Next Article in Journal
Progressive Mitochondrial SOD1G93A Accumulation Causes Severe Structural, Metabolic and Functional Aberrations through OPA1 Down-Regulation in a Mouse Model of Amyotrophic Lateral Sclerosis
Next Article in Special Issue
Anti-Cancer and Electrochemical Properties of Thiogenistein—New Biologically Active Compound
Previous Article in Journal
Novel Nontoxic 5,9-Disubstituted SN38 Derivatives: Characterization of Their Pharmacological Properties and Interactions with DNA Oligomers
Previous Article in Special Issue
In Vitro and In Silico Characterization of an Antimalarial Compound with Antitumor Activity Targeting Human DNA Topoisomerase IB
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Design and Synthesis of Fluoro Analogues of Vitamin D

Faculty of Pharmaceutical Sciences, Teikyo University, 2-11-1 Kaga, Itabashi, Tokyo 173-8605, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(15), 8191; https://doi.org/10.3390/ijms22158191
Submission received: 8 July 2021 / Revised: 27 July 2021 / Accepted: 27 July 2021 / Published: 30 July 2021

Abstract

:
The discovery of a large variety of functions of vitamin D3 and its metabolites has led to the design and synthesis of a vast amount of vitamin D3 analogues in order to increase the potency and reduce toxicity. The introduction of highly electronegative fluorine atom(s) into vitamin D3 skeletons alters their physical and chemical properties. To date, many fluorinated vitamin D3 analogues have been designed and synthesized. This review summarizes the molecular structures of fluoro-containing vitamin D3 analogues and their synthetic methodologies.

Graphical Abstract

1. Introduction

Fluorine is one of the halogens, known as a small and the most electronegative element. Fluorine substitution offers a variety of advantages, such as changing the pKa and dipole moment of the molecule, improving the chemical or metabolic stability, and enhancing the binding affinity to the target protein. Furthermore, it has a small atomic radius, similar to that of a hydrogen atom. Due to their unique properties, fluorine atoms have been incorporated into many drugs, drug candidates, and agricultural chemicals [1,2,3,4,5,6,7,8,9,10,11,12]. The contribution of fluorine to drug development and medicinal chemistry, and the life science as well as material science fields, is widely recognized around the world [13,14,15,16,17]. Many scientists have been engaged in the practical synthesis of organofluorine compounds, including fluorinated vitamin D3 analogues. Vitamin D3 (VD3) is a fat-soluble vitamin whose biological function depends on metabolic activation by CYP enzymes [18]. Bioactivation of VD3 requires sequential oxidation steps at C-25 and C-1 catalyzed by vitamin D 25-hydroxylase (CYP2R1) and 25-hydroxyvitamin D3-1α-hydroxylase (CYP27B1), respectively. The resulting B-secosterol, 1α,25-dihydroxyvitamin D3 [1α,25(OH)2D3 (1)], is the fully active and hormonal form of VD3. Moreover, 1α,25(OH)2D3 (1) and 25-hydroxyvitamin D3 [25(OH)D3 (2)] are degraded via hydroxylation at C23 or C24 catalyzed by 1α,25-dihydroxyvitamin D3-24-hydroxylase (CYP24A1) [19]. C23 hydroxylation and subsequent three-step oxidation lead to vitamin D3-26,23-lactone (3). On the other hand, C24 hydroxylation and subsequent five-step oxidation lead to calcitroic acid (4) (Scheme 1) [19,20].
For slowing or preventing the biological degradation of the VD3 side chain, replacing C-H with C-F bond(s) at appropriate positions should prolong their half-life in vivo, since a C-F bond is stronger than a C-H bond chemically. The introduction of fluorine atoms into VD3 analogues can also alter electron distribution, which can confer lower pKa at the hydroxy group(s), change the dipole moment, and influence the conformation because of their marked electron-withdrawing properties. Because of these unique properties, both academic institutions and industries have designed and synthesized numerous fluorinated VD3 analogues, similarly to other bioactive compounds with fluorine atoms. Most of them show fluorination at the main metabolic site(s) and/or neighboring hydroxy group of the VD3 molecule, namely either an A-ring or a side-chain moiety or both. This review introduces fluorinated VD3 analogues and their synthetic methodologies, including some basic biological activities.

2. A-Ring Fluorinated VD3 Analogues

On the VD3 A-ring, 1α-hydroxylation is the final and essential step to produce the hormonal form of VD3 from 25(OH)D3 (2), and 1α,25(OH)2D3 (1) exerts a range of physiological activities by binding to the vitamin D receptor (VDR). The A-ring moiety is anchored in the VDR ligand-binding pocket through four hydrogen-bonding interactions, i.e., the 1α-hydroxy group to Ser233 and Arg270, and the 3β-hydroxy group to Tyr143 and Ser274 [21]. As expected, based on the importance of the A-ring moiety, many A-ring fluorinated VD3 analogues have been synthesized and evaluated regarding their biological activities.

2.1. 1-Fluorinated VD3 Analogues

DeLuca and coworkers described the first synthesis of 1-fluoro-VD3 for the purpose of studying the possibility of using it as a kind of VD3 antagonist against 1α-hydroxylase in 1979 [22]. They prepared 1α-hydroxyvitamin D3-3-acetate by the selective acetylation of 1α-hydroxyvitamin D3 [1α(OH)D3 (5)] and used it as a starting material. The fluorination step was achieved by N,N-(diethylamino)sulfur trifluoride (DAST) to afford 1-fluorovitamin D3 (6) (Scheme 2). The authors did not assign its C1 configuration in the report. The biological evaluation revealed that 6 demonstrated a relative preference for stimulating bone calcium mobilization with respect to intestinal calcium transport after metabolism in vivo, and 6 was a weak agonist that could not be used as an anti-vitamin D agent.
Next, DeLuca et al. designed and synthesized 1α,25-difluorovitamin D3 (7) in 1981 [23]. As shown in Scheme 3, it was synthesized from either 1α,25(OH)2D3 3-acetate (8) or 1α,25(OH)2-3,5-cyclovitamin D3 (9) using DAST as a fluorination reagent. The authors explained that the stereochemistry at C1 was 1α in the report, but they revised it to 1β in 1984 [24].
The authors evaluated the biological properties of 1α,25-difluorovitamin D3 (7) and demonstrated that it had essentially no vitamin D activity. These results strongly supported the idea that C1 and C25 hydroxylation are essential aspects of vitamin D function. Initially, since both the C1 and C25 positions of VD3 were blocked with fluorine atoms, it might have shown the anti-vitamin D activity of 25-hydroxylation of VD3 in vivo. However, 7 did not exhibit the expected inhibitory activity against the 25-hydroxylation of VD3.
In 1984, 1α-fluorovitamin D3 (10) and 1α-fluoro-25(OH)D3 (11) were synthesized utilizing a steroid A-ring epoxide as the starting material in order to confirm the stereochemistry at C1 (Scheme 4) [24,25].
A comparison of the spectral data of the report [24] and previous ones [22,23] revealed that the reported 1α-fluorovitamin D3 in 1979 [22] and 1α,25-difluorovitamin D3 in 1981 [23] were, in fact, 1β-fluorovitamin D3 and 1β,25-difluorovitamin D3, respectively.
On the other hand, although 1α-fluoro-25(OH)D3 (11) showed no stimulation of intestinal calcium transport or bone calcium mobilization activities at a dosage level of 1.3 μg, its binding affinity to chick intestine VDR was 30 times greater than that of 25(OH)D3.
A convergent synthetic route to 1-fluoro-VD3 analogues was described by Uskoković and coworkers using an A-ring key fragment, a 1α-fluorinated A-ring precursor (12), which was prepared from (S)-(+)-carvone (13), in 1990 (17 steps in 4%) (Scheme 5) [26], as well as an A ring from VD3 in 1991 (12 steps) (Scheme 6) [27]. A lithium anion of the A-ring phosphine oxide (12) underwent a Wittig–Horner coupling reaction with the 8-keto-CD-ring (14) to afford the desired 1α-fluoro-25(OH)D3 (11).
Later, Uskoković’s group utilized the 1α-fluoro-A-ring (12) to synthesize six 1α-fluoro-VD3 analogues with two different side chains at C20 (Gemini analogues). They prepared six 8-keto-CD-rings (16–21) starting from the methyl ester (15) and coupled it with 12 under basic conditions (Scheme 7) [28]. The anticancer activity of these compounds was tested, but 1α-fluorination did not effectively promote the activity.
In 2019, Uesugi and colleagues published the first synthesis of 1-fluoro-19-norVD3 analogues [29]. They constructed 1-hydroxy-19-norvitamin D3 structures with a modified Julia olefination method [30], and the direct deoxyfluorination of C1 yielded the corresponding 1-fluorinated analogues (22,23). In this reaction sequence, shown in Scheme 8, they also obtained 3-fluoro-19-norVD3 analogues (24,25). Some of these analogues were poor VDR binders but showed potent sterol regulatory element-binding protein (SREBP) inhibitory activity via inducing SREBP cleavage-activating protein (SCAP) degradation [29].

2.2. 2-Fluorinated VD3 Analogues

Several 2-fluorinated VD3 analogues have been reported to date, because fluorine substitution at this position can change the A-ring conformation and pKa value of the neighboring 1α- and 3β-hydroxy groups. The first synthesis of 2β-fluoro-VD3 was reported by Ikekawa and coworkers in 1980, in which nucleophilic fluorination using KHF2 to 1,2α-epoxycholesterol gave 2β-fluoro-1α(OH)D3 (26), (Scheme 9) [31]. It was noted that the biological activity of 26 was increased [32].
Next, 2α-fluorovitamin D3 (27) was also synthesized, and its biological activity was tested in 1986 [32]. Electrophilic 2α-fluorination of 3,6β-diacetoxycholest-2-ene (28) using CsSO4F gave 2α-fluoroketone (29). To construct the B-secosteroidal structure, conventional photochemical conversion and subsequent thermal isomerization were applied (Scheme 10) [32]. The biological effects of 27 on intestinal calcium transport and bone calcium mobilization as well as the serum calcium concentration at a dosage level of 500 ng for rats were measured, and the activities were found to be essentially equivalent to those of VD3 itself.
The synthesis and biological activities of 2β-fluoro-1α,25-dihydroxyvitamin D3 (30) were reported by Scheddin et al. in 1998 [33]. The synthetic route was similar to Ikekawa’s [31], as shown in Scheme 11, and the biological evaluation revealed that the synthetic 2β-fluoro-1α,25-(OH)2D3 (30) exhibited greater potency in vitro, for example, six-times higher affinity for VDR, nearly identical affinity for the vitamin D-binding protein (DBP), and 90-times higher antiproliferative activity toward C3H10T1/2 cells under serum-containing conditions, as well as five-times greater adipogenesis inhibitory activity than the natural hormone 1α,25(OH)2D3 (1).
The catalytic asymmetric stereoselective synthesis of the A-ring precursor of the 19-nor type 2α-fluorovitamin D3 analogue (31) and its synthesis were reported by Mikami et al. [34,35]. The regio- and stereo-selective 2α-fluorination was achieved via a ring opening reaction of chiral epoxide (32) mediated by HfF4/Bu4NH2F3, the asymmetric catalytic carbonyl-ene cyclization was used to construct the 6-membered A-ring precursor, and the subsequent coupling reaction with the CD ring afforded 2α-fluoro-19-normaxacalcitol (31) (Scheme 12). This 2α-fluoro-22-oxa-19-nor analogue (31) had very low DBP-binding affinity but four-times stronger VDR-binding potency than its 22-oxa-19-nor counterpart and also showed significant transactivation activity [34]. It was also shown that 31 was highly effective in inhibiting metastatic tumor growth in vivo without toxicity in terms of hypercalcemia and weight loss [35].
Posner’s group showed the first example of synthesis of the 2,2-difluorovitamin D3 analogue in 2002 [36]. They synthesized the racemic 2,2-difluoro substituted A-ring phosphine oxide (33) from trifluoroethanol (7% in 13 steps). A coupling reaction of 33 with the 8-keto-CD-ring and subsequent deprotection yielded the 2,2-difluoro-1,25-dihydroxyvitamin D3 analogues in the ratio of 5:1 (1α,3β (34):1β,3α) (Scheme 13). The diastereomers were separated with reversed-phase HPLC to afford the target 2,2-difluoro-1α,25(OH)2D3 (34). Biological evaluation revealed that 34 exhibited antiproliferative activity similar to that of 1α,25(OH)2D3 (1) and was 2-3 times more transcriptionally active than 1 in rat osteosarcoma cells, even though the human VDR-binding affinity was 9.6% relative to that of 1. Compound 34 showed strong calcemic activity in vivo.

2.3. 3-Fluorinated VD3 Analogues

The C3 position of VD3 has a β-hydroxy group, and substitution of the hydroxy with a fluorine atom is of fundamental interest. There are several reports of replacing the C3-hydroxy group with a fluorine atom in order to demonstrate the expected positive effects on biological activity.
The synthesis of 3β-fluoro-3-deoxyvitamin D3 (35) from 3β-fluorocholesta-5,7-diene (36) via photochemical transformation followed by thermal isomerization was described by Segal et al. in 1976 [37]. It was found that 3β-fluoro-3-deoxyvitamin D3 (35) had an antirachitic effect analogous to VD3. On the other hand, in 1978, Mazur and coworkers designed and synthesized 3β-fluoro-3-deoxy-1α-hydroxyvitamin D3 (37) to elicit VD3 activity. The 3β-fluoro group was constructed from (6R)-hydroxy-3,5-cyclovitamin D3 (38) by treatment with HF (Scheme 14) [38]. The biological activities of the newly synthesized analogue were tested, and the fluoro-analogue (37) could actively induce the formation of a calcium-binding protein and stimulate intestinal calcium absorption in rachitic chicks.
Later, in 1985, Kumar and coworkers reported the synthetic route to 3β-fluoro-3-deoxyvitamin D3 (35) starting from 7-dehydrocholesterol (39) via 3-deoxy-3-fluoro-7-dehydrocholesterol (36) [39]. The B-ring protected PTAD (4-phenyl-1,2,4-triazoline-3,5-dione) adduct (40) was reacted with DAST to yield a fluorinated product (41), and subsequent deprotection of 41 gave 3-deoxy-3-fluoro-7-dehydrocholesterol (36) (Scheme 15). To investigate the influence of the replacement of the C3-hydroxy group with fluorine, they compared the biological activities of VD3, 3-deoxyvitamin D3, and 3β-fluoro-3-deoxyvitamin D3 (35) and revealed that 35 was less active than VD3 and more active than 3-deoxyvitamin D3 in terms of intestinal calcium transport and bone calcium mobilization in vivo.
As mentioned in Section 2.1, Uesugi and colleagues reported the first synthesis of 3-fluoro-19-norVD3 analogues and evaluated their SREBP inhibitory activity vs. VDR activity (see Scheme 8) [29].

2.4. 4-Fluorinated VD3 Analogues

Yamada and coworkers reported 4,4-difluorovitamin D3 (42) and 4,4-difluoro-1α,25(OH)2D3 (43) in an A-ring conformational study in 1999, and these analogues were synthesized starting with enones (44,45), which were constructed from ergosterol, respectively [40]. The difluorination step was achieved by electrophilic fluorination under thermodynamic conditions (Scheme 16). The binding affinity of 4,4-difluoro-1α,25(OH)2D3 (43) for VDR was only ca. 1% of that of the natural hormone, 1α,25(OH)2D3 (1).

3. Introduction of a Fluorine Atom into the Triene Part of VD3

3.1. 19-Fluorinated VD3 Analogues

In 1980, the first attempt to synthesize 19,19-difluorovitamin D3 (46) starting from 19-oxocholesteryl acetate (47) via photoirradiation and thermal isomerization reactions using diene (48) failed, because the final thermal isomerization step by [1,7]-sigmatropic rearrangement did not proceed in the presence of the difluoromethyl group at the C10 position (Scheme 17) [41].
Later, in 1996, Yamada and coworkers showed a novel synthetic route to the first (10E)- and (10Z)-19-fluorovitamin D3 (50,51) [42]. They prepared VD3-SO2 adducts (49) from VD3, and fluorination at the C19 position was achieved by electrophilic fluorination using (PhSO2)2NF in the presence of LiHMDS (Scheme 18A). In 2000, the same group showed the regioselective introduction of a fluorine atom to the C19 position and synthesized both (10E)- and (10Z)-19-fluoro-1α,25(OH)2D3 (52,53) [43]. The binding affinity of (10Z)-19-fluoro-1α,25(OH)2D3 for VDR was ca. 10% compared with that of 1α,25(OH)2D3 (1). The alternative synthetic route to 52 and 53 from a 10-oxo-19-norVD3 derivative was also established by the same group in 2001 (Scheme 18B) [44].

3.2. 6- and 7-Fluorinated VD3 Analogues

The synthesis of 6-fluorovitamin D3 (54) was described by Dauben et al. in 1985, and the synthetic route started from 6-oxo-cholestanyl acetate (55) [45]. The key intermediate, 6-fluoro-7-dehydrocholesteryl acetate (56), was synthesized by allowing 55 to react with piperidinosulfur trifluoride in the presence of sulfuric acid. To construct the triene system, 6-fluoro-7-dehydrocholesterol was irradiated, followed by thermal [1,7]-sigmatropic hydrogen rearrangement, to give the desired 6-fluorovitamin D3 (54) (Scheme 19). The obtained 6-fluorovitamin D3 (54) was air-sensitive, and decomposition proceeded. The biological profile of the analogue was evaluated in vivo, revealing that 54 had no biological effect on either intestinal calcium absorption or bone calcium mobilization. However, it significantly inhibited both VD3- and 1α,25(OH)2D3-mediated intestinal calcium absorption through a direct interaction with VDR [46].
Furthermore, 6-fluoro-1α,25-dihydroxy-19-norvitamin D3 (57) and 7-fluoro-1α,25-dihydroxy-19-norvitamin D3 (58) were synthesized by the Teijin research group in 2004 [47]. Takenouchi et al. prepared a C6-fluorinated A-ring (59) and C7-fluorinated CD-ring (60), respectively. A Ni-catalyzed cross-coupling reaction with each CD-ring- or A-ring-activated alkene counterpart was used to construct the diene structures of 57 and 58 (Scheme 20). These compounds possessed 10–70% VDR binding affinity of that of 1α,25(OH)2D3 and potential activity to induce HL-60 cell differentiation.

4. CD-Ring Fluorinated VD3 Analogues: 11-Fluorinated VD3 Analogues

To our knowledge, only one report has been published on the synthesis of CD-ring fluoro-VD3 analogues. In 1994, De Clercq and coworkers designed and synthesized 11α- and 11β-fluorovitamin D3 analogues (6164) with the aim of inducing a conformational change from s-trans to s-cis at the C6-C7 single bond via expected hydrogen bond formation between the C11-F and C1α-OH groups [48]. Enone (65), readily available from Grundmann’s ketone, was used as a starting material. Epoxidation of 65, followed by reductive opening with lithium dimethylcuprate, gave the C11α-OH functional group (66). After coupling with A-ring phosphine oxides, the C11α-OH group was treated with N-(2-chloro-1,1,2-trifluoroethyl)diethylamine (FAR) to give C11α-F and C11β-F isomers as well as an elimination product at a ratio of 3:1:1 (Scheme 21). NMR spectra analyses of the 11-fluoro-1α(OH)D3 showed that all analogues had a large J6-7 coupling constant, reflecting an almost exclusive s-trans extended geometry without intramolecular hydrogen bonding between C11-F and 1α-OH.

5. Side-Chain Fluorinated VD3 Analogues

The CYP24A1 pathway is well-known as the deactivation pathway of both 25(OH)D3 (2) and 1α,25(OH)2D3 (1) [19]. Varieties of side-chain-fluorinated VD3 analogues have been actively synthesized because of the expected slower catabolism resulting from the presence of fluorine atoms at the oxidation site or adjacent area.

5.1. 22-Fluorinated VD3 Analogues

In 1986, Kumar and coworkers described the synthesis of 22-fluorovitamin D3 (67) starting from (22S)-cholest-5-ene-3β,22-diol (68) [49]. Selective protection of the C3 hydroxy group as an acetate, followed by fluorination at the C22 position by DAST, gave 22-fluorocholest-5-en-3β-acetate (69). After forming the C5-7 diene unit in the B ring, photolysis and thermal isomerization yielded the target 22-fluorovitamin D3 (67) without assigning C22 stereochemistry (Scheme 22). They tested the biological activities of the analogue in vitro and in vivo, referring to the potency of intestinal calcium transport, serum calcium level, calcium-binding protein induction, plasma vitamin D-binding protein (DBP), and VDR affinities, and concluded that the introduction of a fluorine atom to C22 resulted in the compound, with weak biological activities and poor binding to DBP compared with VD3 itself.

5.2. 23-Fluorinated VD3 Analogues

As mentioned in the Introduction, the C23 position of VD3 is one of the essential metabolic sites of CYP24A1; therefore, C23-fluorinated VD3 analogues have been designed and synthesized based on the idea of blocking the oxidative position. Ikekawa and colleagues achieved the first synthesis of a C23-fluoro-VD3 analogue, 23,23-difluoro-25(OH)D3 (70), in 1984 [50]. For the synthesis of 70, the triene structure was constructed by applying the well-established route through 5,7-diene steroids, and the difluoro unit was introduced using DAST into a reactive α-ketoester (71) (Scheme 23).
The same group continued to report 23,23-difluoro-1α,25(OH)2D3 (72) by the enzymatic 1α-hydroxylation of 70 in 1985 [51]. Their biological activities were evaluated, and 23,23-difluoro-25(OH)D3 (70) was 5-10 times less active than 25(OH)D3 in stimulating intestinal calcium transport, bone calcium mobilization, mineralization of rachitic bone, etc., and 23,23-difluoro-1α,25(OH)2D3 (72) was one-seventh as active as 1α,25(OH)2D3 in binding to VDR.
Ikeda and coworkers of the Sumitomo research group reported the synthesis of (23R)-23,26,26,26,27,27,27-heptafluoro-1α,25(OH)2D3 (73) and its 23S isomer (74) in 2000 [52]. The starting methyl ketone was available from VD2, and a subsequent hexafluoroacetone (HFA) aldol reaction gave a 23-oxo derivative, which was reduced to 23R- and 23S-secondary alcohols that could be separated and subjected to deoxyfluorination using DAST to afford 73 and 74 (Scheme 24). Both analogues showed higher VDR-binding affinity and HL-60 cell differentiation activity than falecalcitriol.
Later, in 2019, our group synthesized (23R)-23-fluoro-25(OH)D3 (75) and its 23S-isomer (76) starting from the Inhoffen–Lythgoe diol via the key intermediate 23-hydroxy-CD- rings (77,78) (Scheme 25) [53]. The preliminary biological evaluation revealed that the 23S-isomer (76) showed higher resistance to CYP24A1 metabolism than its 23R-isomer (75).

5.3. 24-Fluorinated VD3 Analogues

Since oxidation of the hydroxy function at the C24 position catalyzed by CYP24A1 is one of the important pathways to deactivate both 25(OH)D3 and 1α,25(OH)2D3, developing practical methods to construct the C24-fluoro unit on the VD3 skeleton has been pursued since 1979. The first synthesis of the 24,24-difluorovitamin D3 analogue was reported independently by Takayama’s group [54] and Kobayashi-Ikekawa’s group [55]. Both synthetic routes involved the key intermediate (79), and the two groups synthesized the same analogue, 24,24-difluoro-25(OH)D3 (80). For the construction of the 24,24-difluoro unit, Takayama and coworkers utilized the reaction of α-ketoester (81), which was derived from lithocholic acid, with DAST. On the other hand, Kobayashi et al. used the reaction of steroidal enol ether, derived from cholic acid, with difluorocarbene (Scheme 26).
In 1980, Kobayashi-DeLuca’s group subsequently demonstrated that kidney homogenates from the chicken converted 24,24-difluoro-25(OH)D3 (80) to 24,24-difluoro-1α,25(OH)2D3 (82) [56]. The results of the biological evaluation revealed that 82 and its nonfluorinated counterpart 1α,25(OH)2D3 (1) equipotently stimulate intestinal calcium transport and bone calcium mobilization in vivo, while in another in vitro system, 82 was found to be four times more potent than 1α,25(OH)2D3 (1).
In 1990, Kumar’s group synthesized 24,24-difluoro-25-hydroxy-26,27-dihomovitamin D3 (83) and its 1α-hydroxy analogue (84) from 3β-hydroxy-22,23-dinorcholenic acid using a Reformatsky reaction in their synthetic route (Scheme 27) [57]. Both analogues showed similar biological activities, such as intestinal calcium transport and bone calcium mobilization in vivo, to those of the respective nonfluorinated counterparts and also 25(OH)D3 or 1α,25(OH)2D3.
In 1992, two alternative linear synthetic routes to 24,24-difluoro-1α,25(OH)2D3 (82) were reported by Takayama et al. using the Reformatsky reaction with ethyl bromodifluoroacetate or Horner–Emmons reaction as the key step, respectively (Scheme 28) [58,59]. The starting material of the former was 1α-hydroxydehydroepiandrosterone, with a 3.8% overall yield of 82 [58], and that of the latter was vitamin D2, with a 9.3% overall yield of 82 [59].
As shown in Scheme 29A, the same group synthesized 24,24-difluoro-1α(OH)D3 (85) from a steroidal skeleton in 1996 [60]. This compound showed higher activity than 24,24-difluoro-1α,25(OH)2D3 (82) in intestinal calcium absorption. In 1998, Iwasaki-Takayama’s group reported the synthesis of (25R)- and (25S)-24,24-difluoro-1α,25,26-trihydroxyvitamin D3 (86,87), including the X-ray crystallographic analysis of a synthetic intermediate to determine C25-stereochemistry, and proved that the 25S-isomer (87) was the main CYP24-metabolite of 82 (Scheme 29B) [61].
The 24,24-difluoro-19-norVD3 analogues including 20-epi-versions (8891) were reported by DeLuca et al. in 2015 [62]. In contrast to the previous synthetic methods starting from steroid skeletons, they demonstrated a convergent method utilizing the Wittig–Horner reaction between 24,24-difluoro-CD-rings (92,93) and a lithium salt of a phosphine oxide anion from A-ring precursors (Scheme 30). The 20S-derivatives showed marked bone-mobilizing activity in vivo; however, 2-methylene substitution was required for such elevated activity in the 20R series.
In 2019, our group developed the convergent synthesis of 24,24-difluoro-25(OH)D3 (80) using a coupling reaction between the 24,24-difluoro-CD ring ketone derived from the Inhoffen–Lythgoe diol via 94 and an A-ring phosphine oxide (Scheme 31A) [63]. We subsequently synthesized a novel vitamin D-based VDR-silent SREBP inhibitor, KK-052, from 94 (Scheme 31B) [64].
Ikekawa’s group reported in 1979 the first synthesis of C24-monofluoro-25(OH)D3 (95) from cholenic acid without assigning the C24 stereochemistry (Scheme 32) [55].
After this, (24R)-24-fluoro-1α,25(OH)2D3 (96) was synthesized by Uskoković’s group through linear (Scheme 33A) and convergent (Scheme 33B) synthetic routes in 1985 [65] and 1988 [66], respectively. In this case, (24R)-24-fluoro-1α,25(OH)2D3 (96) showed a longer plasma half-life and higher anti-rachitogenic activity than 1α,25(OH)2D3 in vivo.

5.4. 25-Fluorinated VD3 Analogues

The 25-hydroxylation is the initial metabolic conversion of VD3 by CYP2R1 or CYP27A1 [18], and 25(OH)D3 (2) is known as the major circulating metabolite in the human body. The first introduction of fluorine to C25, i.e., the synthesis of 25-fluoro-VD3, was reported by DeLuca et al. in 1977 (Scheme 34A) [67]. C3-Selective O-acetylation of 25(OH)D3 (2) and subsequent direct fluorination at the C25 position using DAST, followed by deacetylation, gave 25-fluorovitamin D3 (99). In 1978, Stern and coworkers synthesized 25-fluoro-1α(OH)D3 (100) using the same approach (Scheme 34B) [68]. The results of the biological evaluation indicated that 25-fluoro-1α(OH)D3 (100) was approximately equipotent to 1α(OH)D3 (5) in VDR binding and stimulating bone resorption in vitro, and the C25-fluoro substituent behaved similarly to hydrogen, without elevating its biological potency.
As described in Section 2.1, DeLuca’s group also synthesized 1,25-difluorovitamin D3 in 1981, and this compound was devoid of any biological activity [21].

5.5. 26,27-Hexafluorinated VD3 Analogues

Falecalcitriol (101) is 26,26,26,27,27,27-hexafluorinated VD3 and has been approved for therapeutic use against secondary hyperparathyroidism in Japan [69,70]. Falecalcitriol (101) was ca. 10 times more active in increasing bone calcium mobilization than the natural hormone (1) in vivo [71]. Similarly to other fluorinated VD3 analogues, it was metabolized more slowly than 1α,25(OH)2D3 (1) by CYP24A1, and interestingly, its major metabolite (23S)-23-hydroxyfalecalcitriol (102) was equipotent to the parent compound 101 in its biological activity (Figure 1) [72,73,74]. The 23-hydroxylated 102 was specifically glucuronidated by UGT1A3 [75].
The first synthesis of 26,26,26,27,27,27-hexafluoro-25(OH)D3 (103) was reported by Kobayashi et al. in 1980 [76], and the same group subsequently synthesized 26,26,26,27,27,27-hexafluoro-1α,25(OH)2D3 (101) in 1982 [77]. In this case, 3β-Tetrahydropyranyloxychol-5-en-24-ol tosylate was used as a starting material, and hexafluoroacetone (HFA) was utilized for construction of the 26,26,26,27,27,27-hexafluoro unit (Scheme 35). The biological evaluation revealed that 101 was approximately 5-10-fold more active than 1α,25(OH)2D3 (1) without inducing severe hypercalcemia.
Using the reaction of acetylides with excess HFA gas as the key step, syntheses of the hexafluoroalkynyl-VD3 analogue (104), C-seco-hexafluoroalkynyl-VD3 analogue (105), and the previously mentioned two-side-chain analogues 17-22 including alkene side chains (Section 2.1) were reported by Ohira et al. in 1992 [78], by Wu et al. in 2002 [79], and by Maehr et al. in 2009 [28], respectively (Scheme 36). Compound 104 had a potent inducing effect on the differentiation of cancer cells, with little calcium mobilization activity. Compound 105 showed comparable VDR-binding affinity to the natural hormone 1 and had strong antiproliferative activity against four cancer cell lines in vitro, with 1% calcemic activity compared with 1 in vivo. More recently, Sigüeiro et al. synthesized C22-diyne analogues with a C17-methyl group, including the hexafluoropropanol unit at the terminal (106), which showed potent VDR-binding affinity [80].
Hayashi and colleagues described the introduction of the 26,26,26,27,27,27-hexafluoro unit utilizing an aldol reaction [52]. The C23 ketone was treated with HFA in the presence of LiHMDS to afford the HFA adduct (see Scheme 24 in Section 5.2).
As an alternative path to construct the hexafluoro unit, the nucleophilic trifluoromethylation of methyl esters with Ruppert–Prakash reagent (CF3TMS) can be utilized. In our group, 26,26,26,27,27,27-hexafluoro-25(OH)D3 (103) and 26,26,26,27,27,27-hexafluoro-CD ring (107) were synthesized in 2018 starting from the methyl esters (108,110) through trifluoromethylketones (109,111) as the intermediates (Scheme 37) [81].

6. Summary

This review summarized the historical fluorinated VD3 analogues with modification from the A-ring to the end of the side-chain, including their synthetic methods. With the aim of preventing or slowing their activation or degradation by CYPs, the A-ring and side-chain have been mainly focused on, and numerous VD3 analogues containing the fluorine atom(s) have been synthesized. Hydroxylation at the C25 and C1α positions of VD3 is necessary for the activation process of the molecule by CYP2R1/CYP27A1 and CYP27B1, respectively; therefore, in general, the introduction of fluorine to these positions decreases the biological activity through VDR if compared to non-fluorinated 25(OH)D3 or 1α,25(OH)2D3 as each parent VDR ligand. On the other hand, fluorination at the side chain C23, C24, and C26(27) of VD3, where deactivating hydroxylation occurs based on CYP24A1 metabolism, produces strong VDR agonists that have a long half-life in vivo. Among them, falecalcitriol was successfully approved for the treatment of secondary hyperparathyroidism in Japan. Discovery of the new functions of VD3 continues; for example, the potent SREBP-inhibitory activity of 25(OH)D3 [82] and the new fluorinated analogues with their efficient synthetic methods may contribute to the treatment of patients with VD3-function-related disease in the future.

Author Contributions

Conceptualization, F.K. and A.K.; writing—original draft preparation, F.K. and S.M.; writing—review and editing, A.K.; supervision, A.K.; funding acquisition, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a Grant-in-Aid from the Japan Society for the Promotion of Science (No. 18K06556 to A.K.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Müller, K.; Faeh, C.; Diederich, F. Fluorine in pharmaceuticals: Looking beyond intuition. Science 2007, 317, 1881–1886. [Google Scholar] [CrossRef] [Green Version]
  2. 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] [PubMed]
  3. Böhm, H.-J.; Banner, D.; Bendels, S.; Kansy, M.; Kuhn, B.; Müller, K.; Obst-Sander, U.; Stahl, M. Fluorine in medicinal chemistry. Chembiochem 2004, 5, 637–643. [Google Scholar] [CrossRef]
  4. Kirk, K.L. Fluorine in medicinal chemistry: Recent therapeutic applications of fluorinated small molecules. J. Fluor. Chem. 2006, 127, 1013–1029. [Google Scholar] [CrossRef]
  5. Purser, S.; Moore, P.R.; Swallow, S.; Gouverneur, V. Fluorine in medicinal chemistry. Chem. Soc. Rev. 2008, 37, 320–330. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, J.; Sánchez-Roselló, M.; Aceña, J.L.; Del Pozo, C.; Sorochinsky, A.E.; Fustero, S.; Soloshonok, V.A.; Liu, H. Fluorine in pharmaceutical industry: Fluorine-containing drugs introduced to the market in the last decade (2001–2011). Chem. Rev. 2014, 114, 2432–2506. [Google Scholar] [CrossRef] [PubMed]
  7. 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. Chemistry 2019, 25, 11797–11819. [Google Scholar] [CrossRef]
  8. Bégué, J.-P.; Bonnet-Delpon, D. Recent advances (1995–2005) in fluorinated pharmaceuticals based on natural products. J. Fluor. Chem. 2006, 127, 992–1012. [Google Scholar] [CrossRef]
  9. Shah, P.; Westwell, A.D. The role of fluorine in medicinal chemistry. J. Enzym. Inhib. Med. Chem. 2007, 22, 527–540. [Google Scholar] [CrossRef] [Green Version]
  10. Berkowitz, D.B.; Karukurichi, K.R.; De La Salud-Bea, R.; Nelson, D.L.; McCune, C.D. Use of fluorinated functionality in enzyme inhibitor development: Mechanistic and analytical advantages. J. Fluor. Chem. 2008, 129, 731–742. [Google Scholar] [CrossRef] [Green Version]
  11. Maienfisch, P.; Hall, R.G. The importance of fluorine in the life science industry. Chimia 2004, 58, 93–98. [Google Scholar] [CrossRef]
  12. Prchalová, E.; Štěpánek, O.; Smrček, S.; Kotora, M. Medicinal applications of perfluoroalkylated chain-containing compounds. Future Med. Chem. 2014, 6, 1201–1229. [Google Scholar] [CrossRef]
  13. Kukhar’, V.P.; Soloshonok, V.A. Aliphatic fluorine-containing amino acids. Russ. Chem. Rev. 1991, 60, 850–864. [Google Scholar] [CrossRef]
  14. Liu, P.; Sharon, A.; Chu, C.K. Fluorinated nucleosides: Synthesis and biological implication. J. Fluor. Chem. 2008, 129, 743–766. [Google Scholar] [CrossRef] [Green Version]
  15. Isanbor, C.; O’Hagan, D. Fluorine in medicinal chemistry: A review of anti-cancer agents. J. Fluor. Chem. 2006, 127, 303–319. [Google Scholar] [CrossRef]
  16. Al-Harthy, T.; Zoghaib, W.; Abdel-Jalil, R. Importance of fluorine in benzazole compounds. Molecules 2020, 25, 4677. [Google Scholar] [CrossRef]
  17. Krafft, M.P. Fluorocarbons and fluorinated amphiphiles in drug delivery and biomedical research. Adv. Drug Deliv. Rev. 2001, 47, 209–228. [Google Scholar] [CrossRef]
  18. Sakaki, T.; Kagawa, N.; Yamamoto, K.; Inouye, K. Metabolism of vitamin D3 by cytochromes P450. Front. Biosci. 2005, 10, 119–134. [Google Scholar]
  19. Sakaki, T.; Sawada, N.; Komai, K.; Shiozawa, S.; Yamada, S.; Yamamoto, K.; Ohyama, Y.; Inouye, K. Dual metabolic pathway of 25-hydroxyvitamin D3 catalyzed by human CYP24. Eur. J. Biochem. 2000, 267, 6158–6165. [Google Scholar] [CrossRef] [Green Version]
  20. Yasuda, K.; Nishikawa, M.; Okamoto, K.; Horibe, K.; Mano, H.; Yamaguchi, M.; Okon, R.; Nakagawa, K.; Tsugawa, N.; Okano, T.; et al. Elucidation of metabolic pathways of 25-hydroxyvitamin D3 mediated by Cyp24A1 and Cyp3A using Cyp24a1 knockout rats generated by CRISPR/Cas9 system. J. Biol. Chem. 2021, 296, 100668. [Google Scholar] [CrossRef]
  21. Rochel, N.; Wurtz, J.M.; Mitschler, A.; Klaholz, B.; Moras, D. The crystal structure of the nuclear receptor for vitamin D bound to its natural ligand. Mol. Cell 2000, 5, 173–179. [Google Scholar] [CrossRef]
  22. Napoli, J.L.; Fivizzani, M.A.; Schnoes, H.K.; DeLuca, H.F. 1-Fluorovitamin D3, a vitamin D3 analogue more active on bone-calcium mobilization than on intestinal-calcium transport. Biochemistry 1979, 18, 1641–1646. [Google Scholar] [CrossRef]
  23. Paaren, H.E.; Fivizzani, M.A.; Schnoes, H.K.; DeLuca, H.F. 1α,25-Difluorovitamin D3: An inert vitamin D analog. Arch. Biochem. Biophys. 1981, 209, 579–583. [Google Scholar] [CrossRef]
  24. Ohshima, E.; Takatsuto, S.; Ikekawa, N.; De Luca, H.F. Synthesis of 1α-fluorovitamin D3. Chem. Pharm. Bull. 1984, 32, 3518–3524. [Google Scholar] [CrossRef] [Green Version]
  25. Ohshima, E.; Sai, H.; Takatsuto, S.; Ikekawa, N.; Kobayashi, Y.; Tanaka, Y.; DeLuca, H.F. Synthesis and biological activity of 1α-fluoro-25-hydroxyvitamin D3. Chem. Pharm. Bull. 1984, 32, 3525–3531. [Google Scholar] [CrossRef] [Green Version]
  26. Shiuey, S.-J.; Kulesha, I.; Baggiolini, E.G.; Uskoković, M.R. Total synthesis of 1α-fluoro-25-hydroxycholecalciferol and -ergocalciferol. J. Org. Chem. 1990, 55, 243–247. [Google Scholar] [CrossRef]
  27. Kiegiel, J.; Wovkulich, P.M.; Uskoković, M.R. Chemical conversion of vitamin D3 to its 1,25-dihydroxy metabolite. Tetrahedron Lett. 1991, 32, 6057–6060. [Google Scholar] [CrossRef]
  28. Maehr, H.; Lee, H.J.; Perry, B.; Suh, N.; Uskoković, M.R. Calcitriol derivatives with two different side chains at C-20. V. Potent inhibitors of mammary carcinogenesis and inducers of leukemia differentiation. J. Med. Chem. 2009, 52, 5505–5519. [Google Scholar] [CrossRef] [Green Version]
  29. Nagata, A.; Akagi, Y.; Asano, L.; Kotake, K.; Kawagoe, F.; Mendoza, A.; Masoud, S.S.; Usuda, K.; Yasui, K.; Takemoto, Y.; et al. Synthetic chemical probes that dissect vitamin D activities. ACS Chem. Biol. 2019, 14, 2851–2858. [Google Scholar] [CrossRef]
  30. Ono, K.; Yoshida, A.; Saito, N.; Fujishima, T.; Honzawa, S.; Suhara, Y.; Kishimoto, S.; Sugiura, T.; Waku, K.; Takayama, H.; et al. Efficient synthesis of 2-modified 1α,25-dihydroxy-19-norvitamin D3 with Julia olefination: High potency in induction of differentiation on HL-60 cells. J. Org. Chem. 2003, 68, 7407–7415. [Google Scholar] [CrossRef]
  31. Oshida, J.-I.; Morisaki, M.; Ikekawa, N. Synthesis of 2β-fluoro-1α-hydroxyvitamin D3. Tetrahedron Lett. 1980, 21, 1755–1756. [Google Scholar] [CrossRef]
  32. Kobayashi, Y.; Nakazawa, M.; Kumadaki, I.; Taguchi, T.; Ohshima, E.; Ikekawa, N.; Tanaka, Y.; DeLuca, H.F. Studies on organic fluorine compounds. L. Synthesis and biological activity of 2α-fluorovitamin D3. Chem. Pharm. Bull. 1986, 34, 1568–1572. [Google Scholar] [CrossRef] [Green Version]
  33. Scheddin, D.; Mayer, H.; Schönecker, B.; Gliesing, S.; Reichenbächer, M. Synthesis and biological activities of 2β-chloro-, 2β-fluoro-, and 2β-methoxy-1α,25-dihydroxyvitamin D3. Steroids 1998, 63, 633–643. [Google Scholar] [CrossRef]
  34. Mikami, K.; Ohba, S.; Ohmura, H.; Kubodera, N.; Nakagawa, K.; Okano, T. Asymmetric catalytic ene-cyclization approach to 2-fluoro-19-nor-1,25-dihydroxyvitamin D3 A-ring analog with significant transactivation activity. Chirality 2001, 13, 366–371. [Google Scholar] [CrossRef] [PubMed]
  35. Nakagawa, K.; Okano, T.; Ozono, K.; Kato, S.; Kubodera, N.; Ohba, S.; Itoh, Y.; Mikami, K. Catalytic asymmetric synthesis and anticancer effects of the novel non-calcemic analog of vitamin D, 2α-fluoro-19-nor-22-oxa-1α,25-dihydroxyvitamin D3 in metastatic lung carcinoma. J. Fluor. Chem. 2007, 128, 654–667. [Google Scholar] [CrossRef]
  36. Posner, G.H.; Woodard, B.T.; Crawford, K.R.; Peleg, S.; Brown, A.J.; Dolan, P.; Kensler, T.W. 2,2-Disubstituted analogues of the natural hormone 1α,25-dihydroxyvitamin D3: Chemistry and biology. Bioorg. Med. Chem. 2002, 10, 2353–2365. [Google Scholar] [CrossRef]
  37. Yakhimovich, R.I.; Klimashevskii, V.M.; Segal, G.M. Synthesis and biological activity of 3α-fluoro-9,10-secocholesta-5,7,10(19)-triene, the fluoro derivative of vitamin D3. Khimiko-Farmatsevticheskii Zhurnal 1976, 10, 58–64. [Google Scholar] [CrossRef]
  38. Sheves, M.; Sialom, B.; Mazur, Y. Stereoselective conversion of vitamin D3 into its 3β-halogenated derivatives. The synthesis of a 1α-hydroxy-3β-fluorovitamin D3 analogue. J. Chem. Soc. Chem. Commun. 1978, 554–555. [Google Scholar] [CrossRef]
  39. Revelle, L.K.; Londowski, J.M.; Kost, S.B.; Corradino, R.A.; Kumar, R. Synthesis and biological activity of 3β-fluorovitamin D3: Comparison of the biological activity of 3β-fluorovitamin D3 and 3-deoxyvitamin D3. J. Steroid Biochem. 1985, 22, 469–474. [Google Scholar] [CrossRef]
  40. Shimizu, M.; Iwasaki, Y.; Yamada, S. 4,4-Difluoro-1α,25-dihydroxyvitamin D3: Analog to probe A-ring conformation in vitamin D-receptor complex. Tetrahedron Lett. 1999, 40, 1697–1700. [Google Scholar] [CrossRef]
  41. Sialom, B.; Mazur, Y. Influence of fluorine and oxygen atoms at C-19 on the previtamin D-vitamin D interconversion. J. Org. Chem. 1980, 45, 2201–2204. [Google Scholar] [CrossRef]
  42. Iwasaki, Y.; Shimizu, M.; Hirosawa, T.; Yamada, S. Regioselective synthesis of 19-fluorovitamin D via fluorination of vitamin D-sulfur dioxide adducts. Tetrahedron Lett. 1996, 37, 6753–6754. [Google Scholar] [CrossRef]
  43. Shimizu, M.; Iwasaki, Y.; Ohno, A.; Yamada, S. Synthesis of (10Z)- and (10E)-19-fluoro-1α,25-dihydroxyvitamin D3: Compounds to probe vitamin D conformation in receptor complex by 19F-NMR. Chem. Pharm. Bull. 2000, 48, 1484–1493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Shimizu, M.; Ohno, A.; Yamada, S. (10Z)- and (10E)-19-fluoro-1α,25-dihydroxyvitamin D3: An improved synthesis via 19-nor-10-oxo-vitamin D. Chem. Pharm. Bull. 2001, 49, 312–317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Dauben, W.G.; Kohler, B.; Roesle, A. Synthesis of 6-fluorovitamin D3. J. Org. Chem. 1985, 50, 2007–2010. [Google Scholar] [CrossRef]
  46. Wilhelm, F.; Dauben, W.G.; Kohler, B.; Roesle, A.; Norman, A.W. 6-Fluoro-vitamin D3: A new antagonist of the biological actions of vitamin D3 and its metabolites which interacts with the intestinal receptor for 1α,25(OH)2-vitamin D3. Arch. Biochem. Biophys. 1984, 233, 127–132. [Google Scholar] [CrossRef]
  47. Takenouchi, K.; Ishizuka, S.; Miura, D.; Sato, F.; Hanazawa, T. 6,7-Substituted 19-norvitamin D3 derivatives and pharmaceuticals containing them. Jpn. Kokai Tokkyo Koho. JP 2004-175763. 2004. Available online: https://patents.google.com/patent/US7530024 (accessed on 26 July 2021).
  48. Zhu, G.-D.; Haver, D.V.; Jurriaans, H.; De Clercq, P.J. 11-Fluoro-1α-hydroxyvitamin D3: The quest for experimental evidence of the folded vitamin D conformation. Tetrahedron 1994, 50, 7049–7060. [Google Scholar] [CrossRef]
  49. Gill, H.S.; Londowski, J.M.; Corradino, R.A.; Kumar, R. The synthesis and biological activity of 22-fluorovitamin D3: A new vitamin D analog. Steroids 1986, 48, 93–108. [Google Scholar] [CrossRef]
  50. Taguchi, T.; Mitsuhashi, S.; Yamanouchi, A.; Kobayashi, Y.; Sai, H.; Ikekawa, N. Synthesis of 23,23-difluoro-25-hydroxyvitamin D3. Tetrahedron Lett. 1984, 25, 4933–4936. [Google Scholar] [CrossRef]
  51. Nakada, M.; Tanaka, Y.; DeLuca, H.F.; Kobayashi, Y.; Ikekawa, N. Biological activities and binding properties of 23,23-difluoro-25-hydroxyvitamin D3 and its 1α-hydroxy derivative. Arch. Biochem. Biophys. 1985, 241, 173–178. [Google Scholar] [CrossRef]
  52. Ikeda, M.; Matsumura, H.; Sawada, N.; Hashimoto, K.; Tanaka, T.; Noguchi, T.; Hayashi, M. Synthesis and biological evaluations of C-23-modified 26,26,26,27,27,27-F6-vitamin D3 analogues. Bioorg. Med. Chem. 2000, 8, 1809–1817. [Google Scholar] [CrossRef]
  53. Kawagoe, F.; Yasuda, K.; Mototani, S.; Sugiyama, T.; Uesugi, M.; Sakaki, T.; Kittaka, A. Synthesis and CYP24A1-dependent metabolism of 23-fluorinated vitamin D3 analogues. ACS Omega 2019, 4, 11332–11337. [Google Scholar] [CrossRef] [Green Version]
  54. Yamada, S.; Ohmori, M.; Takayama, H. Synthesis of 24,24-difluoro-25-hydroxyvitamin D3. Tetrahedron Lett. 1979, 20, 1859–1862. [Google Scholar] [CrossRef]
  55. Kobayashi, Y.; Taguchi, T.; Terada, T.; Oshida, J.-I.; Morisaki, M.; Ikekawa, N. Synthesis of 24,24-difluoro- and 24ξ-fluoro-25-hydroxyvitamin D3. Tetrahedron Lett. 1979, 20, 2023–2026. [Google Scholar] [CrossRef]
  56. Tanaka, Y.; DeLuca, H.F.; Schnoes, H.K.; Ikekawa, N.; Kobayashi, Y. 24,24-difluoro-1,25-dihydroxyvitamin D3: In vitro production, isolation, and biological activity. Arch. Biochem. Biophys. 1980, 199, 473–478. [Google Scholar] [CrossRef]
  57. Gill, H.S.; Londowski, J.M.; Corradino, R.A.; Zinsmeister, A.R.; Kumar, R. Synthesis and biological activity of novel vitamin D analogues: 24,24-difluoro-25-hydroxy-26,27-dimethylvitamin D3 and 24,24-difluoro-1α,25-dihydroxy-26,27-dimethylvitamin D3. J. Med. Chem 1990, 33, 480–490. [Google Scholar] [CrossRef]
  58. Konno, K.; Ojima, K.; Hayashi, T.; Takayama, H. An alternative and efficient synthesis of 24,24-difluoro-1α,25-dihydroxyvitamin D3. Chem. Pharm. Bull. 1992, 40, 1120–1124. [Google Scholar] [CrossRef] [Green Version]
  59. Ando, K.; Kondo, F.; Koike, F.; Takayama, H. An improved synthesis of 24,24-difluoro-1α,25-dihydroxyvitamin D3 from vitamin D2. Chem. Pharm. Bull. 1992, 40, 1662–1664. [Google Scholar] [CrossRef] [Green Version]
  60. Kondo, F.; Maki, S.; Konno, K.; Takayama, H. The first synthesis of 24,24-difluoro-1α-hydroxyvitamin D3 by means of radical deoxygenation of alcohols. Chem. Pharm. Bull. 1996, 44, 62–66. [Google Scholar] [CrossRef] [Green Version]
  61. Iwasaki, H.; Hosotani, R.; Miyamoto, Y.; Nakano, Y.; Yamamoto, K.; Yamada, S.; Shinki, T.; Suda, T.; Yamaguchi, K.; Konno, K.; et al. Stereoselective synthesis and structural establishment of (25S)-24,24-difluoro-1α,25,26-trihydroxyvitamin D3, a major metabolite of 24,24-difluoro-1α,25-dihydroxyvitamin D3. Tetrahedron 1998, 54, 14705–14724. [Google Scholar] [CrossRef]
  62. Flores, A.; Massarelli, I.; Thoden, J.B.; Plum, L.A.; DeLuca, H.F. A methylene group on C-2 of 24,24-difluoro-19-nor-1α,25-dihydroxyvitamin D3 markedly increases bone calcium mobilization in nivo. J. Med. Chem. 2015, 58, 9731–9741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kawagoe, F.; Mototani, S.; Yasuda, K.; Nagasawa, K.; Uesugi, M.; Sakaki, T.; Kittaka, A. Introduction of fluorine atoms to vitamin D3 side-chain and synthesis of 24,24-difluoro-25-hydroxyvitamin D3. J. Steroid Biochem. Mol. Biol. 2019, 195, 105477. [Google Scholar] [CrossRef]
  64. Kawagoe, F.; Mendoza, A.; Hayata, Y.; Asano, L.; Kotake, K.; Mototani, S.; Kawamura, S.; Kurosaki, S.; Akagi, Y.; Takemoto, Y.; et al. Discovery of a vitamin D receptor-silent vitamin D derivative that impairs sterol regulatory element-binding protein in vivo. J. Med. Chem. 2021, 64, 5689–5709. [Google Scholar] [CrossRef]
  65. Shiuey, S.-J.; Partridge, J.J.; Chadha, N.K.; Boris, A.; Uskoković, M.R. Stereospecific synthesis of 1α,25-dihydroxy-24R-fluorocholecalciferol (Ro23-0233). In Vitamin D, Chemical, Biochemical and Clinical Update; Walter De Gruyter: Berlin, Germany, 1985; pp. 765–766. [Google Scholar]
  66. Shiuey, S.-J.; Partridge, J.J.; Uskoković, M.R. Triply convergent synthesis of 1α,25-dihydroxy-24(R)-fluorocholecalciferol. J. Org. Chem. 1988, 53, 1040–1046. [Google Scholar] [CrossRef]
  67. Onisko, B.L.; Schnoes, H.K.; DeLuca, H.F. Synthesis of potential vitamin D antagonists. Tetrahedron Lett. 1977, 18, 1107–1108. [Google Scholar] [CrossRef]
  68. Napoli, J.L.; Fivizzani, M.A.; Hamstra, A.H.; Schnoes, H.K.; DeLuca, H.F.; Stern, P.H. The synthesis and activity in vitro of 25-masked 1α-hydroxylated vitamin D3 analogs. Steroids 1978, 32, 453–466. [Google Scholar] [CrossRef]
  69. Leyssens, C.; Verlinden, L.; Verstuyf, A. The future of vitamin D analogs. Front. Physiol. 2014, 5, 122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Maestro, M.A.; Molnár, F.; Carlberg, C. Vitamin D and its synthetic analogs. J. Med. Chem. 2019, 62, 6854–6875. [Google Scholar] [CrossRef]
  71. Tanaka, Y.; DeLuca, H.F.; Kobayashi, Y.; Ikekawa, N. 26,26,26,27,27,27-Hexafluoro-1,25-dihydroxyvitamin D3: A highly potent, long-lasting analog of 1,25-dihydroxyvitamin D3. Arch. Biochem. Biophys. 1984, 229, 348–354. [Google Scholar] [CrossRef]
  72. Harada, M.; Miyahara, T.; Miyata, M.; Tomita, I.; Okayachi, H.; Ikemoto, Y.; Higuchi, S.; Otomo, S.; Kozuka, H.; Ikekawa, N. Effects on cultured neonatal mouse calvaria of 1α,25-dihydroxyvitamin D3, 26,26,26,27,27,27-hexafluoro-1α,25-dihydroxyvitamin D3 and 26,26,26,27,27,27-hexafluoro-1α,23S,25-trihydroxyvitamin D3. Bone Miner. 1992, 18, 41–49. [Google Scholar] [CrossRef]
  73. Inaba, M.; Okuno, S.; Nishizawa, Y.; Imanishi, Y.; Katsumata, T.; Sugata, I.; Morii, H. Effect of substituting fluorine for hydrogen at C-26 and C-27 on the side chain of 1,25-dihydroxyvitamin D3. Biochem. Pharmacol. 1993, 45, 2331–2336. [Google Scholar] [CrossRef]
  74. Sakaki, T.; Sawada, N.; Abe, D.; Komai, K.; Shiozawa, S.; Nonaka, Y.; Nakagawa, K.; Okano, T.; Ohta, M.; Inouye, K. Metabolism of 26,26,26,27,27,27-F6-1α,25-dihydroxyvitamin D3 by CYP24: Species-based difference between humans and rats. Biochem. Pharmacol. 2003, 65, 1957–1965. [Google Scholar] [CrossRef]
  75. Kasai, N.; Sakaki, T.; Shinkyo, R.; Ikushiro, S.; Iyanagi, T.; Ohta, M.; Inouye, K. Metabolism of 26,26,26,27,27,27-F6-1α,23S,25-trihydroxyvitamin D3 by human UDP-glucuronosyltransferase 1A3*. Drug Metab. Dispos. 2005, 33, 102–107. [Google Scholar] [CrossRef]
  76. Kobayashi, Y.; Taguchi, T.; Kanuma, N. Synthesis of 26,26,26,27,27,27-hexafluoro-25-hydroxyvitamin D3. J. Chem. Soc. Chem. Commun. 1980, 459–460. [Google Scholar] [CrossRef]
  77. Kobayashi, Y.; Taguchi, T.; Mitsuhashi, S.; Eguchi, T.; Ohshima, E.; Ikekawa, N. Studies on organic fluorine compounds. XXXIX. Studies on steroids. LXXIX. Synthesis of 1α,25-dihydroxy-26,26,26,27,27,27-hexafluorovitamin D3. Chem. Pharm. Bull. 1982, 30, 4297–4304. [Google Scholar] [CrossRef] [Green Version]
  78. Ohira, Y.; Taguchi, T.; Iseki, K.; Kobayashi, Y. Preparation of (22S)- and (22R)-24-homo-26, 26, 26, 27, 27, 27-hexafluoro-1,22,25-trihydroxy-24-yne-vitamin D3. Chem. Pharm. Bull. 1992, 40, 1647–1649. [Google Scholar] [CrossRef] [Green Version]
  79. Wu, Y.; Sabbe, K.; De Clercq, P.; Vandewalle, M.; Bouillon, R.; Verstuyf, A. Vitamin D3: Synthesis of seco C-9,11,21-trisnor-17-methyl-1α,25-dihydroxyvitamin D3 analogues. Bioorg. Med. Chem. Lett. 2002, 12, 1629–1632. [Google Scholar] [CrossRef]
  80. Sigüeiro, R.; Maestro, M.A.; Mouriño, A. Synthesis of side-chain locked analogs of 1α,25-dihydroxyvitamin D3 bearing a C17 methyl group. Org. Lett. 2018, 20, 2641–2644. [Google Scholar] [CrossRef]
  81. Kawagoe, F.; Sugiyama, T.; Uesugi, M.; Kittaka, A. Recent developments for introducing a hexafluoroisopropanol unit into the vitamin D side chain. J. Steroid Biochem. Mol. Biol. 2018, 177, 250–254. [Google Scholar] [CrossRef]
  82. Asano, L.; Watanabe, M.; Ryoden, Y.; Usuda, K.; Yamaguchi, T.; Khambu, B.; Takashima, M.; Sato, S.; Sakai, J.; Nagasawa, K.; et al. Vitamin D metabolite, 25-hydroxyvitamin D, regulates lipid metabolism by inducing degradation of SREBP/SCAP. Cell Chem. Biol. 2017, 24, 207–217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Deactivation metabolic pathways of 25(OH)D3 (2) and 1α,25(OH)2D3 (1).
Scheme 1. Deactivation metabolic pathways of 25(OH)D3 (2) and 1α,25(OH)2D3 (1).
Ijms 22 08191 sch001
Scheme 2. DeLuca’s direct fluorination approach to 1-fluorovitamin D3 (6).
Scheme 2. DeLuca’s direct fluorination approach to 1-fluorovitamin D3 (6).
Ijms 22 08191 sch002
Scheme 3. Direct C1 and C25 fluorination approach to 1α,25-difluorovitamin D3 (7).
Scheme 3. Direct C1 and C25 fluorination approach to 1α,25-difluorovitamin D3 (7).
Ijms 22 08191 sch003
Scheme 4. Synthesis of 1α-fluoro-VD3 (10) and 1α-fluoro-25(OH)D3 (11) from sterols.
Scheme 4. Synthesis of 1α-fluoro-VD3 (10) and 1α-fluoro-25(OH)D3 (11) from sterols.
Ijms 22 08191 sch004
Scheme 5. Uskoković’s approach to 1α-fluoro-25(OH)D3 (11) using the Wittig–Horner reaction.
Scheme 5. Uskoković’s approach to 1α-fluoro-25(OH)D3 (11) using the Wittig–Horner reaction.
Ijms 22 08191 sch005
Scheme 6. Alternative convergent synthetic approach to 1α-fluoro-25(OH)D3 (11) using the A-ring moiety (12), available from the A-ring of VD3.
Scheme 6. Alternative convergent synthetic approach to 1α-fluoro-25(OH)D3 (11) using the A-ring moiety (12), available from the A-ring of VD3.
Ijms 22 08191 sch006
Scheme 7. Various 1α-fluoro-VD3 analogues with six different double side chains were synthesized using the Wittig–Horner olefination at the coupling stage with the 1α-fluoro-A-ring part (12).
Scheme 7. Various 1α-fluoro-VD3 analogues with six different double side chains were synthesized using the Wittig–Horner olefination at the coupling stage with the 1α-fluoro-A-ring part (12).
Ijms 22 08191 sch007
Scheme 8. Synthesis of 1- and 3-fluoro-19-norvitamin D3 analogues (2225) using direct deoxyfluorination reactions.
Scheme 8. Synthesis of 1- and 3-fluoro-19-norvitamin D3 analogues (2225) using direct deoxyfluorination reactions.
Ijms 22 08191 sch008
Scheme 9. Ikekawa’s approach to 2β-fluoro-1α(OH)D3 (26) via 2β-fluorination of A-ring epoxide using a nucleophilic fluorination reaction.
Scheme 9. Ikekawa’s approach to 2β-fluoro-1α(OH)D3 (26) via 2β-fluorination of A-ring epoxide using a nucleophilic fluorination reaction.
Ijms 22 08191 sch009
Scheme 10. Introduction of the 2α-fluoro group to VD3 using an electrophilic fluorination reaction.
Scheme 10. Introduction of the 2α-fluoro group to VD3 using an electrophilic fluorination reaction.
Ijms 22 08191 sch010
Scheme 11. Synthesis of 2β-fluoro-1α,25-(OH)2D3 (30).
Scheme 11. Synthesis of 2β-fluoro-1α,25-(OH)2D3 (30).
Ijms 22 08191 sch011
Scheme 12. Mikami’s group synthesized 2α-fluoro-19-normaxacalcitol (31) using a convergent coupling method between the 2α-fluoro-A-ring part, prepared from regio- and stereoselective fluorination to 32, followed by asymmetric catalytic carbonyl-ene cyclization, and the appropriate 8-keto-CD-ring.
Scheme 12. Mikami’s group synthesized 2α-fluoro-19-normaxacalcitol (31) using a convergent coupling method between the 2α-fluoro-A-ring part, prepared from regio- and stereoselective fluorination to 32, followed by asymmetric catalytic carbonyl-ene cyclization, and the appropriate 8-keto-CD-ring.
Ijms 22 08191 sch012
Scheme 13. Posner’s synthetic route to 2,2-difluoro-1α,25-dihydroxyvitamin D3 (34) from 2,2-difluoro-A-ring (33) via an inverse-electron-demand Diels–Alder reaction between a pyrone diene and difluorovinyl ether.
Scheme 13. Posner’s synthetic route to 2,2-difluoro-1α,25-dihydroxyvitamin D3 (34) from 2,2-difluoro-A-ring (33) via an inverse-electron-demand Diels–Alder reaction between a pyrone diene and difluorovinyl ether.
Ijms 22 08191 sch013
Scheme 14. Segal’s and Mazur’s 3β-fluoro-3-deoxyvitamin D3 syntheses.
Scheme 14. Segal’s and Mazur’s 3β-fluoro-3-deoxyvitamin D3 syntheses.
Ijms 22 08191 sch014
Scheme 15. Synthesis of 3β-fluoro-3-deoxyvitamin D3 (35) starting from 7-dehydrocholesterol (39).
Scheme 15. Synthesis of 3β-fluoro-3-deoxyvitamin D3 (35) starting from 7-dehydrocholesterol (39).
Ijms 22 08191 sch015
Scheme 16. Yamada’s approach to 4,4-difluorovitamin D3 (42,43) using a thermodynamic fluorination reaction.
Scheme 16. Yamada’s approach to 4,4-difluorovitamin D3 (42,43) using a thermodynamic fluorination reaction.
Ijms 22 08191 sch016
Scheme 17. Mazur’s trial to synthesize 19,19-difluorovitamin D3 (46).
Scheme 17. Mazur’s trial to synthesize 19,19-difluorovitamin D3 (46).
Ijms 22 08191 sch017
Scheme 18. Yamada’s approach to (10E)- and (10Z)-19-fluorovitamin D3 (50,51) in (A) and (10E)- and (10Z)-19-fluoro-1α,25(OH)2D3 (52,53) analogues in (A,B).
Scheme 18. Yamada’s approach to (10E)- and (10Z)-19-fluorovitamin D3 (50,51) in (A) and (10E)- and (10Z)-19-fluoro-1α,25(OH)2D3 (52,53) analogues in (A,B).
Ijms 22 08191 sch018
Scheme 19. Synthesis of 6-fluorovitamin D3 (54) using piperidinosulfur trifluoride as a fluorination reagent.
Scheme 19. Synthesis of 6-fluorovitamin D3 (54) using piperidinosulfur trifluoride as a fluorination reagent.
Ijms 22 08191 sch019
Scheme 20. Efficient introduction of the C6- and C7-fluorovinyl unit to the A-ring or CD-ring followed by construction of the diene structures using a Ni-catalyzed cross-coupling reaction.
Scheme 20. Efficient introduction of the C6- and C7-fluorovinyl unit to the A-ring or CD-ring followed by construction of the diene structures using a Ni-catalyzed cross-coupling reaction.
Ijms 22 08191 sch020
Scheme 21. Stereoselective hydroxylation, coupling with the A-ring precursor, and subsequent fluorination at the C11 position.
Scheme 21. Stereoselective hydroxylation, coupling with the A-ring precursor, and subsequent fluorination at the C11 position.
Ijms 22 08191 sch021
Scheme 22. Synthesis of 22-fluorovitamin D3 (67) from (22S)-22-hydroxycholesterol (68).
Scheme 22. Synthesis of 22-fluorovitamin D3 (67) from (22S)-22-hydroxycholesterol (68).
Ijms 22 08191 sch022
Scheme 23. Synthesis of 23,23-difluoro-25(OH)D3 (70) using α-ketoester (71) as a key intermediate, and the structure of 23,23-difluoro-1α,25(OH)2D3 (72) is also shown.
Scheme 23. Synthesis of 23,23-difluoro-25(OH)D3 (70) using α-ketoester (71) as a key intermediate, and the structure of 23,23-difluoro-1α,25(OH)2D3 (72) is also shown.
Ijms 22 08191 sch023
Scheme 24. Synthesis of 23,26,26,26,27,27,27-heptafluoro-1α,25(OH)2D3 using DAST for the seventh fluorine introduction.
Scheme 24. Synthesis of 23,26,26,26,27,27,27-heptafluoro-1α,25(OH)2D3 using DAST for the seventh fluorine introduction.
Ijms 22 08191 sch024
Scheme 25. Stereoselective C23-fluorination of 23-hydroxy-CD-rings using PyFluor.
Scheme 25. Stereoselective C23-fluorination of 23-hydroxy-CD-rings using PyFluor.
Ijms 22 08191 sch025
Scheme 26. Takayama’s and Kobayashi-Ikekawa’s synthetic routes to 24,24-difluoro-25(OH)D3 (80).
Scheme 26. Takayama’s and Kobayashi-Ikekawa’s synthetic routes to 24,24-difluoro-25(OH)D3 (80).
Ijms 22 08191 sch026
Scheme 27. Kumar’s synthetic approach to 24,24-difluorovitamin D homologues (83,84) using the Reformatsky reaction.
Scheme 27. Kumar’s synthetic approach to 24,24-difluorovitamin D homologues (83,84) using the Reformatsky reaction.
Ijms 22 08191 sch027
Scheme 28. Takayama’s two improved synthetic approaches to 24,24-difluoro-1α,25(OH)2D3 (82).
Scheme 28. Takayama’s two improved synthetic approaches to 24,24-difluoro-1α,25(OH)2D3 (82).
Ijms 22 08191 sch028
Scheme 29. Synthetic routes to 24,24-difluoro-1α(OH)D3 (85) (A) and its CYP24-metabolite (87) (B) by the Takayama group.
Scheme 29. Synthetic routes to 24,24-difluoro-1α(OH)D3 (85) (A) and its CYP24-metabolite (87) (B) by the Takayama group.
Ijms 22 08191 sch029
Scheme 30. DeLuca’s convergent approach to 24,24-difluoro-1α,25(OH)2-19-norVD3 (88,89) and the 2-exomethylene analogues (90,91) including their 20-epi versions using the Wittig–Horner reaction.
Scheme 30. DeLuca’s convergent approach to 24,24-difluoro-1α,25(OH)2-19-norVD3 (88,89) and the 2-exomethylene analogues (90,91) including their 20-epi versions using the Wittig–Horner reaction.
Ijms 22 08191 sch030
Scheme 31. Efficient construction of the 24,24-difluorinated CD-ring unit (94) and the subsequent A-ring coupling reaction (A). Synthesis of the novel SREBP-specific inhibitor KK-052 (B).
Scheme 31. Efficient construction of the 24,24-difluorinated CD-ring unit (94) and the subsequent A-ring coupling reaction (A). Synthesis of the novel SREBP-specific inhibitor KK-052 (B).
Ijms 22 08191 sch031
Scheme 32. Synthesis of 24-monofluoro-25(OH)D3 (95) using nucleophilic fluorination as a key step.
Scheme 32. Synthesis of 24-monofluoro-25(OH)D3 (95) using nucleophilic fluorination as a key step.
Ijms 22 08191 sch032
Scheme 33. (24R)-Fluoro-1α,25(OH)2D3 (96) was synthesized by Uskoković et al. through a linear synthetic route using 97 and optically active L-(‒)-malic acid as a chiral synthon (A) and an alternative convergent route starting from a CD-ring part, 98 (B).
Scheme 33. (24R)-Fluoro-1α,25(OH)2D3 (96) was synthesized by Uskoković et al. through a linear synthetic route using 97 and optically active L-(‒)-malic acid as a chiral synthon (A) and an alternative convergent route starting from a CD-ring part, 98 (B).
Ijms 22 08191 sch033
Scheme 34. Synthesis of 25-fluoro-VD3 analogues via direct C25 fluorination from 25(OH)D3 (A) and 1α,25(OH)2D3 (B) using DAST as a fluorinating reagent.
Scheme 34. Synthesis of 25-fluoro-VD3 analogues via direct C25 fluorination from 25(OH)D3 (A) and 1α,25(OH)2D3 (B) using DAST as a fluorinating reagent.
Ijms 22 08191 sch034
Figure 1. Structures of falecalcitriol (101) and its major metabolite (23S)-23-hydroxyfalecalcitriol (102).
Figure 1. Structures of falecalcitriol (101) and its major metabolite (23S)-23-hydroxyfalecalcitriol (102).
Ijms 22 08191 g001
Scheme 35. Synthesis of 26,26,26,27,27,27-hexafluoro-25(OH)D3 (103) and 26,26,26,27,27,27-hexafluoro-1α,25(OH)2D3 (101) using hexafluoroacetone as a fluorine source.
Scheme 35. Synthesis of 26,26,26,27,27,27-hexafluoro-25(OH)D3 (103) and 26,26,26,27,27,27-hexafluoro-1α,25(OH)2D3 (101) using hexafluoroacetone as a fluorine source.
Ijms 22 08191 sch035
Scheme 36. Combination of the hexafluoropropanol unit derived from hexafluoroacetone (HFA) and triple bond(s), which brought conformational rigidity to the side chain.
Scheme 36. Combination of the hexafluoropropanol unit derived from hexafluoroacetone (HFA) and triple bond(s), which brought conformational rigidity to the side chain.
Ijms 22 08191 sch036
Scheme 37. Synthesis of 26,26,26,27,27,27-hexafluoro-25(OH)D3 (103) and synthetically useful 26,26,26,27,27,27-hexafluoro-CD-ring precursor (107) using efficient two-step trifluoromethylation method.
Scheme 37. Synthesis of 26,26,26,27,27,27-hexafluoro-25(OH)D3 (103) and synthetically useful 26,26,26,27,27,27-hexafluoro-CD-ring precursor (107) using efficient two-step trifluoromethylation method.
Ijms 22 08191 sch037
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kawagoe, F.; Mototani, S.; Kittaka, A. Design and Synthesis of Fluoro Analogues of Vitamin D. Int. J. Mol. Sci. 2021, 22, 8191. https://doi.org/10.3390/ijms22158191

AMA Style

Kawagoe F, Mototani S, Kittaka A. Design and Synthesis of Fluoro Analogues of Vitamin D. International Journal of Molecular Sciences. 2021; 22(15):8191. https://doi.org/10.3390/ijms22158191

Chicago/Turabian Style

Kawagoe, Fumihiro, Sayuri Mototani, and Atsushi Kittaka. 2021. "Design and Synthesis of Fluoro Analogues of Vitamin D" International Journal of Molecular Sciences 22, no. 15: 8191. https://doi.org/10.3390/ijms22158191

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