]fluoropyridine lateral ring is the radioactive portion of this probe. Two developed synthetic procedures are described in Scheme 13.
Scheme 13. Synthesis of Tauvid.
Scheme 13. Synthesis of Tauvid.
Ijms 24 07728 sch013
The first synthetic approach is based on the nucleophilic displacement of the nitro group of precursor 96 with [18F]fluoride and using Kryptofix 222 (K222) as a phase transfer catalyst (PTC) [61]. This approach suffers from some drawbacks related to trace purity; therefore, a different synthetic approach was developed starting with N-Boc protected cation 97, which undergoes nucleophilic displacement to produce radioactive fluorine, followed by acidic Boc removal [62]. This synthetic sequence allows the obtainment of Tauvid in higher yields and purity. Interestingly, this synthesis represents the only example of late-stage fluorination among all the molecules considered in this review. Obviously, the short half-life of 18F forces researchers to follow this peculiar synthetic approach.

5. FDA-Approved Drugs in 2019

In 2019, the FDA approved 48 new molecular entities, including 33 small molecules and 3 diagnostic agents [63]. Among the small molecules, 28 out of 33 contain at least one heterocyclic ring, and 11 out of 33 contain at least one fluorine atom [64,65]. In the following paragraph, four heterocyclic compounds bearing a fluorinated moiety directly linked to the ring are reported (Scheme 14, Scheme 15, Scheme 16 and Scheme 17).

5.1. Alpelisib

Alpelisib is sold under the brand name Rinvoq and was developed by Novartis. It was approved in May 2019 for the treatment of advanced or metastatic breast cancer [66]. Alpelisib is a phosphatidylinositol 3-kinase (PI3K) inhibitor possessing a trifluoro-t-butyl group at position two of the pyridine ring [67]. The presence of the fluorinated moiety induces a higher metabolic stability and excellent oral bioavailability.
Furthermore, the fluorinated group is responsible for high affinity toward the PI3K binding site due to the hydrogen bond with K802, as revealed by X-ray data [68]. Alpelisib synthesis is performed using different approaches [69,70]. One method is shown in Scheme 14 [68].
Fluorinated acid 98 is first converted into the corresponding chloride, 99, via oxalyl chloride under reflux. The acylation of the methyl group of enone 100 is performed at −78 °C using LiHMDS as a strong base. The intermediate diketone is directly cyclized to pyran-4-one 101 after a treatment with trifluoroacetic acid (TFA). The reaction of 101 with ammonium hydroxide produces fluorinated pyridin-4-one 102 via an ANRORC-like reaction. The treatment with POBr3 yields bromopyridine 103, which is, in turn, coupled with acetylaminothiazole 104 using Pd(OAc)2 in a CH activation process. The resulting coupled product, 105, is hydrolyzed using HCl into corresponding amine 106. The treatment with carbonyldiimidazole (CDI) leads to intermediate 107, which is then converted into Alpelisib after a treatment with S-prolinamide.
Scheme 14. Synthesis of Apelisib.
Scheme 14. Synthesis of Apelisib.
Ijms 24 07728 sch014

5.2. Lemborexant

Lemborexant is sold under the brand name Dayvigo and was developed by Purdue Pharma L. Lemborexant is a dual orexin receptor antagonist, with fluorine in the position five of the pyridine moiety [71]. Another fluorine is also present at position three of the central phenyl ring. It was approved in December 2019 for the treatment of insomnia [72].
The presence of each fluorine is crucial to achieve high in vitro binding affinity, good solubility and a good pharmacological profile, as evidenced during the discovery process with the screening of different fluorination patterns [73].
The synthesis of Lemborexant starts with 3-fluorobenzyl cyanide 108, which is converted into chiral cyclopropane derivative 109 in three steps (Scheme 15). Dimethylpyrimidine derivative 110 is used in a reaction under Mitsunobu conditions to obtain ether 111. The conversion of the primary alcoholic function of 111 into corresponding carboxylic acid 112 is performed in two steps with Swern oxidation, followed by Pinnick oxidation. The final production of Lemborexant requires the coupling of acid 112 with 2-amino-5-fluoropyridine 113 using HATU as an activating agent.
Scheme 15. Synthesis of Lemborexant.
Scheme 15. Synthesis of Lemborexant.
Ijms 24 07728 sch015

5.3. Pexidartinib

Pexidartinib is sold under the brand name Turalio and was developed by Daiichi Sankyo Inc. It was approved in August 2019 for the treatment of symptomatic tenosynovial giant cell tumor (TGCT) [74]. Pexidartinib is a tyrosine kinase inhibitor with selective efficacy for colony-stimulating factor (CSF) receptor; thus, it hampers the binding of CSF1 to CSF-receptor 1 (CSF1R).
Three steps in the synthesis of Pexidartinib at the kilogram scale are shown in Scheme 16 [75]. The base-mediated reaction of azaindole 114 at position 3, over aldehyde 115, in the presence of tetrabutylammonium hydrogen sulphate (TBAHS) yields compound 116. The dehydroxylation of 116 with triethylsilane (TES), followed by the Boc-deprotection of the 2-aminopyiridino moiety with TFA, produce compound 117. Pexidartinib is obtained by means of the reductive amination of 117 with 6-(trifluoromethyl)nicotinaldehyde 118 using TES as a reducing agent.
Scheme 16. Synthesis of Pexidartinib.
Scheme 16. Synthesis of Pexidartinib.
Ijms 24 07728 sch016

5.4. Ubrogepant

Ubrogepant is sold under the brand name Ubrelvy and was developed by Allergan. It was approved in December 2019 for the treatment of migraines with or without an aura in adults [76]. Ubrogepant is an effective calcitonin gene-related peptide (CGRP) receptor antagonist, bearing a chiral N-trifluoroethylpiperidinone ring, but its mechanism of action is still unknown. The synthesis of Ubrogepant was patented in 2012 and is reported in Scheme 17 [77]. The synthesis of fluorinated chiral amine 123, starting with phenylacetone 120, which alkylates with iodide 119 in the presence of Cs2CO3 as a base, produces derivative 121 in three steps. The reductive amination of the latter substance using trifluoroethylamine in the presence of sodium triacetoxyborohydride, leads to pyridinone 122 as a single enantiomer, after chiral resolution via normal-phase liquid chromatography (NPLC).
The treatment with HCl deprotects the N-Boc group, yielding 123. The coupling of amine 123 with acidic spyro-subunit 128 (prepared as outlined in Scheme 17) using BOP as an activating agent yields Ubrogepant as a single enantiomer via SFC.
Scheme 17. Synthesis of Ubrogepant.
Scheme 17. Synthesis of Ubrogepant.
Ijms 24 07728 sch017

6. FDA-Approved Drugs in 2018

In 2018, the FDA approved a collection of 59 new molecular entities, including 39 small molecules [78]. Thirty-two out of thirty-eight molecules contain at least one heterocyclic ring, and seventeen out of thirty-eight molecules contain at least one fluorine atom [16]. In the following paragraphs, eight heterocyclic compounds bearing a fluorinated moiety directly linked to the ring are reported (Scheme 18, Scheme 19, Scheme 20, Scheme 21, Scheme 22, Scheme 23, Scheme 24 and Scheme 25).

6.1. Apalutamide

Apalutamide is sold under the brand name Erleada and was discovered by employees of the University of California and developed by Janssen [79]. It was approved in February 2018 for the treatment of prostate cancer (PC) [80]. Apalutamide is a non-steroidal oral androgen receptor inhibitor, presenting a trifluoromethylpyridine moiety linked to the central thiohydantoin core [81]. One of the initial patented Apalutamide synthesis procedures is reported below (Scheme 18) [82].
Chloro-trifluoromethylpyridine 129 is treated with water to induce the nucleophilic displacement of chlorine to obtain pyrimidone 130. This compound can be easily nitrated into 131, and again, converted into corresponding chloropyridine 132 via a treatment with PCl5/POCl3. The hydrogenation on Raney Ni into amine 133 is then followed by N-Boc protection with Boc-anhydride, yielding 134. A Sandmeyer reaction causes cyanation to produce derivative 135, which is subsequently deprotected into aminopyrimidine 136 using TFA. Isothiocyanate 137 is then obtained via the treatment the 136 using thiophosgene. Apalutamide is finally obtained from the construction of the thiohydantoin ring via the reaction of 137 with isocyanide 138 under microwave irradiation. The synthesis of crystalline forms of Apalutamide has recently been reviewed [83].
Scheme 18. Synthesis of Aputalamide.
Scheme 18. Synthesis of Aputalamide.
Ijms 24 07728 sch018

6.2. Baloxavir Marboxil

Baloxavir marboxil is sold under the brand name Xofluza and was developed by Shionogi [84]. It was approved in October 2018 for the treatment of acute uncomplicated influenza types A and B [85].
Baloxavir marboxil is a cap-dependent endonuclease inhibitor characterized by the presence of two fluorine atoms on the 6,11-dihydrodibenzo[b,e]thiepine ring [86]. The patented synthesis of Baloxavir marboxil is depicted in Scheme 19 [87].
Scheme 19. Synthesis of Baloxavir marboxil.
Scheme 19. Synthesis of Baloxavir marboxil.
Ijms 24 07728 sch019
Difluorobenzoic acid 139 is initially formylated via LDA/DMF at low temperature, yielding cyclization product 140. A reaction with thiophenol under acidic conditions produces phenylthiolphthalide 141. The reductive breaking of the C-O bond mediated by the action 1,1,3,3-Tetramethyldisiloxane yields thioether 142, which undergoes intramolecular cyclization by heating at 120 °C. The obtained ketone 143 is reduced into alcohol 144 using NaBH4, and then coupled with chiral compound 145 under acidic conditions to produce compound 146. The deprotection of hexyl ether into 147, and a final reaction with chloromethyl methyl carbonate allows the obtainment of Baloxavir marboxil.

6.3. Binimetinib

Binimetinib is sold under the brand name Mektovi and was developed by Array Biopharma. It was approved in June 2018 for the treatment of metastatic BRAF V600E/K-positive advanced melanoma in association with Encorafenib [88].
Binimetinib is a potent, selective, non-ATP competitive allosteric inhibitor of MEK1 and MEK2, with a fluorobenzimidazole moiety similar to that of Selumetinib (Scheme 12) [89].
As for the analogue, Selumetinib, the synthetic route of Binimetinib is based on ester 87 (Scheme 20) [90]. The nucleophilic displacement of 2-fluoroaniline 148 to obtain derivative 149 is followed by hydrogenation in the presence of formic acid to directly yield benzimidazole 150. NBS-mediated bromination and methylation at N-1 give compounds 151 and 152, respectively. Ester 117 is, therefore, hydrolyzed using NaOH, and the resulting acid, 153, after activation using EDC/HOBt is converted into hydroxamic acid 155 upon a reaction with O-alkyl hydroxylamine 154. Binimetinib is finally obtained by the acidic hydrolysis of the vinyl ether group of 155.
Scheme 20. Synthesis of Binimetinib.
Scheme 20. Synthesis of Binimetinib.
Ijms 24 07728 sch020

6.4. Doravirine

Doravirine is sold under the brand name Pifeltro and was developed by Merck. It was approved in August 2018 for the treatment of human immunodeficiency virus 1 (HIV-1) infections [91]. Doravirine is a non-nucleoside reverse transcriptase inhibitor, presenting improved ADME properties due the presence of a CF3-substituted pyridone central ring. In fact, the presence of this strong electron-withdrawing group is correlated with a longer elimination half-life in rats and dogs compared to that of unfluorinated analogs [92]. The patented method of the synthesis of Doravirine is shown in Scheme 21 [93].
Fluorinated building block 2-chloro-3-fluoro-4-(trifluoromethyl)pyridine 156 is first used in a reaction with phenol 157 using K2CO3 as a base to induce fluoride displacement and obtain ether 158. The hydrolysis of chloropyridine 158 into pyridinone 159 is then performed via a treatment with KOH. The cyanation of the C-Br bond of the phenyl portion with CuCN yields substrate 160, which reacts under basic condition with chloromethyltriazole 161 to produce pyridone N-alkylation product 162. Doravirine is finally obtained by the regio-selective methylation of the 1,2,4-triazole ring at N-4 with methyliodide in DMF in the presence of K2CO3 as a base. Other synthetic approaches have recently been reviewed [16].
Scheme 21. Synthesis of Doravirine.
Scheme 21. Synthesis of Doravirine.
Ijms 24 07728 sch021

6.5. Fostamatinib

Fostamatinib is sold under the brand name Tavalisse and was developed by Rigel Pharmaceuticals. It was approved in April 2018 for the treatment of thrombocytopenia in adults with persistent or chronic immune thrombocytopenia (ITP) [94].
Fostamatinib is a potent spleen tyrosine kinase (Syk) inhibitor, bearing a 5-fluoropyrimidine ring, and it is used to improve membranes’ permeability [95]. Actually, Fostamatinib is a pro-drug of compound 168, and its synthesis is described in Scheme 22 [96].
5-Fluoropyrimidine-2,4-diol 163 is converted into dichloro-derivative 164 after a treatment with POCl3.
Two subsequent chloride displacements with different amines then occur. The first one at C-4 with amino-pyridoxazinone 165 yields 166; the second one at C-2 with 3,4,5-trimethoxyaniline 167 produces compound 168. As mentioned above, this compound is converted into the pro-drug Fostamatinib via a treatment with di-tert-butyl(chloromethyl)phosphate to produce ester 169, which is then hydrolyzed and converted into target phosphate disodium salt.
Scheme 22. Synthesis of Fostamatinib.
Scheme 22. Synthesis of Fostamatinib.
Ijms 24 07728 sch022

6.6. Ivosidenib

Ivosidenib is sold under the brand name Tibsovo and was developed by Servier. It was approved in July 2018 for the treatment of relapsed or refractory acute myeloid leukemia [97]. Ivosidenib is an inhibitor of mutated cytosolic isocitrate dehydrogenase 1 (IDH1); thus, it lowers the level of oncometabolite D-2-hydroxyglutarate (2-HG) [98]. Fluorine at the position five of the pyridine ring is crucial in order to ensure a high level of potency and metabolic stability [98]. The synthesis of Ivosidenib at the multi-gram scale is outlined in Scheme 23 [99]. The synthetic process is based on a multi-component Ugi reaction of 3-amino-5-fluoropyridine 170 with 2-chlorobenzaldehyde 171, followed by L-pyroglutamic acid and isocyanide 172, to produce peptide derivative 173. N-H coupling with bromopyridine 174 under Buchwald conditions produces Ivosidenib as a single enantiomer after crystallization.
Scheme 23. Synthesis of Ivosidenib.
Scheme 23. Synthesis of Ivosidenib.
Ijms 24 07728 sch023

6.7. Talazoparib

Talazoparib is sold under the brand name Talzenna and was developed by Pfizer [100]. It was approved in October 2018 for the treatment of locally advanced or metastatic breast cancer patients with a germline BRCA mutation [101]. Talazoparib, a fluorine-containing tetrahydropyridophthlazinones is active as a poly(ADP-ribose) polymerase (PARP) inhibitor [101]. To achieve inhibitory activity and metabolic stability, as well as to increase the number of interactions at the binding site via H-bonding, 5-fluoro substitution and the 4-fluorophenyl groups are crucial [101]. The synthetic method for the preparation of Talazoparib at the 30 g scale is described in Scheme 24 [102].
Fluorinated dihydroquinolinone 177 is constructed with a two steps acylation/reductive amination of fluoroaniline 175 with β-ketoacid 176. The following chiral resolution with (-)-tartaric acid is crucial to obtain the desired enantiomer, 178. The base-induced reaction between 178 and 5-chloro-1-methyl-1,2,4-triazole 179 leads to the desired trans stereoisomer, 180, which is finally converted into Talazoparib via a reaction with hydrazine in EtOH under reflux.
Scheme 24. Synthesis of Talazoparib.
Scheme 24. Synthesis of Talazoparib.
Ijms 24 07728 sch024

6.8. Tezacaftor

Tezacaftor is sold under the brand name Symdeko, as a co-formulation with ivacaftor and was developed by Vertex Pharms Inc. It was approved in February 2018 for the treatment of cystic fibrosis [103]. Tezacaftor improves the processing and trafficking of cystic fibrosis transmembrane conductance regulator (CFTR) in vitro and improves CFTR’s function alone and in combination with other drugs [104].
The synthesis of Tezacaftor has been recently reviewed [105], the second generation process is described in Scheme 25 [106].
The scheme is based on the initial formation of 6-fluoroindole’s nucleus using 3-fluoro-4-nitroaniline 181. Bromination with elemental bromine in acetic acid yields compound 182, and the nucleophilic ring opening by the anilino moiety on chiral epoxide 183 produces compound 184. Nitro-group reduction with Zinc and salt formation with PTSA from amine 185,produce tosylate 186. The latter one is coupled with terminal alkyne 187 under Sonogashira conditions to produce compound 188. Indole ring formation is achieved via a Pd-catalyzed reaction using Pd(CH3CN)2Cl2. 6-Fluoroindole 189, obtained as a single enantiomer, is then used in a reaction with chloride 190 to yield Bn-protected derivative 191, which is finally converted into Tezacaftor by means of hydrogenation over Pd/C.
Scheme 25. Synthesis of Tezacaftor.
Scheme 25. Synthesis of Tezacaftor.
Ijms 24 07728 sch025

7. FDA-Approved Drugs in 2017

In 2017, the FDA approved 46 new drugs, including 34 small molecules [107,108]. Thirty-one out of thirty-four molecules contain at least one heterocyclic ring, and ten out of thirty-four molecules contain at least one fluorine atom. In the following paragraphs, seven heterocyclic compounds bearing a fluorinated moiety directly linked to the ring are reported (Scheme 26, Scheme 27, Scheme 28, Scheme 29, Scheme 30, Scheme 31 and Scheme 32).

7.1. Abemaciclib

Abemaciclib is sold under the brand name Verzenio and was developed by Eli Lilly. It was approved in September 2017 for the treatment of advanced or metastatic breast cancers. Abemaciclib is a cyclin-dependent kinase (CDK) inhibitor that is selective for isoforms CDK4 and CDK6 [109]. Abemaciclib contains two fluorinated heterocycles, namely 4-fluorobenzimidazole and 5-fluoropyrimidine, which are directly linked to form a 6-(pyrimidin-4-yl)-benzimidazole core (Scheme 26). The synthesis described by Frederick et al. starts with the formation of this bond via a Suzuki reaction between fluorobenzimidazolyl pinacol boronate 192 and 2,4-dichloro-5-fluoropyrimidine 193 [110]. The reaction occurs selectively with the displacement of chlorine at position 4, producing 194, while less-reactive chlorine at position 2 is then used in a reaction with aminopyridine 195 under Buchwald–Hartwig conditions to yield intermediate 196. The latter substance is converted into Abemaciclib through reductive amination with ethylpyperazine 197 via a Leuckart–Wallach reaction, with trimethyl orthoformate as a dehydrating agent. A further improvement has introduced a more convergent scheme, which involves the performance of flow synthesis [111].
Scheme 26. Synthesis of Abemaciclib.
Scheme 26. Synthesis of Abemaciclib.
Ijms 24 07728 sch026

7.2. Delafloxacin

Delafloxacin is sold under the brand name Baxdela and was developed by Melinta. It was approved in June 2017 for the treatment of acute bacterial skin and skin structure infections [112]. Such as other members of the fluoroquinolone family, it is a DNA gyrase topoisomerase IV inhibitor that is active against Gram-positive bacteria, including methicillin-resistant Staphylococcus Aureus (MRSA), and Gram-negative organisms, such as Escherichia Coli and Pseudomonas Aeruginosa [113].
Additionally, some quinolone-resistant strains are susceptible to Delafloxacin. The synthetic process is in line with the classical fluoroquinolone method (Scheme 27) [114].
Scheme 27. Synthesis of Delafloxacin.
Scheme 27. Synthesis of Delafloxacin.
Ijms 24 07728 sch027
Trifluorobenzoic acid 198 is initially converted into the corresponding chloride with thionylchloride, and then into β-ketoester 200 via a treatment with potassium monoethylmalonate 199. Compound 200 is then converted into an intermediate vinylether, which is directly transformed into enamine 202 after the reaction with 2,6-diamino-3,5-difluoropyridine 201. The cyclization of compound 202 into the corresponding quinolone by the nucleophilic displacement of ortho fluorine is induced by the addition of DBU. The second aromatic nucleophilic substitution, involving fluorine at position seven, is performed with 3-hydroxyazetidine 203. Compound 204 is then obtained via the protection of an hydroxyl group as an ester to avoid competitive oxidation in the following chlorination step. Chlorination at position eight of the quinolone ring is selectively performed using NCS as a chlorinating agent in an acidic environment. Finally, Delafloxacin is obtained after the deprotection of the hydroxyazetine portion by means of ester hydrolysis with NaOH.

7.3. Enasidenib

Enasidenib is sold under the brand name Idhifa and was developed by Celgene. It was approved in August 2017 for the treatment of relapsed or refractory acute myeloid leukemia in patients with specific mutations of the isocitrate dehydrogenase 2 (IDH2) gene [115]. Enasidenib is a first-in-class small-molecule inhibitor of the IDH2-mutant enzyme with oral bioavailability [116]. This drug contains two trifluoromethylpyridine rings, as demonstrated by ab initio calculations with X-ray data; one trifluoromethyl group is important for the CF3···O tetrel bond with Asp312 [117]. The same CF3-group is also responsible for C-H···F bonding with Asp312 and N-H···F bonding with Gln316. The synthesis of Enasidenib was patented in 2013 (Scheme 28) [118].
Scheme 28. Synthesis of Enasidenib.
Scheme 28. Synthesis of Enasidenib.
Ijms 24 07728 sch028
Trifluoromethylpycolinate 205 is condensed using biuret 206 in refluxing EtOH in the presence of sodium metal to produce 1,3,5-triazin-2,4-dione 207. Chlorination with PCl5 in POCl3 produces dichlorotriazine 208. The nucleophilic displacement of aminoalcohol 209 produces compound 210. The Buchwald–Hartwig Pd-catalyzed N-arylation of 4-amino-2-(trifluoromethyl)pyridine 211 with chloride 210 forms Enasidenib.

7.4. Glecaprevir

Glecaprevir is sold under the brand name Mavyret, as a co-formulation with Pibrentasvir (see Section 7.6), and was developed by AbbVie Inc. It was approved in August 2017 for the treatment of chronic hepatitis C virus (HCV) in adults [119].
In 2019, the FDA expanded the use to children. Glecaprevir is a non-structural (NS) protein 3/4A protease inhibitor, presenting a macrocyclic ring with a difluoromethylene moiety directly linked to a quinoxaline ring [120]. The enabling synthesis of Glecaprevir to produce the quantity needed for Phase I clinical trials is based on ring-closing metathesis (RCM) for the production of the 18-membered macrocycle (Scheme 29) [121].
Scheme 29. Synthesis of Glecaprevir.
Scheme 29. Synthesis of Glecaprevir.
Ijms 24 07728 sch029
The synthetic route starts with the formation of fluorinated α-hydroxyester 214 via the Indium-mediated allylation of ethyl glyoxylate 213 starting with 3-bromo-3,3-difluoro-propene 212. The Swern oxidation of propylphosphonic anhydride (T3P) into intermediate α-ketoester is followed by condensation with ortho-phenylenediamine 215 to produce gem-difluoro quinoxaline 216. Chlorination with thionyl chloride produces derivative 217, possessing a good leaving group for nucleophilic aromatic substitution with Boc-protected hydroxyproline 218. Concurrent methyl ester formation and the removal of Boc protection via treating 219 with HCl in MeOH produces amine 220, which is one of the two main building blocks for macrocycle formation. The second main component, acid 221, is coupled, inducing amide bond formation when employing HATU as an activating agent. Diene 222 is then subjected to RCM using Zhan 1B catalyst after careful screening of the reaction conditions, thereby optimizing the formation of the desired trans macrocycle, 223. Glecaprevir is obtained in two steps via the hydrolysis of 223 into acid 224 and its coupling with the fluorinated aminocyclopropane 225 side-chain, combining EDC and 2-hydroxypyridine-N-oxide (HOPO) as activating agents. Another approach based on ether-bond macrocyclization has also been developed for large-scale synthesis [122].

7.5. Letermovir

Letermovir is sold under the brand name Prevymis and was developed by Merck & Co. It was approved in November 2017 for the treatment of infections caused by cytomegalovirus (CMV) after a bone marrow transplant [123]. Letermovir’s mode of action is different from that of other antiviral agents, which act on DNA polymerase; in fact, it interferes with the activity of terminase complex of the virus [124]. The asymmetric synthesis of Letermovir is performed in seven steps with a key part: PTC-mediated aza-Michael cyclization to obtain the chiral fluorinated dihydroquinazoline core (Scheme 30) [125]. The formation of aminocinnamate 228 is based on a Heck reaction between fluoroaniline 226 and methyl acrylate 227. Carbamate 229 is then obtained treating 228 with phenyl chloroformate. Urea 231 is formed via a reaction with anisidine 230. Compound 231 is dehydrated with PCl5 into carbodiimide 232 and directly converted into guanidine 234 via a treatment with piperazine 233. Compound 234 is the key intermediate used for asymmetric cyclization into compound 236 using fluorinated cinchona-based derivative 235 as a PTC catalyst. Precursor 236 is converted into Letermovir via the hydrolysis of the methyl ester moiety. Other asymmetric approaches to Letermovir synthesis were developed later [126,127].
Scheme 30. Synthesis of Letermovir.
Scheme 30. Synthesis of Letermovir.
Ijms 24 07728 sch030

7.6. Pibrentasvir

Pibrentasvir is sold under the brand name Mavyret, as a co-formulation with Glecaprevir (see Scheme 29), and was developed by AbbVie Inc. It was approved in August 2017 for the treatment of chronic hepatitis C virus (HCV) in adults [119]. Pibrentasvir is an NS5A inhibitor antiviral agent with two symmetric fluorobenzimidazole rings linked to a central trans pyrrolidine core [128]. The method for the patented synthesis of Pibrentasvir is shown in Scheme 31 [129].
Scheme 31. Synthesis of Pibrentasvir.
Scheme 31. Synthesis of Pibrentasvir.
Ijms 24 07728 sch031
Fluoro-acetophenone 237 was brominated using the methyl group, producing α-bromoketone 238. The ZnCl2-mediated C-C coupling of 237 with 238 produces diketone 239. Stereoselective reduction in the presence of prolinol-derived catalyst 240 yields an intermediate diol, which is directly converted into dimesylate 241. Double nucleophilic displacement with aniline 242 produces trans pyrrolidine 243, which is then treated with N-Boc prolinamide under Buchwald conditions to yield N-arylation at both rings in compound 244. The latter substance is converted into bis-benzimidazole 245 via the hydrogenation of nitro groups, AcOH-mediated cyclization and TFA-induced deprotection. Diamine 245 is finally converted into Pibrentasvir via coupling with protected O-methyl-threonine 246 with EDC and HOBt as an activating agent.

7.7. Voxilaprevir

Voxilaprevir is sold under the brand name Vosevi, as a co-formulation with sofosbuvir (see Section 8.2) and velpatasvir and was developed by Gilead. It was approved in July 2017 for the treatment of chronic hepatitis C virus (HCV) in adults [130]. Voxilaprevir is an NS protein 3/4A protease inhibitor, possessing an 18-membered gem-difluoro methylene quinoxaline portion similar to that of Glecaprevir (Scheme 29). Additionally, the synthetic pathways are quite similar (Scheme 32) [131].
Scheme 32. Synthesis of Voxilaprevir.
Scheme 32. Synthesis of Voxilaprevir.
Ijms 24 07728 sch032
The synthetic method starts with the formation of fluorinated α-ketoester 247 via the lithium exchange and allylation of diethyl oxalate, starting with 3-bromo-3,3-difluoro-propene 212. Condensation with methoxy ortho-phenylenediamine 248 produces gem-difluoro quinoxaline 249, which is chlorinated with POCl3 to produce derivative 250, which possesses a good leaving group for nucleophilic aromatic substitution with Boc-protected ethyl hydroxyproline 251. The Boc removal of 252 with HCl furnishes amine 253, which is one of the two main building blocks for macrocycle formation. The second main component, acid 254, is coupled to it by inducing amide bond formation and employing HATU as an activating agent. Diene 255 is then subjected to RCM using Zhan 1B catalyst, causing the formation of the desired trans macrocycle, 256. Glecaprevir is obtained in two steps by means of the hydrogenation of the double bond of 256 and t-Bu ester removal to produce acid 257 and its coupling with fluorinated aminocyclopropane 225 side-chain using HATU as an activating agent.

8. FDA-Approved Drugs in 2016

In 2016, the FDA approved 22 new drugs, including 13 small molecules [132]. Eleven out of thirteen molecules contain at least one heterocyclic ring, and four out of thirteen molecules have at least one fluorine atom. In the following paragraphs, two heterocyclic compounds bearing a fluorinated moiety directly linked to the ring are reported (Scheme 33 and Scheme 34).

8.1. Rucaparib

Rucaparib is sold under the brand name Rubraca and was developed by Clovis Oncology. It was approved in December 2016 for the treatment of ovarian cancer [133]. Rucaparib is the first-in-class inhibitor of DNA repair enzyme poly-ADP ribose polymerase-1 (PARP-1); notably, the presence of fluorine at the indole ring enhances the in vitro potency tenfold in comparison to that of the unfluorinated analogue [134]. Rucaparib synthesis is completed in five steps, starting with fluoroindole 258 (Scheme 33) [135]. The alkylation of 258 with nitroacetate 259 yields compound 260, which is reduced using Zinc in an acid medium and directly cyclized under basic conditions into azepino-indole 261. The bromuration of indole C-2 produces 262, which undergoes to a Suzuki reaction with formyl boronic acid 263. The obtained aldehyde 264 is converted into Rucaparib by means of reductive amination with NaBH3CN.
Scheme 33. Synthesis of Rucaparib.
Scheme 33. Synthesis of Rucaparib.
Ijms 24 07728 sch033

8.2. Sofosbuvir

Sofosbuvir is sold under the brand name Epclusa, as a co-formulation with velpatavsir, and was developed by Gilead Science. It was approved in June 2016 for the treatment of six major forms of HCV [136]. Sofsbuvir has also been FDA-approved since 2014 for the treatment of HCV, alone or co-administered with other drugs [137]. Sofosbuvir acts as an HCV NS5B polymerase inhibitor and is administered as a prodrug [138]. The synthesis of Sofosbuvir was described by Bao et al. in 2010 (Scheme 34). Sofosbuvir was originally obtained from cytidine derivative 270 via the hydrolysis of the amino group and Bz removal, followed by the installation of a phosphoramidate side-chain on 272 in the presence of N-methylimidazole (NMI) [139]. Compound 270 has been previously synthesized, starting with chiral cyclic sulfate 265 [140]. Sulfate opening using fluoride is followed by a hydrolytic step, allowing the obtainment of fluorinated chiral compound 266. Acetonide hydrolysis, performed with HCl in EtOH, straightforwardly produces lactone 267, which is protected as it is Bz ester 268, and then reduced and coupled with pyrimidine derivative 269 to produce compound 270.
Scheme 34. Synthesis of Sofosbuvir.
Scheme 34. Synthesis of Sofosbuvir.
Ijms 24 07728 sch034

9. Conclusions

The use and the importance of fluorinated heterocyclic compounds is continuously increasing in the field of medicinal chemistry. Drug discovery is often linked to the presence of a fluorinated moiety that is able to improve the drug’s potency or the metabolic stability of different heterocycles. These features allow the constant increase in these molecular entities, among other new drugs and among blockbuster and best-selling compounds. The only restriction, up to now, could be due to the limitation of fluorinated building blocks for synthetic purposes. In fact, all the reported synthetic schemes are based on available fluorinated heterocyclic compounds or their precursors, while late fluorination strategies are limited to 18F probes due to the short half-life of this isotope. The installation of fluorinated moieties at the end of the synthesis could be a different approach that could characterize future research in this field.

Author Contributions

C.R., S.A., I.P., A.P., S.B. and A.P.P. equally contributed to writing this review. All authors have read and agreed to the published version of the manuscript.

Funding

C.R. thanks PNR: Next-Generation EU, DM737/2021, CUP B79J21038330001, for funding. This work was supported by PNRR—PE0000019—“HEAL ITALIA—Health Extended Alliance for Innovative Therapies, Advanced Lab-research, and Integrated Approaches of Precision Medicine”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The University of Palermo is gratefully acknowledged.

Conflicts of Interest

There are no conflict to declare.

References

  1. Kerru, N.; Gummidi, L.; Maddila, S.; Gangu, K.K.; Jonnalagadda, S.B. A Review on Recent Advances in Nitrogen-Containing Molecules and Their Biological Applications. Molecules 2020, 25, 1909. [Google Scholar] [CrossRef]
  2. Martorana, A.; Giacalone, V.; Bonsignore, R.; Pace, A.; Gentile, C.; Pibiri, I.; Buscemi, S.; Lauria, A.; Piccionello, P.A. Heterocyclic Scaffolds for the Treatment of Alzheimer’s Disease. Curr. Pharm. Des. 2016, 22, 3971–3995. [Google Scholar] [CrossRef]
  3. Pathania, S.; Narang, R.K.; Rawal, R.K. Role of sulphur-heterocycles in medicinal chemistry: An update. Eur. J. Med. Chem. 2019, 180, 486–508. [Google Scholar] [CrossRef]
  4. Wetzel, C.; Lonneman, M.; Wu, C. Polypharmacological drug actions of recently FDA approved antibiotics. Eur. J. Med. Chem. 2021, 209, 112931. [Google Scholar] [CrossRef] [PubMed]
  5. Heravi, M.M.; Zadsirjan, V. Prescribed drugs containing nitrogen heterocycles: An overview. RSC Adv. 2020, 10, 44247–44311. [Google Scholar] [CrossRef]
  6. Inoue, M.; Sumii, Y.; Shibata, N. Contribution of Organofluorine Compounds to Pharmaceuticals. ACS Omega 2020, 5, 10633–10640. [Google Scholar] [CrossRef] [PubMed]
  7. Upadhyay, C.; Chaudhary, M.; De Oliveira, R.N.; Borbas, A.; Kempaiah, P.S.; Rathi, B. Fluorinated scaffolds for antimalarial drug discovery. Expert Opin. Drug Discov. 2020, 15, 705–718. [Google Scholar] [CrossRef]
  8. O’Hagan, D. Fluorine in health care: Organofluorine containing blockbuster drugs. J. Fluor. Chem. 2010, 131, 1071–1081. [Google Scholar] [CrossRef]
  9. Fried, J.; Sabo, E.F. 9α-Fluoro Derivatives of Cortisone and Hydrocortisone. J. Am. Chem. Soc. 1954, 76, 1455–1456. [Google Scholar] [CrossRef]
  10. Purser, S.; Moore, P.R.; Swallow, S.; Gouverneur, V. Fluorine in medicinal chemistry. Chem. Soc. Rev. 2008, 37, 320–330. [Google Scholar] [CrossRef] [PubMed]
  11. 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] [PubMed]
  12. 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]
  13. Morgenthaler, M.; Schweizer, E.; Hoffmann-Röder, A.; Benini, F.; Martin, R.E.; Jaeschke, G.; Wagner, B.; Fischer, H.; Bendels, S.; Zimmerli, D.; et al. Predicting and Tuning Physicochemical Properties in Lead Optimization: Amine Basicities. ChemMedChem 2007, 2, 1100–1115. [Google Scholar] [CrossRef] [PubMed]
  14. Rowley, M.; Hallett, D.J.; Goodacre, S.; Moyes, C.; Crawforth, J.; Sparey, T.J.; Patel, S.; Marwood, R.; Patel, S.; Thomas, S.; et al. 3-(4-Fluoropiperidin-3-yl)-2-phenylindoles as High Affinity, Selective, and Orally Bioavailable h5-HT2A Receptor Antagonists. J. Med. Chem. 2001, 44, 1603–1614. [Google Scholar] [CrossRef] [PubMed]
  15. Shah, P.; Westwell, A.D. The role of fluorine in medicinal chemistry. J. Enzym. Inhib. Med. Chem. 2007, 22, 527–540. [Google Scholar] [CrossRef]
  16. Mei, H.; Han, J.; Fustero, S.; Medio-Simon, M.; Sedgwick, D.M.; Santi, C.; Ruzziconi, R.; Soloshonok, V.A. Fluorine-Containing Drugs Approved by the FDA in 2018. Chem. A Eur. J. 2019, 25, 11797–11819. [Google Scholar] [CrossRef] [PubMed]
  17. Knight, J.C.; Edwards, P.G.; Paisey, S.J. Fluorinated contrast agents for magnetic resonance imaging; a review of recent developments. RSC Adv. 2011, 1, 1415–1425. [Google Scholar] [CrossRef]
  18. Van der Born, D.; Pees, A.; Poot, A.J.; Orru, R.V.A.; Windhorst, A.D.; Vugts, D.J. Fluorine-18 labelled building blocks for PET tracer synthesis. Chem. Soc. Rev. 2017, 46, 4709–4773. [Google Scholar] [CrossRef]
  19. Das, P.; Delost, M.D.; Qureshi, M.H.; Smith, D.T.; Njardarson, J.T. A Survey of the Structures of US FDA Approved Combination Drugs. J. Med. Chem. 2019, 62, 4265–4311. [Google Scholar] [CrossRef]
  20. Ivasyshyn, V.; Smit, H.; Chiechi, R.C. Synthesis of a Hominal Bis(difluoromethyl) Fragment. ACS Omega 2019, 4, 14140–14150. [Google Scholar] [CrossRef]
  21. Liang, T.; Neumann, C.N.; Ritter, T. Introduction of Fluorine and Fluorine-Containing Functional Groups. Angew. Chem. Int. Ed. 2013, 52, 8214–8264. [Google Scholar] [CrossRef]
  22. López, S.E.; Salazar, J. Trifluoroacetic acid: Uses and recent applications in organic synthesis. J. Fluor. Chem. 2013, 156, 73–100. [Google Scholar] [CrossRef]
  23. Shibata, N.; Matsnev, A.; Cahard, D. Shelf-stable electrophilic trifluoromethylating reagents: A brief historical perspective. Beilstein J. Org. Chem. 2010, 6, 65. [Google Scholar] [CrossRef]
  24. Eisenberger, P.; Gischig, S.; Togni, A. Novel 10-I-3 Hypervalent Iodine-Based Compounds for Electrophilic Trifluoromethylation. Chem. A Eur. J. 2006, 12, 2579–2586. [Google Scholar] [CrossRef]
  25. Magnier, E.; Blazejewski, J.-C.; Tordeux, M.; Wakselman, C. Straightforward One-Pot Synthesis of Trifluoromethyl Sulfonium Salts. Angew. Chem. Int. Ed. 2006, 45, 1279–1282. [Google Scholar] [CrossRef] [PubMed]
  26. FDA. Novel Drug Approvals for 2022; FDA: Silver Spring, MD, USA, 2022.
  27. De la Torre, B.G.; Albericio, F. The Pharmaceutical Industry in 2022: An Analysis of FDA Drug Approvals from the Perspective of Molecules. Molecules 2023, 28, 1038. [Google Scholar] [CrossRef] [PubMed]
  28. Paik, J. Lenacapavir: First Approval. Drugs 2022, 82, 1499–1504. [Google Scholar] [CrossRef]
  29. Segal-Maurer, S.; DeJesus, E.; Stellbrink, H.-J.; Castagna, A.; Richmond, G.J.; Sinclair, G.I.; Siripassorn, K.; Ruane, P.J.; Berhe, M.; Wang, H.; et al. Capsid Inhibition with Lenacapavir in Multidrug-Resistant HIV-1 Infection. N. Engl. J. Med. 2022, 386, 1793–1803. [Google Scholar] [CrossRef]
  30. Graupe, M.; Henry, S.J.; Link, J.O.; Rowe, C.W.; Saito, R.D.; Schroeder, S.; Stefanidis, D.; Tse, W.C.; Zhang, J.R. Therapeutic Compounds Useful for the Prophylactic or Treatment of an HIV Virus Infection. Patent WO 2018035359, 22 February 2018. [Google Scholar]
  31. Hoy, S.M. Oteseconazole: First Approval. Drugs 2022, 82, 1017–1023. [Google Scholar] [CrossRef]
  32. Sobel, J.D.; Nyirjesy, P. Oteseconazole: An advance in treatment of recurrent vulvovaginal candidiasis. Future Microbiol. 2021, 16, 1453–1461. [Google Scholar] [CrossRef] [PubMed]
  33. Hoekstra, W.J.; Schotzinger, R.J.; Rafferty, S.W. Metalloenzyme Inhibitor Compounds. Patent WO 2011/133875, 27 October 2011. [Google Scholar]
  34. Benedetto Tiz, D.; Bagnoli, L.; Rosati, O.; Marini, F.; Sancineto, L.; Santi, C. New Halogen-Containing Drugs Approved by FDA in 2021: An Overview on Their Syntheses and Pharmaceutical Use. Molecules 2022, 27, 1643. [Google Scholar] [CrossRef] [PubMed]
  35. Deeks, E.D. Atogepant: First Approval. Drugs 2022, 82, 65–70. [Google Scholar] [CrossRef]
  36. Hay, D.L.; Walker, C.S.; Harris, P.W.R. Atogepant (Qulipta®) for migraine prevention. Trends Pharmacol. Sci. 2022, 43, 701–702. [Google Scholar] [CrossRef] [PubMed]
  37. Belyk, K.M.; Cleator, E.; Kuo, S.; Maligres, P.E.; Xiang, B.; Yasuda, N.; Yin, J. Process for Making CGRP Receptor Antagonists. Patent WO 2013/138418, 19 September 2013. [Google Scholar]
  38. Wood, M.R.; Bell, I.M.; Gallicchio, S.N.; Selnick, H.G.; Stump, C.A.; Zartman, C.B. Substituted Spirocyclic CGRP Receptor Antagonists. Patent WO 2007133491, 22 December 2007. [Google Scholar]
  39. Keam, S.J. Piflufolastat F 18: Diagnostic First Approval. Mol. Diagn. Ther. 2021, 25, 647–656. [Google Scholar] [CrossRef]
  40. Chen, Y.; Pullambhatla, M.; Foss, C.A.; Byun, Y.; Nimmagadda, S.; Senthamizhchelvan, S.; Sgouros, G.; Mease, R.C.; Pomper, M.G. 2-(3-{1-Carboxy-5-[(6-[18F]fluoro-pyridine-3-carbonyl)-amino]-pentyl}-ureido)-pentanedioic acid, [18F]DCFPyL, a PSMA-based PET imaging agent for prostate cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2011, 17, 7645–7653. [Google Scholar] [CrossRef] [PubMed]
  41. Olberg, D.E.; Arukwe, J.M.; Grace, D.; Hjelstuen, O.K.; Solbakken, M.; Kindberg, G.M.; Cuthbertson, A. One Step Radiosynthesis of 6-[18F]Fluoronicotinic Acid 2,3,5,6-Tetrafluorophenyl Ester ([18F]F-Py-TFP): A New Prosthetic Group for Efficient Labeling of Biomolecules with Fluorine-18. J. Med. Chem. 2010, 53, 1732–1740. [Google Scholar] [CrossRef] [PubMed]
  42. Canon, J.; Rex, K.; Saiki, A.Y.; Mohr, C.; Cooke, K.; Bagal, D.; Gaida, K.; Holt, T.; Knutson, C.G.; Koppada, N.; et al. The clinical KRAS(G12C) inhibitor AMG 510 drives anti-tumour immunity. Nature 2019, 575, 217–223. [Google Scholar] [CrossRef]
  43. Lanman, B.A.; Allen, J.R.; Allen, J.G.; Amegadzie, A.K.; Ashton, K.S.; Booker, S.K.; Chen, J.J.; Chen, N.; Frohn, M.J.; Goodman, G.; et al. Discovery of a Covalent Inhibitor of KRASG12C (AMG 510) for the Treatment of Solid Tumors. J. Med. Chem. 2020, 63, 52–65. [Google Scholar] [CrossRef]
  44. Lanman, B.A.; Chen, J.; Reed Anthony, B.; Cee Victor, J.; Liu, L.; Kopecky, D.J.; Lopez, P.; Wurz, R.P.; Nguyen, T.T.; Booker, S.; et al. Kras G12C Inhibitors and Method of Using the Same. Patent US 2018/0334454, 22 December 2018. [Google Scholar]
  45. Zhang, L.; Griffin, D.J.; Beaver, M.G.; Blue, L.E.; Borths, C.J.; Brown, D.B.; Caille, S.; Chen, Y.; Cherney, A.H.; Cochran, B.M.; et al. Development of a Commercial Manufacturing Process for Sotorasib, a First-in-Class KRASG12C Inhibitor. Org. Process Res. Dev. 2022, 26, 3115–3125. [Google Scholar] [CrossRef]
  46. Lunning, M.; Vose, J.; Nastoupil, L.; Fowler, N.; Burger, J.A.; Wierda, W.G.; Schreeder, M.T.; Siddiqi, T.; Flowers, C.R.; Cohen, J.B.; et al. Ublituximab and umbralisib in relapsed/refractory B-cell non-Hodgkin lymphoma and chronic lymphocytic leukemia. Blood 2019, 134, 1811–1820. [Google Scholar] [CrossRef]
  47. Weiss, M.; Miskin, H.; Sportelli, P.S.; Vakkalanka, K.V.S. Combination of Anti-CD20 Antibody and PI3 Kinase Selective Inhibitor. Patent WO 2014071125, 8 May 2014. [Google Scholar]
  48. Follmann, M.; Ackerstaff, J.; Redlich, G.; Wunder, F.; Lang, D.; Kern, A.; Fey, P.; Griebenow, N.; Kroh, W.; Becker-Pelster, E.-M.; et al. Discovery of the Soluble Guanylate Cyclase Stimulator Vericiguat (BAY 1021189) for the Treatment of Chronic Heart Failure. J. Med. Chem. 2017, 60, 5146–5161. [Google Scholar] [CrossRef] [PubMed]
  49. Yuan, S.; Luo, Y.Q.; Zuo, J.H.; Liu, H.; Li, F.; Yu, B. New drug approvals for 2020: Synthesis and clinical applications. Eur. J. Med. Chem. 2021, 215, 113284. [Google Scholar] [CrossRef]
  50. Ohsawa, I.; Honda, D.; Suzuki, Y.; Fukuda, T.; Kohga, K.; Morita, E.; Moriwaki, S.; Ishikawa, O.; Sasaki, Y.; Tago, M.; et al. Oral berotralstat for the prophylaxis of hereditary angioedema attacks in patients in Japan: A phase 3 randomized trial. Allergy 2021, 76, 1789–1799. [Google Scholar] [CrossRef] [PubMed]
  51. Yahya El-Kattan, Y.S.B. Crystalline Salts of a Plasma Kallikrein Inhibitor. Patent US 202010662160, 7 May 2020. [Google Scholar]
  52. Dhillon, S. Decitabine/Cedazuridine: First Approval. Drugs 2020, 80, 1373–1378. [Google Scholar] [CrossRef] [PubMed]
  53. Ferraris, D.; Duvall, B.; Delahanty, G.; Mistry, B.; Alt, J.; Rojas, C.; Rowbottom, C.; Sanders, K.; Schuck, E.; Huang, K.-C.; et al. Design, Synthesis, and Pharmacological Evaluation of Fluorinated Tetrahydrouridine Derivatives as Inhibitors of Cytidine Deaminase. J. Med. Chem. 2014, 57, 2582–2588. [Google Scholar] [CrossRef]
  54. Brown, K.; Dixey, M.; Weymouth-Wilson, A.; Linclau, B. The synthesis of gemcitabine. Carbohydr. Res. 2014, 387, 59–73. [Google Scholar] [CrossRef] [PubMed]
  55. Markham, A. Pralsetinib: First Approval. Drugs 2020, 80, 1865–1870. [Google Scholar] [CrossRef] [PubMed]
  56. Subbiah, V.; Shen, T.; Terzyan, S.S.; Liu, X.; Hu, X.; Patel, K.P.; Hu, M.; Cabanillas, M.; Behrang, A.; Meric-Bernstam, F.; et al. Structural basis of acquired resistance to selpercatinib and pralsetinib mediated by non-gatekeeper RET mutations. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2021, 32, 261–268. [Google Scholar] [CrossRef]
  57. Brubaker, J.D.; Kim, J.L.; Wilson, K.J.; Wilson, D.; Di Pietro, L.V. Inhibitors of Ret. Patent US 20170121312, 4 May 2017. [Google Scholar]
  58. Markham, A.; Keam, S.J. Selumetinib: First Approval. Drugs 2020, 80, 931–937. [Google Scholar] [CrossRef]
  59. Wallace, E.M.; Lyssikatos, J.P.; Hurley, B.T.; Marlow, A.L. N3 Alkylated Benzimiadazole Derivates Ad MEK Inhibitors. Patent WO 2003077914, 25 September 2003. [Google Scholar]
  60. Mu, L.; Jie, C.V.M.L.; Treyer, V.; Schibli, R. Tauvid™: The First FDA-Approved PET Tracer for Imaging Tau Pathology in Alzheimer’s Disease. Pharmaceuticals 2021, 14, 110. [Google Scholar] [CrossRef]
  61. Xia, C.F.; Arteaga, J.; Chen, G.; Gangadharmath, U.; Gomez, L.F.; Kasi, D.; Lam, C.; Liang, Q.; Liu, C.; Mocharla, V.P.; et al. [(18)F]T807, a novel tau positron emission tomography imaging agent for Alzheimer’s disease. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2013, 9, 666–676. [Google Scholar] [CrossRef] [PubMed]
  62. Attardo, G.; Lister-James, J.; Xiong, H.; Lim, N. Compounds and Their Use for Preparation of Tau Imaging Agents and Tau Imaging Formulations. Patent WO 2015047902, 2 April 2015. [Google Scholar]
  63. Patridge, E.; Gareiss, P.; Kinch, M.S.; Hoyer, D. An analysis of FDA-approved drugs: Natural products and their derivatives. Drug Discov. Today 2016, 21, 204–207. [Google Scholar] [CrossRef] [PubMed]
  64. Yuan, S.; Yu, B.; Liu, H.M. New drug approvals for 2019: Synthesis and clinical applications. Eur. J. Med. Chem. 2020, 205, 112667. [Google Scholar] [CrossRef] [PubMed]
  65. Mei, H.; Remete, A.M.; Zou, Y.; Moriwaki, H.; Fustero, S.; Kiss, L.; Soloshonok, V.A.; Han, J. Fluorine-containing drugs approved by the FDA in 2019. Chin. Chem. Lett. 2020, 31, 2401–2413. [Google Scholar] [CrossRef]
  66. Markham, A. Alpelisib: First Global Approval. Drugs 2019, 79, 1249–1253. [Google Scholar] [CrossRef] [PubMed]
  67. André, F.; Ciruelos, E.; Rubovszky, G.; Campone, M.; Loibl, S.; Rugo, H.S.; Iwata, H.; Conte, P.; Mayer, I.A.; Kaufman, B.; et al. Alpelisib for PIK3CA-Mutated, Hormone Receptor-Positive Advanced Breast Cancer. N. Engl. J. Med. 2019, 380, 1929–1940. [Google Scholar] [CrossRef] [PubMed]
  68. Furet, P.; Guagnano, V.; Fairhurst, R.A.; Imbach-Weese, P.; Bruce, I.; Knapp, M.; Fritsch, C.; Blasco, F.; Blanz, J.; Aichholz, R.; et al. Discovery of NVP-BYL719 a potent and selective phosphatidylinositol-3 kinase alpha inhibitor selected for clinical evaluation. Bioorg. Med. Chem. Lett. 2013, 23, 3741–3748. [Google Scholar] [CrossRef] [PubMed]
  69. Caravatti, G.; Fairhutst, R.A.; Guagnano, V.; Imbach, P.; Furet, P. Thiazole Derivate Used as PI 3 KINASE Inhibitors. Patent WO 2009080694, 2 July 2009. [Google Scholar]
  70. Caravatti, G.; Fairhutst, R.A.; Guagnano, V.; Imbach, P.; Furet, P. Organic Compounds. Patent WO 2010029082, 10 September 2010. [Google Scholar]
  71. Beuckmann, C.T.; Suzuki, M.; Ueno, T.; Nagaoka, K.; Arai, T.; Higashiyama, H. In Vitro and In Silico Characterization of Lemborexant (E2006), a Novel Dual Orexin Receptor Antagonist. J. Pharmacol. Exp. Ther. 2017, 362, 287–295. [Google Scholar] [CrossRef] [PubMed]
  72. Beuckmann, C.T.; Ueno, T.; Nakagawa, M.; Suzuki, M.; Akasofu, S. Preclinical in vivo characterization of lemborexant (E2006), a novel dual orexin receptor antagonist for sleep/wake regulation. Sleep 2019, 42, zsz076. [Google Scholar] [CrossRef]
  73. Yoshida, Y.; Naoe, Y.; Terauchi, T.; Ozaki, F.; Doko, T.; Takemura, A.; Tanaka, T.; Sorimachi, K.; Beuckmann, C.T.; Suzuki, M.; et al. Discovery of (1R,2S)-2-{[(2,4-Dimethylpyrimidin-5-yl)oxy]methyl}-2-(3-fluorophenyl)-N-(5-fluoropyridin-2-yl)cyclopropanecarboxamide (E2006): A Potent and Efficacious Oral Orexin Receptor Antagonist. J. Med. Chem. 2015, 58, 4648–4664. [Google Scholar] [CrossRef]
  74. Lamb, Y.N. Pexidartinib: First Approval. Drugs 2019, 79, 1805–1812. [Google Scholar] [CrossRef]
  75. Ibrahim, P.N.; Jin, M.; Matsuura, S. Synthesis of 1H-Pyrrolo[2,3-B]pyridin Derivates That Modulate Kinases. Patent WO 2016179412, 10 November 2016. [Google Scholar]
  76. Dodick, D.W.; Lipton, R.B.; Ailani, J.; Lu, K.; Finnegan, M.; Trugman, J.M.; Szegedi, A. Ubrogepant for the Treatment of Migraine. N. Engl. J. Med. 2019, 381, 2230–2241. [Google Scholar] [CrossRef] [PubMed]
  77. Bell, I.M.; Fraley, M.E.; Gallicchio, S.N.; Ginnetti, A.; Mitchell, H.; Paone, D.V.; Wang, S.D.D.C.; Zartman, C.B.; Stevenson, H.E. Piperidone Carbozamide CGRP Receptor Antagonists. Patent WO 2012064910, 18 May 2012. [Google Scholar]
  78. Flick, A.C.; Leverett, C.A.; Ding, H.X.; McInturff, E. Synthetic Approaches to New Drugs Approved during 2018. J. Med. Chem. 2020, 63, 10652–10704. [Google Scholar] [CrossRef]
  79. Al-Salama, Z.T. Apalutamide: First Global Approval. Drugs 2018, 78, 699–705. [Google Scholar] [CrossRef] [PubMed]
  80. Smith, M.R.; Saad, F.; Chowdhury, S.; Oudard, S.; Hadaschik, B.A.; Graff, J.N.; Olmos, D.; Mainwaring, P.N.; Lee, J.Y.; Uemura, H.; et al. Apalutamide Treatment and Metastasis-free Survival in Prostate Cancer. N. Engl. J. Med. 2018, 378, 1408–1418. [Google Scholar] [CrossRef] [PubMed]
  81. Clegg, N.J.; Wongvipat, J.; Joseph, J.D.; Tran, C.; Ouk, S.; Dilhas, A.; Chen, Y.; Grillot, K.; Bischoff, E.D.; Cai, L.; et al. ARN-509: A novel antiandrogen for prostate cancer treatment. Cancer Res. 2012, 72, 1494–1503. [Google Scholar] [CrossRef]
  82. Jung, M.; Sawyers, E.C.; Ouk, S.L.; Tran, C.; Wongvipat, J. Androgen Receptor Modulator for the Treatment of Prostate Cancer and Androgen Receptor-Associated Disease. Patent WO 2007126765, 8 November 2007. [Google Scholar]
  83. Hughes, D.L. Review of Synthetic Routes and Crystalline Forms of the Antiandrogen Oncology Drugs Enzalutamide, Apalutamide, and Darolutamide. Org. Process Res. Dev. 2020, 24, 347–362. [Google Scholar] [CrossRef]
  84. Heo, Y.A. Baloxavir: First Global Approval. Drugs 2018, 78, 693–697. [Google Scholar] [CrossRef]
  85. Noshi, T.; Kitano, M.; Taniguchi, K.; Yamamoto, A.; Omoto, S.; Baba, K.; Hashimoto, T.; Ishida, K.; Kushima, Y.; Hattori, K.; et al. In vitro characterization of baloxavir acid, a first-in-class cap-dependent endonuclease inhibitor of the influenza virus polymerase PA subunit. Antivir. Res. 2018, 160, 109–117. [Google Scholar] [CrossRef]
  86. Takashita, E.; Morita, H.; Ogawa, R.; Nakamura, K.; Fujisaki, S.; Shirakura, M.; Kuwahara, T.; Kishida, N.; Watanabe, S.; Odagiri, T. Susceptibility of Influenza Viruses to the Novel Cap-Dependent Endonuclease Inhibitor Baloxavir Marboxil. Front. Microbiol. 2018, 9, 3026. [Google Scholar] [CrossRef]
  87. Shibahara, S.F.; Nobuaki; Toshikatsu, M. Method for Producing Substituted Polycyclic Derivate and Crystal of Same. Patent JP 20176212678, 28 December 2017. [Google Scholar]
  88. Koelblinger, P.; Dornbierer, J.; Dummer, R. A review of binimetinib for the treatment of mutant cutaneous melanoma. Future Oncol. 2017, 13, 1755–1766. [Google Scholar] [CrossRef]
  89. Ascierto, P.A.; Schadendorf, D.; Berking, C.; Agarwala, S.S.; van Herpen, C.M.; Queirolo, P.; Blank, C.U.; Hauschild, A.; Beck, J.T.; St-Pierre, A.; et al. MEK162 for patients with advanced melanoma harbouring NRAS or Val600 BRAF mutations: A non-randomised, open-label phase 2 study. Lancet Oncol. 2013, 14, 249–256. [Google Scholar] [CrossRef] [PubMed]
  90. Chen, J. Synthetic Method of Binimetinib. Patent CN 2016105820124, 3 August 2016. [Google Scholar]
  91. Colombier, M.A.; Molina, J.M. Doravirine: A review. Curr. Opin. HIV AIDS 2018, 13, 308–314. [Google Scholar] [CrossRef] [PubMed]
  92. Côté, B.; Burch, J.D.; Asante-Appiah, E.; Bayly, C.; Bédard; Blouin, L.M.; Campeau, L.C.; Cauchon, E.; Chan, M.; Chefson, A.; et al. Discovery of MK-1439, an orally bioavailable non-nucleoside reverse transcriptase inhibitor potent against a wide range of resistant mutant HIV viruses. Bioorganic Med. Chem. Lett. 2014, 24, 917–922. [Google Scholar] [CrossRef]
  93. Burch, J.; Nguyen, N.; Li, C.S.; St-onge, M.; Gauvreau, D.; Cote, B. Non-Nucleoside Reverse Transcriptase Inhibitors. Patent WO 2011120133, 6 October 2011. [Google Scholar]
  94. Markham, A. Fostamatinib: First Global Approval. Drugs 2018, 78, 959–963. [Google Scholar] [CrossRef] [PubMed]
  95. Pine, P.R.; Chang, B.; Schoettler, N.; Banquerigo, M.L.; Wang, S.; Lau, A.; Zhao, F.; Grossbard, E.B.; Payan, D.G.; Brahn, E. Inflammation and bone erosion are suppressed in models of rheumatoid arthritis following treatment with a novel Syk inhibitor. Clin. Immunol. 2007, 124, 244–257. [Google Scholar] [CrossRef]
  96. Felfer, U.; Giselbrecht, K.-H.; Wolberg, M. Syntesis of N4-(2,2-Dimethyl-4-[(Dihydrogen Phosphonoxy]-3-Oxo-5-Pyrydo [1,4] Oxazin6-yl)-5-Fluoro-N2-(3,4,5,-Trimethoxyphenyl)-2,4-Pyrimidinediamine Disodium Salt. Patent WO 2011002999, 6 January 2011. [Google Scholar]
  97. Dhillon, S. Ivosidenib: First Global Approval. Drugs 2018, 78, 1509–1516. [Google Scholar] [CrossRef]
  98. Popovici-Muller, J.; Lemieux, R.M.; Artin, E.; Saunders, J.O.; Salituro, F.G.; Travins, J.; Cianchetta, G.; Cai, Z.; Zhou, D.; Cui, D.; et al. Discovery of AG-120 (Ivosidenib): A First-in-Class Mutant IDH1 Inhibitor for the Treatment of IDH1 Mutant Cancers. ACS Med. Chem. Lett. 2018, 9, 300–305. [Google Scholar] [CrossRef]
  99. Lemieux, R.M.; Popovici-Muller, J.; Travins, J.; Cai, Z.; Cui, D.; Zhou, D. Therapeutically Active Compounds and Their Methods of Use. Patent WO 2013107291, 25 July 2013. [Google Scholar]
  100. Hoy, S.M. Talazoparib: First Global Approval. Drugs 2018, 78, 1939–1946. [Google Scholar] [CrossRef]
  101. Wang, B.; Chu, D.; Feng, Y.; Shen, Y.; Aoyagi-Scharber, M.; Post, L.E. Discovery and Characterization of (8S,9R)-5-Fluoro-8-(4-fluorophenyl)-9-(1-methyl-1H-1,2,4-triazol-5-yl)-2,7,8,9-tetrahydro-3H-pyrido[4,3,2-de]phthalazin-3-one(BMN673,Talazoparib), a Novel, Highly Potent, and Orally Efficacious Poly(ADP-ribose) Polymerase-1/2 Inhibitor, as an Anticancer Agent. J. Med. Chem. 2016, 59, 335–357. [Google Scholar] [CrossRef] [PubMed]
  102. Xu, Y.Y.; Peter, W.; Michael, X.; Douglas, C. Synthesis of Parp Inhibitor Talazoparib. Patent US 20179708319, 21 December 2017. [Google Scholar]
  103. Mospan, C.; Mospan, G.; Byland, E.; Whitaker, W.B.; Xiong, L.; Dunlap, J.; Canupp, K. Drug updates and approvals: 2018 in review. Nurse Pract. 2018, 43, 23–32. [Google Scholar] [CrossRef]
  104. Donaldson, S.H.; Pilewski, J.M.; Griese, M.; Cooke, J.; Viswanathan, L.; Tullis, E.; Davies, J.C.; Lekstrom-Himes, J.A.; Wang, L.T. Tezacaftor/Ivacaftor in Subjects with Cystic Fibrosis and F508del/F508del-CFTR or F508del/G551D-CFTR. Am. J. Respir. Crit. Care Med. 2018, 197, 214–224. [Google Scholar] [CrossRef]
  105. Hughes, D.L. Patent Review of Synthetic Routes and Crystalline Forms of the CFTR-Modulator Drugs Ivacaftor, Lumacaftor, Tezacaftor, and Elexacaftor. Org. Process Res. Dev. 2019, 23, 2302–2322. [Google Scholar] [CrossRef]
  106. Tanoury, G.J.; Harrison, C.; Littler, B.J.; Rose, P.J.; Hughes, R.M.; Jung, Y.C.; Siesel, D.A.; Lee, E.C.; Belmont, D.T. Process of Producing Cycloalkylcarboxamido-Indole Compounds. Patent WO 2011133751, 27 October 2011. [Google Scholar]
  107. De la Torre, B.G.; Albericio, F. The Pharmaceutical Industry in 2017. An Analysis of FDA Drug Approvals from the Perspective of Molecules. Molecules 2018, 23, 533. [Google Scholar] [CrossRef]
  108. Mullard, A. 2017 FDA drug approvals, Nature reviews. Drug Discov. 2018, 17, 81–85. [Google Scholar] [CrossRef] [PubMed]
  109. Martin, M.; Garcia-Saenz, J.A. Abemaciclib, a CDK4 and CDK6 inhibitor for the treatment of metastatic breast cancer. Futur. Oncol. 2020, 16, 2763–2778. [Google Scholar] [CrossRef] [PubMed]
  110. Frederick, M.O.; Kjell, D.P. A synthesis of abemaciclib utilizing a Leuckart–Wallach reaction. Tetrahedron Lett. 2015, 56, 949–951. [Google Scholar] [CrossRef]
  111. Frederick, M.O.; Pietz, M.A.; Kjell, D.P.; Richey, R.N.; Tharp, G.A.; Touge, T.; Yokoyama, N.; Kida, M.; Matsuo, T. Development of a Leuckart–Wallach Reaction in Flow for the Synthesis of Abemaciclib. Org. Process Res. Dev. 2017, 21, 1447–1451. [Google Scholar] [CrossRef]
  112. Markham, A. Delafloxacin: First Global Approval. Drugs 2017, 77, 1481–1486. [Google Scholar] [CrossRef] [PubMed]
  113. Harnett, S.J.; Fraise, A.P.; Andrews, J.M.; Jevons, G.; Brenwald, N.P.; Wise, R. Comparative study of the in vitro activity of a new fluoroquinolone, ABT-492. J. Antimicrob. Chemother. 2004, 53, 783–792. [Google Scholar] [CrossRef]
  114. Barnes, D.M.; Christesen, A.C.; Engstrom, K.M.; Haight, A.R.; Hsu, M.C.; Lee, E.C.; Peterson, M.J.; Plata, D.J.; Raje, P.S.; Stoner, E.J.; et al. Chlorination at the 8-Position of a Functionalized Quinolone and the Synthesis of Quinolone Antibiotic ABT-492. Org. Process Res. Dev. 2006, 10, 803–807. [Google Scholar] [CrossRef]
  115. Kim, E.S. Enasidenib: First Global Approval. Drugs 2017, 77, 1705–1711. [Google Scholar] [CrossRef] [PubMed]
  116. Stein, E.M.; DiNardo, C.D.; Pollyea, D.A.; Fathi, A.T.; Roboz, G.J.; Altman, J.K.; Stone, R.M.; DeAngelo, D.J.; Levine, R.L.; Flinn, I.W.; et al. Enasidenib in mutant IDH2 relapsed or refractory acute myeloid leukemia. Blood 2017, 130, 722–731. [Google Scholar] [CrossRef]
  117. García-Llinás, X.; Bauzá, A.; Seth, S.K.; Frontera, A. Importance of R–CF3···O Tetrel Bonding Interactions in Biological Systems. J. Phys. Chem. A 2017, 121, 5371–5376. [Google Scholar] [CrossRef] [PubMed]
  118. Cianchetta, B.D.G.; Popovici-Muller, J.F.; Salituro, G.; Saunders, J.O.; Travins, J.; Yan, S.; Guo, T.; Zhang, L. Therapeutically Active Compounds and Their Methods of Use. Patent WO 2013102431, 7 November 2013. [Google Scholar]
  119. Lamb, Y.N. Glecaprevir/Pibrentasvir: First Global Approval. Drugs 2017, 77, 1797–1804. [Google Scholar] [CrossRef]
  120. Lawitz Eric, J.; O’Riordan William, D.; Asatryan, A.; Freilich Bradley, L.; Box Terry, D.; Overcash, J.S.; Lovell, S.; Ng Teresa, I.; Liu, W.; Campbell, A.; et al. Potent Antiviral Activities of the Direct-Acting Antivirals ABT-493 and ABT-530 with Three-Day Monotherapy for Hepatitis C Virus Genotype 1 Infection. Antimicrob. Agents Chemother. 2016, 60, 1546–1555. [Google Scholar] [CrossRef]
  121. Cink, R.D.; Lukin, K.A.; Bishop, R.D.; Zhao, G.; Pelc, M.J.; Towne, T.B.; Gates, B.D.; Ravn, M.M.; Hill, D.R.; Ding, C.; et al. Development of the Enabling Route for Glecaprevir via Ring-Closing Metathesis. Org. Process Res. Dev. 2020, 24, 183–200. [Google Scholar] [CrossRef]
  122. Kallemeyn, J.M.; Engstrom, K.M.; Pelc, M.J.; Lukin, K.A.; Morrill, W.H.; Wei, H.; Towne, T.B.; Henle, J.; Nere, N.K.; Welch, D.S.; et al. Development of a Large-Scale Route to Glecaprevir: Synthesis of the Macrocycle via Intramolecular Etherification. Org. Process Res. Dev. 2020, 24, 1373–1392. [Google Scholar] [CrossRef]
  123. Kim, E.S. Letermovir: First Global Approval. Drugs 2018, 78, 147–152. [Google Scholar] [CrossRef] [PubMed]
  124. Goldner, T.; Hewlett, G.; Ettischer, N.; Ruebsamen-Schaeff, H.; Zimmermann, H.; Lischka, P. The novel anticytomegalovirus compound AIC246 (Letermovir) inhibits human cytomegalovirus replication through a specific antiviral mechanism that involves the viral terminase. J. Virol. 2011, 85, 10884–10893. [Google Scholar] [CrossRef]
  125. Humphrey, G.R.; Dalby, S.M.; Andreani, T.; Xiang, B.; Luzung, M.R.; Song, Z.J.; Shevlin, M.; Christensen, M.; Belyk, K.M.; Tschaen, D.M. Asymmetric Synthesis of Letermovir Using a Novel Phase-Transfer-Catalyzed Aza-Michael Reaction. Org. Process Res. Dev. 2016, 20, 1097–1103. [Google Scholar] [CrossRef]
  126. Chung, C.K.; Liu, Z.; Lexa, K.W.; Andreani, T.; Xu, Y.; Ji, Y.; DiRocco, D.A.; Humphrey, G.R.; Ruck, R.T. Asymmetric Hydrogen Bonding Catalysis for the Synthesis of Dihydroquinazoline-Containing Antiviral, Letermovir. J. Am. Chem. Soc. 2017, 139, 10637–10640. [Google Scholar] [CrossRef]
  127. Wang, P.S.; Shen, M.L.; Wang, T.C.; Lin, H.C.; Gong, L.Z. Access to Chiral Hydropyrimidines through Palladium-Catalyzed Asymmetric Allylic C-H Amination. Angew. Chem. 2017, 56, 16032–16036. [Google Scholar] [CrossRef] [PubMed]
  128. Ng Teresa, I.; Krishnan, P.; Pilot-Matias, T.; Kati, W.; Schnell, G.; Beyer, J.; Reisch, T.; Lu, L.; Dekhtyar, T.; Irvin, M.; et al. In Vitro Antiviral Activity and Resistance Profile of the Next-Generation Hepatitis C Virus NS5A Inhibitor Pibrentasvir. Antimicrob. Agents Chemother. 2017, 61, e02558-16. [Google Scholar] [CrossRef] [PubMed]
  129. Degoey, D.A.; Kati, W.M.; Hutchins, C.W.; Donner, P.L.; Krueger, A.C.; Randolph, J.T.; Motter, C.E.; Nelson, L.T.; Patel, S.V.; Matulenko, M.A.; et al. Anti-Viral Compounds. Patent WO 2012051361, 19 April 2012. [Google Scholar]
  130. Heo, Y.-A.; Deeks, E.D. Sofosbuvir/Velpatasvir/Voxilaprevir: A Review in Chronic Hepatitis C. Drugs 2018, 78, 577–587. [Google Scholar] [CrossRef]
  131. Bjornson, K.; Canales, E.; Cotell, J.; Karki, J.K.; Katana, K.A.; Kato, A.D.; Kobayashi, T.; Link, J.O.; Martinez, R.; Phillips, B.; et al. Inhibitors of Hepatitis C Virus. Patent WO 2014008285, 9 January 2014. [Google Scholar]
  132. Mullard, A. 2016 FDA drug approvals. Nat. Rev. Drug Discov. 2017, 16, 73–76. [Google Scholar] [CrossRef]
  133. Musella, A.; Bardhi, E.; Marchetti, C.; Vertechy, L.; Santangelo, G.; Sassu, C.; Tomao, F.; Rech, F.; D’Amelio, R.; Monti, M.; et al. Rucaparib: An emerging parp inhibitor for treatment of recurrent ovarian cancer. Cancer Treat. Rev. 2018, 66, 7–14. [Google Scholar] [CrossRef]
  134. Thomas, H.D.; Calabrese, C.R.; Batey, M.A.; Canan, S.; Hostomsky, Z.; Kyle, S.; Maegley, K.A.; Newell, D.R.; Skalitzky, D.; Wang, L.-Z.; et al. Preclinical selection of a novel poly(ADP-ribose) polymerase inhibitor for clinical trial. Mol. Cancer Ther. 2007, 6, 945–956. [Google Scholar] [CrossRef]
  135. Webber, S.E.; Canan-Koch, S.S.; Tikhe, J.G.; Thoresen, L.H. Tricyclic Inhibitors of Poly(ADP-Ribose) Polymerases. Patent WO 2000042040, 20 July 2000. [Google Scholar]
  136. Bonaventura, A.; Montecucco, F. Sofosbuvir/velpatasvir: A promising combination. World J. Hepatol. 2016, 8, 785–789. [Google Scholar] [CrossRef] [PubMed]
  137. Keating, G.M.; Vaidya, A. Sofosbuvir: First Global Approval. Drugs 2014, 74, 273–282. [Google Scholar] [CrossRef]
  138. Asselah, T. Sofosbuvir for the treatment of hepatitis C virus. Expert Opin. Pharmacother. 2014, 15, 121–130. [Google Scholar] [CrossRef] [PubMed]
  139. Sofia, M.J.; Bao, D.; Chang, W.; Du, J.; Nagarathnam, D.; Rachakonda, S.; Reddy, P.G.; Ross, B.S.; Wang, P.; Zhang, H.-R.; et al. Discovery of a β-d-2′-Deoxy-2′-α-fluoro-2′-β-C-methyluridine Nucleotide Prodrug (PSI-7977) for the Treatment of Hepatitis C Virus. J. Med. Chem. 2010, 53, 7202–7218. [Google Scholar] [CrossRef] [PubMed]
  140. Wang, P.; Chun, B.-K.; Rachakonda, S.; Du, J.; Khan, N.; Shi, J.; Stec, W.; Cleary, D.; Ross, B.S.; Sofia, M.J. An Efficient and Diastereoselective Synthesis of PSI-6130: A Clinically Efficacious Inhibitor of HCV NS5B Polymerase. J. Org. Chem. 2009, 74, 6819–6824. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of representative fluorinated heterocyclic drugs approved by FDA, with the year of approval in brackets.
Figure 1. Structures of representative fluorinated heterocyclic drugs approved by FDA, with the year of approval in brackets.
Ijms 24 07728 g001
Figure 2. Representative reaction schemes for introduction of F atoms into organic molecules. Introduction of (a) two, (b) one or (c) three fluorine atoms into alkyl chains. Introduction of (d) one fluorine or (e) trifluoromethyl into aromatics.
Figure 2. Representative reaction schemes for introduction of F atoms into organic molecules. Introduction of (a) two, (b) one or (c) three fluorine atoms into alkyl chains. Introduction of (d) one fluorine or (e) trifluoromethyl into aromatics.
Ijms 24 07728 g002
Scheme 1. Synthesis of Lenacapavir, part (I–II). (I) Synthesis of intermediate 6; (II) Synthesis of intermediate 9.
Scheme 1. Synthesis of Lenacapavir, part (I–II). (I) Synthesis of intermediate 6; (II) Synthesis of intermediate 9.
Ijms 24 07728 sch001
Scheme 2. Synthesis of Lenacapavir, part (III).
Scheme 2. Synthesis of Lenacapavir, part (III).
Ijms 24 07728 sch002
Scheme 3. Synthesis of Oteseconazole.
Scheme 3. Synthesis of Oteseconazole.
Ijms 24 07728 sch003
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

Rizzo, C.; Amata, S.; Pibiri, I.; Pace, A.; Buscemi, S.; Palumbo Piccionello, A. FDA-Approved Fluorinated Heterocyclic Drugs from 2016 to 2022. Int. J. Mol. Sci. 2023, 24, 7728. https://doi.org/10.3390/ijms24097728

AMA Style

Rizzo C, Amata S, Pibiri I, Pace A, Buscemi S, Palumbo Piccionello A. FDA-Approved Fluorinated Heterocyclic Drugs from 2016 to 2022. International Journal of Molecular Sciences. 2023; 24(9):7728. https://doi.org/10.3390/ijms24097728

Chicago/Turabian Style

Rizzo, Carla, Sara Amata, Ivana Pibiri, Andrea Pace, Silvestre Buscemi, and Antonio Palumbo Piccionello. 2023. "FDA-Approved Fluorinated Heterocyclic Drugs from 2016 to 2022" International Journal of Molecular Sciences 24, no. 9: 7728. https://doi.org/10.3390/ijms24097728

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

Article Metrics

Back to TopTop