Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Feb 1.
Published in final edited form as: J Fluor Chem. 2018 Dec 11;218:90–98. doi: 10.1016/j.jfluchem.2018.12.005

Synthesis of Leu-Enkephalin Peptidomimetics Containing Trifluoromethylalkenes as Amide Isopolar Mimics

Venkateswararao Eeda 1, Manikandan Selvaraju 1, Ryan A Altman 1
PMCID: PMC6499489  NIHMSID: NIHMS1516969  PMID: 31061541

Abstract

Fluorinated peptidomimetics are valuable substrates for exploring peptide backbone conformations and for perturbing physicochemical properties of probe compounds. However, in some cases synthetic limitations restrict installation of the fluorinated peptidomimetics into the desired probe compounds. For instance, trifluoromethylalkenes have served as amide isopolar mimics, but are rarely utilized, because many standard peptide-coupling conditions promote the isomerization of the alkene to thermodynamically favored positions. To address this challenge, we report the conversion of a naturally occurring amino acid to a Tyr1-ψ/[CF3C=CH]-Gly2 dipeptide mimetic, and notably, successful peptide coupling reactions that avoid alkene isomerization. Using this strategy, we generated trifluoromethylalkene-containing Leu-enkephalin peptidomimetics in high purity and good yield. This sequence suggests that the trifluoromethylalkene peptidomimetics can be incorporated into other target molecules with appropriate optimization.

Keywords: Leu-Enkephalin, Trifluoromethylalkene, Peptidomimetics, Peptide coupling, Trifluoromethylation

Graphical Abstract

graphic file with name nihms-1516969-f0001.jpg

1. Introduction

Peptides are key messengers in biological signaling pathways and are therapeutically useful for modulating several pathological conditions [1]. Unfortunately, many peptide-based drug candidates have struggled clinically, because of poor physicochemical properties and in vivo stability, which restrict systemic distribution and half-lives [2-3]. These drawbacks can be primarily attributed to the peptide backbones that make strong H-bonding interactions with H2O, thus disfavoring desolvation, and that are subject to rapid enzymatic hydrolysis [4]. To improve the drug-like properties of peptide-based probes, the labile amide bonds can be replaced with stable amide isosteric- or isopolar-mimics. One such strategy involves fluorinated substructures that can possess similar dipole moments, electrostatic potentials, and steric-, resonance-, and inductive-properties to amides [5], but are not substrates for endogenous peptidases [6-9]

Amongst fluorinated peptidomimetics, the trifluoromethylalkene group (CF3-alkene) can serve as an isopolar replacement for the amide linkage and can be used to probe chemical biology [10-11]. This substructure closely mimics the electrostatic potential and direction of polarization of the carbonyl group, as well as the sp2-hybridization of the carbonyl carbon (Figure 1A) [12]. However, incorporation of the CF3-alkene group would impart greater thermal, proteolytic and conformational stability, and greater lipophilicity to a polypeptide [10, 13]. This substructure has been exploited in few probe molecules, including (a) a trifluoromethylalkene analogue of Gramicidin S that showed similar solution-state and solid-state conformations to the natural peptide [12], and (b) an X-ray crystallographic study of the CF3-alkene analogue of the L-Ala-D-Ala dipeptide, in which the CF3-alkene analogue adopted a considerably more organized structure than other peptidomimetics (Figure 1A) [5].

Fig. 1.

Fig. 1.

Trifluoromethylalkene dipeptides for probing chemical biology

Beyond these intriguing examples, exploitation of CF3-alkene peptidomimetic biological probes remains limited, presumably because of the potential for the alkene to isomerize from the β,γ-position to the thermodynamically favorable α,β-position of the amide during peptide coupling (Figure 1B). For the two successful examples, the bulky α-substituents of Gramicidin S and L-Ala-D-Ala CF3-alkene may disfavor the base-mediated isomerization of the double bond, which might enable coupling (Figure 1A, blue highlight).

To investigate synthetic routes for incorporating the CF3-alkene into a peptide backbone, while minimizing alkene isomerization, we selected Leu-enkephalin (1, Figure 1C) as a model substrate. Leu-enkephalin is an endogenous pentapeptide that exhibits moderate selectivity for the δ–opioid receptor (DOR), and several preclinical studies have demonstrated that selective activation of the DOR induces analgesia without promoting the undesirable side effects (tolerance, dependence, and constipation) typically associated with stimulation of the μ–opioid receptor (MOR) [14]. However, as with many peptides, the poor proteolytic stability and bioavailability of Leu-enkephalin limit in vivo therapeutic utility [15-17]. Previous studies of Leu-enkephalin identified the Tyr1–Gly2 amide bond as both metabolically labile and amenable to modification with minimal loss of opioid activity (Figure 1C) [18]. Therefore, we elected to replace the Tyr1–Gly2 peptide bond of Leu-enkephalin with a CF3-alkene, which might improve the pharmacokinetic and physicochemical properties of Leu-enkephalin. As part of our ongoing effort to probe the therapeutic potential of fluorinated peptidomimetics, we herein describe the synthesis of Tyr1-ψ[CF3C=CH]-Gly2 peptidomimetic analogs of Leu-enkephalin (2a-b) [19-20]. These syntheses overcame the isomerization of the alkene that has likely limited the use of CF3-alkenes as peptidomimetics.

2. Results and Discussion

Our retrosynthetic analysis of enkephalin-mimic 2 involved an initial peptide bond-disconnection to provide trifluoromethylalkene 10 and known tripeptide Gly-Phe-Leu-OR 11 (Scheme 1) [21]. Trifluoromethylalkene 10 could be constructed via a Wittig olefination reaction between trifluoromethylketone 6 and phosphonium ylide 7. Trifluoromethylketone 6 would be accessed through a series of functional group transformations from commercially available L-tyrosine.

Scheme. 1.

Scheme. 1.

Peptide coupling and Wittig olefination strategy to access Tyr1-ψ[CF3C=CH]-Gly2-Leu-enkephalin peptidomimetic 2.

Initial unsuccessful attempts to prepare key building block, trifluoromethylketone 6 involved reactions of various tyrosine derivatives under Dakin–West reaction conditions using trifluoroacetic anhydride and pyridine (Scheme 2A) [22]. However, these reactions generally decomposed the substrate and produced complex mixtures of products. Next, using Kolb’s modification of the Dakin–West reaction [23] we accessed only a hydrate form (3a) of the CF3-ketone intermediate which under various conditions, failed to dehydrate to the corresponding ketone, and instead cyclized to produce oxazole derivatives (Scheme 2B) [24-25]. Subsequent attempts to couple CF3-keto hydrate 3a via Wittig reaction with ylide 7, generally decomposed the substrate. Next, we explored a strategy involving the Ruppert–Prakash reagent (TMSCF3), which can convert esters [26] or Weinreb amides [27] into trifluoromethyl ketones. However, these reactions generated inseparable undesired products (Scheme 2C), presumably by 1,2–addition of CF3 to the carbonyl of phthalimide. Over the course of this preliminary work, we realized that the above–mentioned general strategies would not afford the desired trifluoromethyl ketone.

Scheme 2.

Scheme 2.

Attempted strategies to synthesize Tyr-derived trifluoromethylketone 6

Our revised synthesis of CF3 ketone 6 involved nucleophilic addition of TMSCF3 and subsequent oxidation as key steps. In practice, amino alcohol 4 (prepared according to reference 28, but also available commercially) was protected as a phthalimide, and subsequent Swern oxidation afforded tyrosine–derived aldehyde 5 (Scheme 2D) [28]. Next, we investigated the nucleophilic addition of TMSCF3 to aldehyde 5 under various conditions. Reactions of 5 with TMSCF3 in the presence of catalytic amount of TBAF at room temperature gave an inseparable mixture of products resulting from the competitive 1,2-addition of CF3 to the aldehyde and phthalimide carbonyl groups. However, lowering the reaction temperature to −78 °C, and slowing the addition of TBAF improved the 1,2-addition of CF3 to the aldehyde, and afforded a diastereomeric mixture of alcohols, which was oxidized using Dess-Martin periodinane (DMP) to provide CF3-ketone 6. However, compound 6 was obtained as a hydrate form that converted to CF3-ketone 6 in 63% yield by refluxing for 2 h in dry toluene with molecular sieves (4 Å).

Subsequent effort enabled the conversion of CF3-ketone 6 to the Tyr1-ψ[CF3C=CH]-Gly2 dipeptide mimic by a sequence involving olefination, deprotection, and subsequent oxidation. First, CF3 ketone 6 and triphenylphosphonium bromide-derived ylide 7 (prepared from the corresponding 1,3-propanediol) [29] were coupled using a Wittig olefination reaction (Scheme 3). Initial attempts to optimize the reaction to obtain high Z-selectivity failed over a range of temperatures (0 → −78°C), with most of the conditions producing a 1:1 mixture of E/Z isomers (8a–b). Optimal yields were obtained by adding a THF solution of the CF3-ketone to a solution of the in-situ generated ylide at −78 °C, followed by an aqueous workup. Following this procedure, both benzyl (8a) and THP (8b) protected alkene derivatives were obtained in good yields, which were subsequently deprotected with BCl3 or tetrabutylammonium tribromide (TBATB) respectively, to generate a 1:1 E:Z mixture of isomeric alcohols (9a–b), which were separated by SiO2 gel chromatography.

Scheme 3.

Scheme 3.

Wittig olefination strategy for preparing protected Tyr1-ψ[CF3C=CH]-Gly2 10

Next, we investigated the oxidation of allylic alcohols 9 (both E and Z), bearing the sensitive CF3–alkene, to provide acid 10. Oxidation of 9a (Scheme 3) with either the Dess-Martin periodinane or using Swern oxidation conditions generally decomposed the substrate, potentially via oxidation of the free phenol. In contrast, treatment of 9b (R3 = benzyl) with Dess-Martin periodinane, followed by Pinnick oxidation provided an inseparable mixture of intermediate aldehyde and other undetermined side products. Subsequently, we explored a facile one-step oxidation of primary alcohol 9a to 10a using a CrO3-catalyzed oxidation using H5IO6 in wet MeCN at 0 °C. In this reaction, the aldehyde intermediate may have been sufficiently short-lived to minimize decomposition [30]. Additionally, the mild acidic conditions presumably minimized isomerization of the trifluoromethylalkene, and other decomposition pathways. Using these conditions, 9a converted to the corresponding acid 10a in 30–33% yield while also leaving unreacted alcohol (~65%). Considering that free phenol might contribute to the low yield of the reaction, we subjected benzyl-protected E-9b and Z-9b to the same oxidation conditions, which provided improved yields of carboxylates E-10b and Z-10b.

Having accomplished the successful synthesis of Tyr-ψ[CF3C=CH]-Gly dipeptide mimics 10, the next task involved the peptide coupling reaction with the tripeptide (11) to form the desired trifluoromethylalkene containing Leu-enkephalin mimetic. For this step, mild conditions would be required to inhibit the base catalyzed isomerization of the CF3-alkene to the thermodynamically favored α,β-unsaturated position. Hence, we optimized the reaction conditions to achieve the desired coupling product (Table 1). Using standard 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC) coupling conditions, the attempted coupling of Z-10a and 11a generated a complex mixture of unreacted acid Z-10a, along with isomerized product 13 and isomerized acid 14 (Table 1, entries 1–3 & 6). As the moderate acidity of the free phenol of Z-9a might catalyze the isomerization of the CF3-alkene, we next evaluated the coupling reaction of benzyl protected analog Z–10b. Attempts to couple Z–10b with tripeptide 11a using a variety of BOP [(benzotriazol-1-yloxy)tris(dimethylamino)phosphonium hexafluorophosphate]-derived coupling reagents (entries 4, 5, 7 & 8) provided similar results. Further optimization using dual functional reagents such as 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methylmorpholinium chloride (DMT-MM, entry 9) [31] and 3-(diethoxyphosphoryloxy)-1,2,3-benzotriazin-4(3H)-one (DEPBT, entry 10) [32], that act as both in-situ acid activator and coupling reagents, was not fruitful.

Table 1.

Coupling of Z–Tyr-ψ[CF3C=CH]-Gly dipeptide mimics 10a–b with Gly-Phe-Leu tripeptide a

graphic file with name nihms-1516969-t0007.jpg

entry Z10a (or)
Z10b
coupling
reagent
(1 equiv.)
base
(2 equiv.)
additive
(1 equiv.)
time
(h)
solvent 19F NMR Yield(%)
Z-10 Z-12 13 14
1 Z-10a EDC DIPEA HOBt 12 THF 90
2 Z-10a EDC DMAP HOBt 12 DCM 11 16 71
3 Z-10a EDC DMAP HOBt 20 CHCl3 29 2 6 61
4 Z-10a PyBOP DIPEA 12 CHCl3 22 11 51 17
5 Z-10a PyBOP NMM HOBt 12 DMF 14 6 78
6 Z-10b EDC DIPEA HOBt 12 THF 50 50
7 Z-10b PyBOP DIPEA HOBt 12 THF 19 80
8 Z-10b BOP NMM 12 DMF Undefined mixture
9 Z-10b DMT-MM 5 MeOH No Reaction
10 Z-10b DEPBT DIPEA 5 DCM No Reaction
11 Z-10b 12 DCM 7 10 21 60
12 Z-10b COMU DIPEA 0.25 DCM 30 69
13 Z-10b HBTU DIPEA 1 DCM 14 34 51
14 Z-10b HBTU DIPEA 12 THF 10 20 64
15 Z-10b HCTU DIPEA 1 DMF 10 80
16 Z-10b HCTU DIPEA 25 DCM 14 87
17 Z-10b HATU DIPEA 1 DCM 64 35
18 Z-10b HATU DIPEA 25 DCM 68 31
19b Z-10b SOCl2 TEA 3 DCM 78 11 9
20b Z-10b SOCl2 TEA 1 DCM 11 84 4
a

Reactions were performed with 0.039 mmol of 10, and 0.20 mL of solvent.

b

Reactions were performed with 0.039 mmol of 10, 0.20 mL of solvent, 3.0 equiv. of SOCl2, 3.0 equiv. base Conversion and yield data were determined by 19F NMR (α,α,α-trifluorotoluene as internal standard).

Considering the low conversion and high recovery of isomerized peptide 13 and isomerized acid 14, we investigated the cause of isomerization. As determined by 19F NMR and MS analysis, 67% of substrate Z–10b isomerized to 14 (12 h, rt, THF). However, subjection of the substrate to weak bases (e.g. imidazole) accelerated the isomerization, leaving no detectable substrate at 20 min. These results prompted us to explore less basic or base-free coupling conditions that might avoid the isomerization.

Further attempts to minimize isomerization included the use of base-free conditions and uronium salt-derived coupling reagents. Employment of a recent ynamide coupling reagent that avoids the use of base provided the desired product (Z–12) in low yield together with a mixture of isomerized peptide (13) and unreacted isomerized acid (14, entry 11) [33] though purification of this reaction mixture was not fruitful. Next, we explored various uronium salts as coupling reagents. The use of COMU [(1-cyano-2-ethoxy-2-oxoethylidenaminooxy)dimethylamino-morpholino-carbenium hexafluorophosphate] afforded 30% of the desired peptide Z–12 (entry 12). Encouraged by this result, we screened other uronium salts. The use of HBTU [(2-(1H-benzotriazol-1-yl)-1,1,3,3-tetramethyluronium hexafluorophosphate] in DCM and THF provided the desired peptide Z–12 in 34% (1 h) and 10% (12 h) respectively (entries 13–14). Surprisingly, the use of HCTU [2-(6-chloro-1-H-benzotriazole-1-yl)-1,1,3,3-tetramethylaminium hexafluorophosphate] in DMF generated 80% of isomerized acid 14 with no evidence of productive coupling (entry 15). Conversely, changing the solvent to DCM afforded 87% of isomerized coupled peptide 13 (entry 16). Considering these results, a general solvent–isomerization relationship emerged, more polar solvents promoting the isomerization of the CF3-alkene in Z–10, potentially because of the improved ability to stabilize the charged cation-anion pair (Figure 1B). Thus, we continued our optimization studies exclusively in DCM. In this solvent, HATU {1-[bis(dimethylamino)methylene]-1H-1,2,3-triazolo[4,5-b]pyridinium 3-oxid hexafluorophosphate} finally provided coupled CF3-alkene Z–12 as the major product (64% yield, 2:1 selectivity with 13, entry 17). Although this reaction provided ~60% of the desired product, chromatographic separation of 13 proved challenging, and a pure product was not obtained.

Subsequent efforts eventually provided optimized conditions for generating Z–12 that minimized formation of isomerized products 13 and 14. Specifically, treatment of acid Z–10b with excess thionyl chloride (SOCl2) in anhydrous DCM at ambient temperature, followed by addition to 11a at 0 °C readily afforded Z–12b in 78% yield (entry 20), but also produced 11% of the inseparable isomerized peptide 13. To improve the ratio of products, we reversed the addition strategy (adding a mixture of base and tripeptide to the solution of acid chloride) and shortened the reaction time (3 h 1 h) to provide Z–12b in 84% yield with only 4% of isomerized peptide 14 and 11% unreacted acid Z–10b (entry 21). With these conditions, the product mixture was separable by column chromatography (HPLC purity = 97.0%). Employing the same conditions, the other isomer E-10b was successfully coupled to afford the t-Bu ester variant of E–12b in 83% yield (HPLC purity = 99.5%; eq. 1).

graphic file with name nihms-1516969-f0006.jpg

Having obtained the CF3-alkene Leu-enkephalin peptidomimetic, we briefly explored the late-stage removal of the phthalimide, benzyl ether, and C-terminal alkyl ester groups. However, our attempted efforts either decomposed the substrate or isomerized the trifluoromethylalkene [34-36]. We speculate that alternate protecting strategies that avoid basic conditions might provide the desired analogs.

3. Conclusion

In conclusion, we successfully coupled phthalimide protected Tyr-ψ[CF3C=CH]-Gly dipeptide mimics to a C-terminal tripeptide to obtain a new protected Leu-enkephalin analog. The optimized mild conditions minimize isomerization of the double bond from the β,γ-position to thermodynamically favored α,β-position. The amide coupling reactions delivered products bearing both acid-labile and base-labile protecting groups, which provides multiple strategies for late stage deprotection of the final peptides. Most importantly, these successful coupling reactions suggest that, with appropriate optimization, other CF3-alkene peptidomimetics might be effectively coupled to access target analogs of a lead peptide. Ongoing work in our laboratory focuses on deprotecting both the C- and N-termini to provide analogs of Leu-enkephalin, and on characterizing pharmacodynamic and physicochemical perturbations imparted by the trifluoromethyl alkene isopolar mimic.

4. Experimental Section

4.1. General Methods

Unless otherwise noted, all reactions were performed using oven–dried glassware under an atmosphere of dry N2, which were sealed with rubber septa. Stainless steel syringes were used to transfer air or moisture–sensitive liquid reagents. Reactions were monitored by thin–layer chromatography (TLC) on UNIPLATE Silica Gel HLF 250 μm glass plates precoated with 230–400 mesh silica impregnated with a fluorescent indicator (250 nm), visualizing by quenching of fluorescence, KMnO4 solution, or p–anisaldehyde solution. A Teledyne Isco CombiFlash® RF–4x purification system was used for chromatographic purifications, silica gel was purchased from Sorbent Technologies (cat. #30930M–25, 60 Å, 40–63 μm). Commercial reagents were purchased and used as received. Anhydrous toluene (PhMe), acetonitrile (CH3CN), methanol (MeOH), dichloromethane (DCM), tetrahydrofuran (THF), and triethylamine (NEt3) were dispensed from a solvent purification system, in which the solvent was dried by passage through two columns of activated alumina under argon.

Proton nuclear magnetic resonance (1H NMR) spectra were recorded at 400 or 500 MHz. Carbon nuclear magnetic resonance (13C NMR) spectra were recorded at 101 or 126 MHz. Fluorine nuclear magnetic resonance (19F NMR) spectra were recorded at 376 MHz. Chemical shifts (δ) for protons are reported in parts per million (ppm) down–field from trimethylsilane and are referenced to the proton resonance of residual CHCl3 in the NMR solvent (δ = 7.27 ppm). Chemical shifts (δ) for carbon are reported in parts per million down–field from trimethylsilane and are referenced to the carbon resonances of the solvent peak (δ = 77.16 ppm). Chemical shifts (δ) for fluorine are reported in parts per million up-field from trichlorofluoromethane and are referenced to PhCF3 (δ = −63.72 ppm). NMR data are represented as follows: chemical shift (ppm), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, sept = septet, m = multiplet), coupling constants in hertz (Hz), and integration.

Exact mass determinations were obtained by: electrospray ionization (ESI) using a TOF mass analyzer (Waters LCT Premiere). Melting points are uncorrected and were measured on a Thomas–Hoover Capillary melting point apparatus. Optical rotation measurements were made on a Perkin–Elmer 241 polarimeter and are quoted in units of 10−1 deg cm2 g1. Infrared spectra were measured using a Shimadzu FTIR–8400S Fourier Transform Infrared Spectrometer. HPLC preparative purification was performed using Chiralpak IC Acquity T3 C–18 column by using 0.02% formic acid, acetonitrile mobile phases, purity was determined by UV area % from HPLC analysis.

4.2. (S)–2–(1–(4–(Benzyloxy)phenyl)–3–hydroxypropan–2–yl)isoindoline–1,3–dione (4a)

Phthalic anhydride (0.82 g, 5.5 mmol) and NEt3 (1.0 mL, 7.7 mmol) were sequentially added to a solution of (S)–2–Amino–3–(4–(benzyloxy)phenyl)propan–1–ol 4 (1.30 g, 5.0 mmol) in toluene (15 mL). The reaction was refluxed for 6 h. After the completion of the reaction, the organic solvents were removed under reduced pressure to provide a sticky oily residue. The resulting residue was purified by silica gel-column chromatography (EtOAc/hexanes, 20 % to 30 %) to obtain the title compound 4a as a white solid (1.03 g, 53%); m.p. 119–121 °C; 1H NMR (400 MHz, CDCl3) δ 7.85 – 7.67 (m, 4H), 7.43 – 7.30 (m, 5H), 7.19 – 7.12 (m, 2H), 6.90 – 6.81 (m, 2H), 5.00 (s, 2H), 4.62 (qd, J = 7.9, 3.5 Hz, 1H), 4.14 – 3.90 (m, 2H), 3.17 (d, J= 8.1 Hz, 2H), 2.66 (s, 1H); 13C NMR (126 MHz, CDCl3) δ 169.0, 157.5, 137.0, 134.1, 131.6, 130.0, 129.7, 128.5, 127.9, 127.5, 123.3, 114.9, 69.9, 62.8, 55.3, 33.9; [α]d24 = −130.8 (c 0.73, CHCl3); IR (film) 3496, 3468, 2943, 2883, 1770, 1697, 1614, 1512, 1398, 1369, 1253, 1174, 1120, 1018, 8733, 721 cm−1; HRMS (ESI): m/z [M+Na]+ calc’d for C24H21NNaO4: 410.1363; found 410.1370.

4.3. (S)–3–(4–(Benzyloxy)phenyl)–2–(1,3–dioxoisoindolin–2–yl)propanal (5)

To a solution of oxalyl chloride (0.66 mL, 7.7 mmol) in DCM (10 mL) was added DMSO (1.1 mL, 15.4 mmol) at – 78 °C, and the resulting solution was stirred at the same temperature for 30 minutes. Then, a solution of 4a (2.0 g, 5.1 mmol) in DCM (20 mL) was added dropwise to the reaction mixture. The reaction mixture was stirred at −78 °C for 2 h, followed by the addition of NEt3 (3.5 mL, 25.5 mmol). The cooling bath was removed after 5 min, and the reaction mixture was allowed to warm to room temperature and stirred for 2 h. The reaction was quenched with 10% aqueous citric acid solution (10 mL), and the mixture was stirred at ambient temperature for an additional 10 min and extracted with DCM (2 X 25 mL). The organic layer was washed with 1.0 N aqueous HCl solution (15 mL), followed by brine, dried over anhydrous MgSO4, and evaporated in vacuo. The residue was crystalized in diethyl ether to get the title compound 5 as a white solid (1.71 g, 86 %). m.p. 152–156 °C; 1H NMR (400 MHz, CDCl3) δ 9.79 (s, 1H), 7.79 (ddd, J = 36.3, 5.5, 3.1 Hz, 4H), 7.44 – 7.29 (m, 5H), 7.14 – 7.07 (m, 2H), 6.88 – 6.81 (m, 2H), 4.99 (s, 2H), 4.96 (dd, J = 10.7, 5.6 Hz, 1H), 3.54 (dd, J = 14.4, 5.6 Hz, 1H), 3.33 (dd, J = 14.4, 10.7 Hz, 1H); 13C NMR (126 MHz, CDCl3) δ 196.3, 167.6, 157.7, 136.8, 134.3, 131.5, 129.9, 128.5, 128.2, 127.9, 115.0, 69.9, 60.1, 32.2; [α]d24 = −137.6 (c 0.51, CHCl3); IR (film) 2926, 1772, 1737, 1710, 1681, 1614, 1583, 1514, 1467, 1450, 1386, 1334, 1313, 1300, 1255, 1178, 1066, 1028, 721 cm−1; HRMS (ESI): m/z [M–H] calc’d for C24H18NO4: 384.1236; found 384.1276.

4.4. (S)–2–(1–(4–(Benzyloxy)phenyl)–4,4,4–trifluoro–3–oxobutan–2–yl)isoindoline–1,3–dione (6)

To a solution of 5 (1.0 g, 2.6 mmol) in THF (10 mL) at −78 °C was added TMSCF3 (1.2 mL, 7.8 mmol), followed by catalytic TBAF (1.0 M in THF). The reaction was stirred at room temperature for 3 h. After complete disappearance of the aldehyde, desilylation was carried out by adding 4.0 N aqueous HCl (0.5 mL) and stirred at ambient temperature for 5 h. The resulting product was extracted with EtOAc. The organic layer was washed with brine solution, dried over anhydrous MgSO4, and evaporated in vacuo. The residue was crystalized in EtOAc to provide the 1,2-addition product in 59% yield (0.70 g), which was subsequently oxidized as described below.

To a solution of 2–((2S)–1–(4–(benzyloxy)phenyl)–4,4,4–trifluoro–3–hydroxybutan–2–yl)isoindoline–1,3–dione (0.70 g, 1.5 mmol) in dry DCM (20 mL) was added Dess–Martin periodinane (0.78 g, 1.8 mmol), and the mixture was stirred at ambient temperature for 3 h. The reaction mixture was then diluted with DCM (5 mL) and poured into a 0.26 M solution of sodium thiosulfate in saturated aqueous sodium bicarbonate. The layers were separated, and the organic phase was washed sequentially with saturated aqueous sodium bicarbonate and water. The combined organic phases were dried over anhydrous MgSO4, filtered, and the solvent was removed in vacuo. The crude compound was purified by silica-gel column chromatography (EtOAc/hexanes, 10% to 30%) to get a hydrate form of (S)–2–(1–(4–(benzyloxy)phenyl)–4,4,4–trifluoro–3–oxobutan–2–yl)isoindoline–1,3–dione. The obtained hydrate was refluxed in toluene (10 mL) in the presence of molecular sieves (4 Å) for 2 h. After cooling to rt, and filtering off the sieves, the solvent was evaporated to afford (S)–2–(1–(4–(benzyloxy)phenyl)–4,4,4–trifluoro–3–oxobutan–2–yl)isoindoline–1,3–dione as a white solid (0.7 g, 95%). m.p. 131–134 °C; 1H NMR (400 MHz, CDCl3) δ 7.78–7.57 (m, 4H), 7.32–7.18 (m, 5H), 6.97 (d, J = 8.4 Hz, 2H), 6.75–6.67 (m, 2H), 5.31 (dd, J = 10.2, 5.5 Hz, 1H), 4.86 (s, 2H), 3.38 (qd, J = 14.4, 7.9 Hz, 2H); 13C NMR (126 MHz, CDCl3) δ 186.0 (q, J = 35.1 Hz), 166.8, 157.9, 136.7, 134.5, 131.1, 130.0, 128.5, 127.9, 127.4, 127.2, 117.2 (q, J = 295.9 Hz), 93.9 (q, J = 32.3 Hz), 69.9, 55.7, 32.1; 19F NMR (376 MHz, CDCl3) δ −75.33 (s, 3F); [α]d24 = −6.6 (c 0.65, CHCl3); IR (film) 3367, 3034, 1768, 1720, 1701, 1610, 1583, 1512, 1467, 1452, 1383, 1247, 1178, 968, 719 cm−1; HRMS (ESI): m/z [M–H] calc’d for C25H17NF3NO4: 452.1104; found 452.1099.

4.5. General Procedure A:

To a solution of triphenylphosphonium bromide derivative 7 (1.5 equiv.) in THF (0.04 M) at −78 °C was added a solution of n–butyl lithium (2.5 M in hexane, 1.5 equiv.). After stirring for 1 h at the same temperature, a solution of CF3–ketone (1.0 equiv.) in THF (0.05 M) was added slowly over a period of 10–15 min, and the reaction was stirred at −78 °C for 1 h. The cooling bath was removed, and the reaction mixture was allowed to warm to ambient temperature and stir overnight. Saturated aqueous NH4Cl was added to the reaction, and the mixture was extracted twice with EtOAc. The organic layer was washed with brine solution, dried over anhydrous MgSO4, and evaporated in vacuo to afford 1:1 E/Z mixture of trifluoromethylalkene products 8.

4.5.1. (S)–2–(6–(Benzyloxy)–1–(4–(benzyloxy)phenyl)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione (8a-Bn)

Following general procedure A, 2.40 g (4.9 mmol) of compound 7a was reacted with 1.50 g (3.3 mmol) of 6. Workup and purification by column chromatography (EtOAc/hexane, 20% to 30%) afforded the title compound as a colorless oil (1.20 g, 62%): 1H NMR (400 MHz, CDCl3) δ 7.84 – 7.65 (m, 4H), 7.48 – 7.28 (m, 10H), 7.12 (dd, J = 27.6, 8.6 Hz, 2H), 6.83 (dd, J = 24.0, 8.6 Hz, 2H), 6.69 – 6.43 (m, 1H), 5.36 (ddd, J = 89.1, 10.0, 6.0 Hz, 1H), 4.98 (d, J = 12.7 Hz, 2H), 4.57 – 4.49 (m, 2H), 3.85 – 3.16 (m, 4H), 2.71 (d, J = 15.5 Hz, 2H); 13C NMR (126 MHz, CDCl3) δ 168.0, 167.6, 157.6, 138.9 (q, J = 3.0 Hz), 137.2 (q, J= 6.4 Hz), 133.9, 131.6, 131.4, 130.1, 129.9, 128.5, 128.4, 127.9, 127.7, 127.6, 124.8 (q, J = 276.3 Hz), 114.8, 72.9, 69.9, 69.8, 68.7, 68.0, 50.8, 50.6, 35.2, 29.0, 28.1, 25.6; 19F NMR (376 MHz, CDCl3) δ −59.80 (3F), −62.45 (3F); [α]d24 = −10.7 (c 0.41, CHCl3); IR (film) 3064, 3032, 1776, 1714, 1610, 1512, 1454, 1384, 1354, 1244, 1174, 1118, 1026, 875, 717, 698, 530 cm−1; HRMS (ESI): m/z [M+Na] + calc’d for C35H30F3NNaO4 608.2019, found 608.2012.

4.5.2. 2–((2S)–1–(4–(Benzyloxy)phenyl)–6–((tetrahydro–2H–pyran–2–yl)oxy)–3–(trifluoromethyl)hex–3-en–2–yl)isoindoline–1,3–dione (8b-OTHP)

Following general procedure A, 2.10 g (5.2 mmol) of 7b, was reacted with 1.60 g (3.5 mmol) of 6. Workup and purification by column chromatography (EtOAc/Hexane, 20% to 30%) afforded the title compound as a colorless oil (1.50 g, 87%): 1H NMR (400 MHz, CDCl3) δ 7.82 – 7.66 (m, 4H), 7.44 – 7.29 (m, 5H), 7.21 – 7.08 (m, 2H), 6.90 – 6.79 (m, 2H), 6.69 – 6.43 (m, 1H), 5.37 (ddd, J = 94.7, 10.3, 6.0 Hz, 1H), 5.03 – 4.96 (m, 2H), 4.70 – 4.59 (m, 1H), 3.96 – 3.03 (m, 5H), 2.74 – 2.35 (m, 2H), 1.93 – 1.43 (m, 6H); 13C NMR (126 MHz, CDCl3) δ 168.0, 167.6, 157.6, 138.9, 138.8, 137.3, 137.2, 136.9, 136.8, 133.9, 131.6, 130.1, 129.9, 129.1, 128.5, 124.8 (q, J = 274.3 Hz), 114.8, 98.8, 69.9, 65.8, 65.4, 62.3, 62.0,50.7, 35.3, 30.5, 29.0, 28.2, 25.4, 19.4, 19.2; 19F NMR (376 MHz, CDCl3) δ −59.78 (dd, J = 19.0, 2.6 Hz, 3F), −62.52 (d, J = 10.4 Hz, 3F); [α]d24 = −0.86 (c 0.46, CHCl3); IR (film) 2945, 2359, 2341, 1712, 1691, 1678, 1629, 1583, 1562, 1546, 1512, 1383, 1352, 1330, 1244, 1165, 1118, 1033, 970, 875, 731, 717 cm−1; HRMS (ESI): m/z [M+Na]+ calc’d for C33H32F3NNaO5: 602.2072, found 602.2084.

4.6.1. (S)–2–(6–Hydroxy–1–(4–hydroxyphenyl)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione (9a)

To a stirred solution of (S)–2–(6–(Benzyloxy)–1–(4–(benzyloxy)phenyl)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione 8a-Bn (0.50 g, 0.8 mmol) and pentamethylbenzene (0.37 g, 2.5 mmol) in dry DCM (5 mL) was added BCl3 (1.0 M in DCM, 2.5 mL, 2.5 mmol) dropwise over 10 min via syringe at −78 °C. After 30 min, the temperature was brought to −30 °C and the reaction stirred for 1h. The progress of the reaction was monitored by TLC, and after the complete consumption of the starting materials, the temperature brought back to −78 °C and quenched with CHCl3–MeOH (10:1, 20 mL). The resulting mixture was warmed to room temperature. The excess organic solvents were removed under reduced pressure. The residue was purified by column chromatography (MeOH/DCM, 5% to 10%) to separate E and Z isomers of 9a.

4.6.1. (S,Z)–2–(6–Hydroxy–1–(4–hydroxyphenyl)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione (Z-9a)

32% yield; White solid, m.p. 132–135 °C; 1H NMR (400 MHz, DMSO–d6) δ 9.16 (s, 1H), 7.91 – 7.75 (m, 4H), 6.95 – 6.90 (m, 2H), 6.69 (t, J = 7.5 Hz, 1H), 6.57 – 6.48 (m, 2H), 5.07 (dd, J = 11.0, 4.9 Hz, 1H), 4.80 (t, J = 5.3 Hz, 1H), 3.54 (td, J = 6.5, 5.3 Hz, 2H), 3.38 (dd, J = 13.6, 11.0 Hz, 1H), 3.13 (dd, J = 13.5, 5.0 Hz, 1H), 2.50 (d, J = 1.9 Hz, 2H); 13C NMR (126 MHz, DMSO–d6) δ 167.4, 156.3, 140.4 (q, J = 3.3 Hz), 135.2, 130.9, 130.2, 127.3, 126.1 (q, J = 27.2 Hz), 125.4, 124.3 (q, J = 276.8 Hz), 123.6, 115.5, 60.0, 50.8, 34.7, 32.4; 19F NMR (376 MHz, DMSO–d6) δ −58.49 (s, 3F); [α]d24 = −5.1 (c 0.62, MeOH); IR (film) 3350, 3131, 1770, 1708, 1683, 1558, 1519, 1379, 1180, 1114, 1041, 713, 682 cm−1; HRMS (ESI): m/z [M–H] calc’d for C21H17F3NO4: 404.1110, found 404.1119.

4.6.2. (S,E)–2–(6–Hydroxy–1–(4–hydroxyphenyl)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione (E-9a)

38% yield; White solid, m.p. 136–140 °C; 1H NMR (400 MHz, DMSO–d6) δ 9.28 (s, 1H), 7.90 (s, 4H), 7.10–7.02 (m, 2H), 6.71–6.64 (m, 2H), 6.53 (td, J = 7.1, 1.7 Hz, 1H), 5.42 (dd, J = 9.3, 7.1 Hz, 1H), 4.83 (t, J = 5.2 Hz, 1H), 3.72 (dd, J = 13.7, 9.4 Hz, 1H), 3.54–3.44 (m, 2H), 3.39–3.31 (m, 1H), 2.56–2.29 (m, 2H); 13C NMR (126 MHz, DMSO–d6) δ 167.9, 156.4, 139.0 (q, J = 6.5 Hz), 135.3, 135.0, 131.4, 131.1, 130.3, 130.2, 128.6, 127.9, 126.0 (q, J = 26.2 Hz), 125.6 (q, J = 274.5 Hz), 123.6, 115.6, 115.5, 59.3, 50.9, 34.9, 31.2; 19F NMR (376 MHz, DMSO–d6) δ −62.80 (s, 3F); [α]d24=1.08 (c 0.46, CHCl3); IR (film) 3317, 3180, 1778, 2873, 1708, 1654, 1599, 1522, 1412, 1315, 1265, 1244, 1114, 1072, 905, 806, 698, 656 cm−1; HRMS (ESI): m/z [M–H] calc’d for C21H17F3NO4: 404.1110, found 404.1129.

4.7. (S,Z)–2–(1–(4–(Benzyloxy)phenyl)–6–hydroxy–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione (9b)

2–((2S)–1–(4–(Benzyloxy)phenyl)–6–((tetrahydro–2H–pyran–2–yl)oxy)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione 8b-OTHP (0.75 g, 1.3 mmol) in methanol (20 mL) was added TBATB (0.12 g, 0.2 mmol) at ambient temperature and stirred for 2.5 h. The reaction was quenched with acetone, diluted with EtOAc and washed with 5% aqueous solution of sodium bisulfite. The organic phase was washed with brine, dried over MgSO4, filtered, evaporated and purified by column chromatography (MeOH/DCM, 1% to 2%) to provide isomers of compound 9b.

4.7.1. (S,Z)–2–(1–(4–(Benzyloxy)phenyl)–6–hydroxy–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione (Z-9b)

30% yield; Clear oil; 1H NMR (400 MHz, CDCl3) δ 7.76 (dd, J = 5.5, 3.1 Hz, 2H), 7.69 (dd, J = 5.5, 3.1 Hz, 2H), 7.41 – 7.35 (m, 4H), 7.35 – 7.29 (m, 1H), 7.09 (d, J = 8.3 Hz, 2H), 6.82 (d, J = 8.4 Hz, 2H), 6.63 (t, J = 7.8 Hz, 1H), 5.27 (dd, J = 11.0, 5.2 Hz, 1H), 4.98 (s, 2H), 3.83 (td, J = 6.0, 2.3 Hz, 2H), 3.54 (dd, J = 13.8, 10.9 Hz, 1H), 3.23 (dd, J = 13.8, 5.2 Hz, 1H), 2.83 – 2.56 (m, 2H); 13C NMR (126 MHz, CDCl3) δ 167.6, 157.6, 139.1 (q, J =3.1 Hz), 136.8, 134.0, 131.3, 129.9, 127.9 (q, J = 28.0 Hz), 128.8, 128.5, 127.9, 127.4, 123.6 (q, J = 277.0 Hz), 123.3, 114.9, 69.8, 61.6, 50.5, 35.1, 31.9; 19F NMR (376 MHz, CDCl3) δ −59.52 (S, 3F); [α]d24 = −0.55 (c 0.54, MeOH); IR (film) 3518, 3475, 2931, 2879, 1776, 1712, 1610, 1512, 1467, 1454, 1379, 1354, 1244, 1174, 1116, 1039, 964, 910, 873, 790, 717, 696, 680 cm−1; HRMS (ESI): m/z [M+H]+ calc’d for C28H25F3NO4: 496.1657, found 496.1665.

4.7.2. (S,E)–2–(1–(4–(Benzyloxy)phenyl)–6–hydroxy–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione (E-9b)

41% yield; Clear oil; 1H NMR (400 MHz, CDCl3) δ 7.76 (ddd, J = 39.1, 5.5, 3.1 Hz, 4H), 7.47–7.29 (m, 5H), 7.21–7.14 (m, 2H), 6.92–6.86 (m, 2H), 6.42 (tq, J = 8.2, 1.6 Hz, 1H), 5.46 (t, J = 8.2 Hz, 1H), 5.03 (s, 2H), 3.78 (dd, J = 14.1, 8.5 Hz, 1H), 3.68 (td, J = 6.8, 6.0, 3.8 Hz, 2H), 3.57 (dd, J = 14.0, 8.1 Hz, 1H), 2.70–2.21 (m, 2H), 1.87–1.74 (m, 1H); 13C NMR (126 MHz, CDCl3) δ 168.2, 157.6, 137.3 (d, J = 6.4 Hz), 136.9, 134.1, 131.5, 130.3, 129.7, 128.5, 127.9, 127.3, 127.1 (q, J = 27.8 Hz), 127.0, 123.8 (q, J = 275.1 Hz), 123.3, 114.9, 69.9, 61.0, 51.2, 35.4, 30.8; 19F NMR (376 MHz, CDCl3) δ −62.51 (s, 3F); [α]d24 = 1.1 (c 0.49, CHCl3); IR (film) 3518, 3475, 2931, 2879, 1776, 1712, 1610, 1512, 1467, 1454, 1379, 1354, 1244, 1174, 1116, 1039, 964, 910, 873, 790, 717, 696, 680 cm−1; HRMS (ESI): m/z [M+H]+ calc’d for C28H25F3NO4: 496.1657, found 496.1624.

4.8. General Procedure B

To a solution of (S)–2–(6–Hydroxy–1–(4–hydroxyphenyl)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione Z-9a (0.10 g, 0.2 mmol, 1.0 equiv.) in wet acetonitrile (0.5 M), (0.75 v % water), a stock solution of H5IO6/CrO3 (1.2 mL, 0. 4 mmol, 2.0 equiv.) [prepared by dissolving H5IO6 (11.40 g, 50.0 mmol) and CrO3 (0. 020 g, 1.2 mol %) in wet MeCN (0.75 v % water) to a volume of 114 mL] was added slowly at 0 °C for 30–60 minutes at 0 °C. The mixture was stirred at 0 °C for 0.5 h and the completion of the reaction was confirmed by 19F NMR. The reaction was quenched with aqueous solution of Na2HPO4 (0.60 g in 10 mL of water). The reaction mixture was extracted with EtOAc (2 X 12 mL) and the organic layer was separated and washed with aqueous NaHSO3 solution (15 mL) followed by brine (2 × 10 mL). The organic layer was concentrated in vacuo, and the resulting residue which was purified by silica-gel column chromatography.

4.8.1. (S, Z)–5–(1,3–Dioxoisoindolin–2–yl)–6–(4–hydroxyphenyl)–4–(trifluoromethyl)hex–3–enoic acid (Z-10a)

Following general procedure B, Z-9a (0.10 g, 0.2 mmol, 1.0 equiv.) was reacted with stock solution of H5IO6/CrO3 (1.2 mL, 0.4 mmol, 2.0 equiv.) Workup and chromatographic purification (MeOH/DCM, 1% to 2%) afforded compound Z-10a as an oil (0.03 g, 32%): 1H NMR (400 MHz, Methanol–d4) δ 7.78 (s, 4H), 6.99 – 6.90 (m, 3H), 6.60 – 6.56 (m, 2H), 5.20 (dd, J = 11.5, 4.8 Hz, 1H), 3.60 – 3.37 (m, 3H), 3.20 (dd, J = 13.6, 4.8 Hz, 1H); 13C NMR (126 MHz, DMSO–d6) δ 171.4, 170.8, 167.4, 166.2, 156.4, 137.0, 135.5, 135.3, 130.9, 130.3, 130.0 (d, J = 28.6 Hz), 127.2, 127.1, 127.0, 126.9, 125.1 (q, J = 276.6 Hz), 123.7, 122.9, 115.6, 60.2, 50.6, 34.7, 33.7, 21.2, 14.5; 19F NMR (376 MHz, CDCl3) δ −59.14 (s, 3F); [α]d24 = + 0.86 (c 0.65, CHCl3); IR (film) 3745, 2360, 1770, 1708, 1683, 1558, 1519, 1379, 1361, 1259, 1190, 873, 837, 738, 713 cm−1; HRMS (ESI): m/z [M–H] calcd for C21H15F3NO5: 418.0908, found 418.0887.

4.8.2. (S, E)–5–(1,3–Dioxoisoindolin–2–yl)–6–(4–hydroxyphenyl)–4–(trifluoromethyl)hex–3–enoic acid (E-10a)

General procedure B was followed using (S, E)–2–(6–Hydroxy–1–(4–hydroxyphenyl)–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione E-9a (0.50 g, 1.2 mmol) and added stock solution of H5IO6/CrO3 (7.3 mL, 2.4 mmol). Workup and chromatographic purification (MeOH/DCM, 1% to 2%) afforded compound E-10a as an oil (0.17 g, 33%); 1H NMR (400 MHz, DMSO–d6) δ 12.70 (s, 1H), 9.24 (s, 1H), 7.85 (s, 4H), 7.04 – 6.97 (m, 2H), 6.66 – 6.58 (m, 3H), 5.24 (dd, J = 9.5, 6.7 Hz, 1H), 3.68 – 3.41 (m, 2H), 3.33 – 3.19 (m, 2H); 13C NMR(126 MHz, DMSO–d6) δ 171.4, 167.7, 166.5, 156.5, 135.3, 134.6, 134.6, 134.5 (q, J = 3.0 Hz), 131.1, 130.3, 130.1 (q, J = 25.6 Hz), 130.0, 127.6, 126.8, 126.6, 125.3 (q, J = 277.3 Hz), 123.7, 123.5, 115.6, 115.4, 67.4, 50.7, 34.6, 33.1, 32.7, 25.5; 19F NMR (376 MHz, DMSO–d6) δ −61.75 (s, 3F); [α]d24 = −0.88 (c 0.65, CH2Cl2); IR (film) 3748, 2889, 2343, 1770, 1708, 1683, 1558, 1519, 1379, 1361, 1259, 1190, 873, 837, 756, 713, 682 cm−1.

4.8.3. (S,Z)–6–(4–(Benzyloxy)phenyl)–5–(1,3–dioxoisoindolin–2–yl)–4–(trifluoromethyl)hex–3–enoic acid (Z-10b)

General procedure B was followed using (S,Z)–2–(1–(4–(Benzyloxy)phenyl)–6–hydroxy–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione Z-9b (0.10 g, 0.2 mmol) and a stock solution of H5IO6/CrO3 (1.2 mL, 0.4 mmol). Workup and chromatographic purification (MeOH/DCM, 1% to 2%) afforded the title compound as colorless oil (0.07 g, 70%): 1H NMR (400 MHz, CDCl3) δ 7.77 (ddd, J = 38.0, 5.5, 3.1 Hz, 4H), 7.47 – 7.29 (m, 5H), 7.16 (d, J = 8.2 Hz, 2H), 6.89 (d, J = 8.2 Hz, 2H), 6.63 (t, J = 6.9 Hz, 1H), 5.32 (t, J = 8.2 Hz, 1H), 5.02 (s, 2H), 3.73 (dd, J = 14.1, 8.7 Hz, 1H), 3.61 – 3.50 (m, 2H), 3.33 – 3.21 (m, 1H); 13C NMR (126 MHz, CDCl3) δ 174.3, 167.8, 157.7, 136.8, 134.1, 131.7, 131.6, 131.5, 131.4 (d, J = 6.4 Hz), 130.3, 130.1, 129.2, 128.5, 128.2, 128.0 (d, J = 27.6 Hz), 127.9, 127.8, 127.4, 126.8, 125.0 (d, J = 275.0 Hz), 124.6, 122.4, 120.2, 115.0, 114.9, 69.9, 61.0, 51.2, 50.8, 35.1, 32.1, 30.7; 19F NMR(376 MHz, CDCl3) δ −63.16 (s, 3F), [α]d24 = −1.3 (c 0.44, CH2Cl2); IR (film) 3034, 2827, 1772, 1716, 1508, 1456, 1375, 1361, 1329, 1317, 1301, 1286, 1184, 1120, 1097, 1085, 1045, 837, 738, 711, 696, 530, 434 cm−1; HRMS (ESI): m/z [M–H] calc’d for C28H21F3NO5: 508.1377, found 508.1370.

4.8.4. (S,E)–6–(4–(Benzyloxy)phenyl)–5–(1,3–dioxoisoindolin–2–yl)–4–(trifluoromethyl)hex–3–enoic acid (E-10b)

General procedure B was followed using (S,E)–2–(1–(4–(Benzyloxy)phenyl)–6–hydroxy–3–(trifluoromethyl)hex–3–en–2–yl)isoindoline–1,3–dione E-9b (0.66 g, 1.3 mmol) and added stock solution of H5IO6/CrO3 (7.9 mL, 3.3 mmol). Workup and chromatographic purification (DCM/MeOH 1:0 → 9.8:0.2) afforded the title compound as an oil (0.45 g, 67%): 1H NMR (400 MHz, CDCl3) δ 7.73 (ddt, J = 26.7, 5.6, 3.1 Hz, 4H), 7.43 – 7.29 (m, 5H), 7.10 (d, J = 8.3 Hz, 2H), 6.87 (t, J = 7.1 Hz, 1H), 6.81 (d, J = 8.3 Hz, 2H), 5.27 (dd, J = 11.2, 5.0 Hz, 1H), 4.98 – 4.94 (m, 2H), 3.68 – 3.51 (m, 3H), 3.26 (dd, J = 13.8, 5.1 Hz, 1H); 13C NMR (126 MHz, CDCl3) δ 174.4, 167.5, 157.6, 136.8, 134.0, 132.4 (q, J = 3.0 Hz), 132.3, 131.3, 129.9, 129.2 (q, J = 27.2 Hz), 128.9, 128.6, 128.5, 127.9, 127.4, 126.7, 124.5 (q, J = 277.3 Hz), 123.3, 122.3, 120.1, 114.9, 69.8, 50.4, 35.1, 32.8, 14.2; 19F NMR (376 MHz, CDCl3) δ −60.49 (s, 3F); [α]d24 = − 1.3 (c 0.6, CH2Cl2); IR (film) 3034, 2827, 1772, 1716, 1508, 1456, 1375, 1361, 1329, 1317, 1301, 1286, 1184, 1120, 1097, 1085, 1045, 837, 738, 711, 696, 530, 434 cm−1; HRMS (ESI): m/z [M+HCOOH]+ calc’d for C28H22F3NO5HCOOH: 555.1505, found 555.1493.

4.9. Procedure for Peptide Coupling

To a solution of (S)–6–(4–(Benzyloxy)phenyl)–5–(1,3–dioxoisoindolin–2–yl)β4–(trifluoromethyl)hex–3–enoic acid 10b (0.10 g, 0.2 mmol, 1.0 equiv.) in DCM (6 mL) was added thionyl chloride (0.028 mL, 0.4 mmol, 2.0 equiv.) at ambient temperature and the progress of the reaction was monitored by 19F NMR. After completion of the reaction, the solvent was evaporated under reduced pressure. The obtained residue is dissolved in dry DCM (6 mL), added NEt3 (0.18 mL, 0.5 mmol, 2.5 equiv) and Gly–Phe–Leu–OR4 tripeptide 11 (0.15 g, 0.4 mmol, 2.0 equiv.) in DCM (5 mL) at 0 °C and the reaction mixture is stirred at the same temperature for 1 h. After the completion, quenched with 1N aqueous HCl, extracted with EtOAc, and washed with saturated NaHCO3 solution. The combined organic phases were washed with brine, dried over MgSO4, filtered, evaporated and purified by silica gel chromatography (EtOAc/Hexanes 40% to 60%) to afford the respective peptide.

4.9.1. Methyl ((S,Z)–6–(4–(benzyloxy)phenyl)–5–(1,3–dioxoisoindolin–2–yl)–4–(trifluoromethyl)hex–3–enoyl)glycyl–L–phenylalanyl–L–leucinate (Z-12b)

The peptide coupling procedure was followed using (S,Z)–6–(4–(Benzyloxy)phenyl)–5–(1,3–dioxoisoindolin–2–yl)–4–(trifluoromethyl)hex–3–enoic acid Z-10b (0.23 g, 0.4 mmol), SOCl2 (0.098 mL) NEt3 (0.18 mL) and Gly–Phe–Leu–OMe tripeptide 11a (0.34 g, 0.8 mmol). Purified by silica gel chromatography (EtOAc/Hexanes 40% to 60%) to afford the respective peptide as a white solid (0.305 g, 80%). HPLC purity (UV area percent) 97.0 %; m.p. 85–89 °C; 1H NMR (400 MHz, CDCl3) δ 7.83 (dq, J = 7.0, 4.3, 3.5 Hz, 2H), 7.73 (dd, J = 5.5, 3.1 Hz, 2H), 7.44 – 7.31 (m, 6H), 7.26 – 7.13 (m, 5H), 6.89 (d, J = 8.3 Hz, 2H), 6.70 (dd, J = 12.2, 6.8 Hz, 2H), 6.60 – 6.52 (m, 1H), 6.30 (dd, J = 10.4, 8.0 Hz, 1H), 5.35 (t, J = 8.2 Hz, 1H), 5.02 (s, 2H), 4.71 (q, J = 7.1 Hz, 1H), 4.58 – 4.48 (m, 1H), 3.91 (dt, J = 9.3, 4.9 Hz, 2H), 3.71 (d, J = 1.2 Hz, 4H), 3.59 (dd, J = 14.1, 8.2 Hz, 1H), 3.33 (dd, J = 16.2, 8.6 Hz, 1H), 3.19 – 2.95 (m, 3H), 1.62 – 1.43 (m, 2H), 0.89 (dd, J = 6.1, 2.3 Hz, 6H); 13C NMR (126 MHz, CDCl3) δ 172.7, 170.1, 169.3, 168.5, 168.1, 157.7, 136.7, 136.2, 134.3, 133.1, 132.9 (q, J = 6.2 Hz), 131.4, 128.6, 128.5, 128.0, 127.5, 127.1, 126.8, 124.6 (q, J = 275.2 Hz), 122.5, 120.3, 115.0, 77.2, 69.9, 54.3, 52.3, 51.1, 50.9, 43.6, 41.2, 38.0, 37.9, 35.2, 35.0, 24.7, 22.6, 21.9; 19F NMR (376 MHz, CDCl3) δ −62.90 (d, J = 9.5 Hz, 3F); [α]d24 −9.5 (c 0.58, CHCl3); IR (film) 3296, 3086, 3064, 3032, 2958, 2931, 1776, 1712, 1614, 1631, 1546, 1529, 1512, 1467, 1383, 1369, 1352, 1246, 1124, 1026, 732, 698, 667 cm−1; HRMS (ESI): m/z [M+Na]+ calcd for C46H47F3N4O8Na: 863.3244, found 863.3251.

4.9.2. Tert–butyl ((S,E)–6–(4–(benzyloxy)phenyl)–5–(1,3–dioxoisoindolin–2–yl)–4–trifluoromethyl)hex–3–enoyl)glycyl–L–phenylalanyl–L–leucinate (E-12b)

Above coupling procedure was followed using (S,E)–6–(4–(benzyloxy)phenyl)–5–(1,3–dioxoisoindolin–2–yl)–4–(trifluoromethyl)hex–3–enoic acid E-10b (0.20 g, 0.4 mmol) SOCl2 (0.085 mL, 1.1 mmol) and NEt3 (0.18 mL, 0.8 mmol), and Gly–Phe–Leu–OMe tripeptide 11b (0.30 g, 0.8 mmol). After workup and purification by silica gel chromatography (EtOAc/Hexanes, 40% to 60%) afforded the respective peptide as a white solid (0.28 g, 83%). HPLC purity (UV area percent) 99.5%; m.p. 80–86 °C; 1H NMR (400 MHz, CDCl3) δ 7.79 – 7.66 (m, 4H), 7.43 – 7.29 (m, 5H), 7.29 – 7.13 (m, 5H), 7.08 (dd, J = 8.6, 3.5 Hz, 2H), 6.83 – 6.76 (m, 4H), 6.48 – 6.30 (m, 1H), 5.27 (dd, J = 11.1, 4.9 Hz, 1H), 4.97 (d, J = 2.5 Hz, 2H), 4.75 (dd, J = 7.6, 3.4 Hz, 1H), 4.42 (t, J = 5.4 Hz, 1H), 4.07 – 3.85 (m, 1H), 3.57 – 3.43 (m, 2H), 3.40 – 3.18 (m, 2H), 3.12 (dd, J = 6.6, 4.1 Hz, 2H), 1.57 (dd, J = 8.8, 4.9 Hz, 2H), 1.46 (s, 9H), 0.91 (dd, J = 6.0, 4.1 Hz, 6H); 13C NMR (126 MHz, CDCl3) δ 171.6, 171.5, 169.9, 169.5, 169.4, 168.4, 167.6, 157.7, 136.8, 136.2, 136.1, 134.1, 133.5, 133.5, 131.2, 130.0, 129.4, 129.3, 128.6, 128.5, 127.9, 127.4, 127.0, 126.6, 124.4, 123.4, 120.0, 114.9, 81.9, 69.8, 54.2, 51.5, 50.4, 43.7, 43.6, 41.6, 38.2, 38.1, 35.4, 35.3, 35.0, 27.9, 24.8, 22.6, 22.1; 19F NMR (376 MHz, CDCl3) δ −59.79 (d, J = 17.6 Hz); [α]d24 = −12.6 (c 0.58, CHCl3); IR (film) 3280, 3086, 3064, 2958, 2931, 1776, 1712, 1614, 1631, 1546, 1529, 1512, 1467, 1383, 1369, 1352, 1246, 1124, 1026, 732, 698, 667, 401 cm−1; HRMS (ESI): m/z [M+H]+ calc’d for C49H54F3N4O8Na: 883.3894, found 883.3887.

Supplementary Material

1

Highlights.

  • A trifluoromethylalkene isopolar peptidomimetic is introduced into the Leu-Enkephalin framework.

  • A Wittig olefination strategy enables access to the trifluoromethylalkene containing dipeptide mimetic.

  • Conditions were optimized to overcome a problematic amide coupling reaction and minimize undesired isomerization of the alkene.

Acknowledgements

Research reported in this publication was supported by the National Institute on Drug Abuse of the National Institutes of Health under award number DA036730. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. Support for the NMR instrumentation was provided by NIH Shared Instrumentation Grant S10D016360. We are thankful to Benjamin Neuenswander for HPLC purification and analysis support.

Footnotes

Appendix A. Supplementary data

Supplementary data related to this article can be found at https://doi.org/10.1016/j.jfluchem.XXXX.XX.XXX

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Luca G, Alessandra T, Federico S, Peptides and peptidomimetics in medicine, surgery and biotechnology, Curr. Med. Chem 13 (2006) 2449–2466. [DOI] [PubMed] [Google Scholar]
  • 2.Cornford EM, Braun LD, Crane PD, Oldendorf WH, Blood-Brain Barrier Restriction of Peptides and the Low Uptake of Enkephalins*, Endocrinology. 103 (1978) 1297–1303. [DOI] [PubMed] [Google Scholar]
  • 3.Roques B P NF, Dual inhibitors of enkephalin-degrading enzymes (neutral endopeptidase 24.11 and aminopeptidase N) as potential new medications in the management of pain and opioid addiction, NIDA Res Monogr 147 (1995) 104–145. [PubMed] [Google Scholar]
  • 4.Veber DF, Freidinger RM, The design of metabolically-stable peptide analogs, Trends Neurosci. 8 (1985) 392–396. [Google Scholar]
  • 5.Wipf P, Henninger TC, Geib SJ, Methyl- and (trifluoromethyl)alkene peptide isosteres: Synthesis and evaluation of their potential as β-Turn promoters and peptide mimetics, J. Org. Chem 63 (1998) 6088–6089. [DOI] [PubMed] [Google Scholar]
  • 6.Knipp GT, Vander Velde DG, Siahaan TJ, Borchardt RT, The effect of β-turn structure on the passive diffusion of peptides across caco-2 cell monolayers, Pharm. Res. 14 (1997) 1332–1340. [DOI] [PubMed] [Google Scholar]
  • 7.Borchardt RT, Optimizing oral absorption of peptides using prodrug strategies, J. Controlled Release 62 (1999) 231–238. [DOI] [PubMed] [Google Scholar]
  • 8.Gudmundsson OS, Jois SDS, Velde DGV, Siahaan TJ, Borchardt RT, Wang B, The effect of conformation on the membrane permeation of coumarinic acid- and phenylpropionic acid-based cyclic prodrugs of opioid peptides, J. Peptide. Res 53 (2003) 383–392. [DOI] [PubMed] [Google Scholar]
  • 9.Wang B, Wang W, Zhang H, Shan D, Nimkar K, Gudmundsson O, Gangwar S, Siahaan T, Borchardt RT, Synthesis and evaluation of the physicochemical properties of esterase-sensitive cyclic prodrugs of opioid peptides using coumarinic acid and phenylpropionic acid linkers, J. Peptide. Res 53 (2003) 370–382. [DOI] [PubMed] [Google Scholar]
  • 10.Molteni M, Bellucci MC, Bigotti S, Mazzini S, Volonterio A, Zanda M, Ψ[CH(CF3)NH]Gly-peptides: synthesis and conformation analysis, Org. Biomol. Chem 7 (2009) 2286–2296. [DOI] [PubMed] [Google Scholar]
  • 11.Sani M, Volonterio A, Zanda M, The trifluoroethylamine function as peptide bond replacement, ChemMedChem 2 (2007) 1693–1700. [DOI] [PubMed] [Google Scholar]
  • 12.Xiao J, Weisblum B, Wipf P, Electrostatic versus Steric Effects in Peptidomimicry: Synthesis and Secondary Structure Analysis of Gramicidin S Analogues with (E)-Alkene Peptide Isosteres, J. Am. Chem. Soc 127 (2005) 5742–5743. [DOI] [PubMed] [Google Scholar]
  • 13.Koksch B, Sewald N, Hofmann H-J, Burger K, Jakubke H-D, Proteolytically stable peptides by incorporation of α-Tfm amino acids, J. Pept. Sci 3 (1998) 157–167. [DOI] [PubMed] [Google Scholar]
  • 14.Bodnar RJ, Klein GE, Endogenous opiates and behavior: 2005, Peptides. 27 (2006) 3391–3478. [DOI] [PubMed] [Google Scholar]
  • 15.Weinberger SB, Martinez JL, Characterization of hydrolysis of [leu]enkephalin and D-ala2-[L-leu]enkephalin in rat plasma, J. Pharmacol. Exp. Ther 247 (1988) 129–135. [PubMed] [Google Scholar]
  • 16.Thompson SE, Audus KL, Leucine-enkephalin metabolism in brain microvessel endothelial cells, Peptides. 15 (1994) 109–116. [DOI] [PubMed] [Google Scholar]
  • 17.Bak A, Fich M, Larsen BD, Frokjaer S, Friis GJ, N-terminal 4-imidazolidinone prodrugs of Leu-enkephalin: synthesis, chemical and enzymatic stability studies, Eur. J. Pharm. Sci 7 (1999) 317–323. [DOI] [PubMed] [Google Scholar]
  • 18.Proteau-Gagné A, Bournival V, Rochon K, Dory YL, Gendron L, Exploring the Backbone of Enkephalins To Adjust Their Pharmacological Profile for the δ-Opioid Receptor, ACS Chem. Neurosci 1 (2010) 757–769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Karad SN, Pal M, Crowley RS, Prisinzano TE, Altman RA, Synthesis and opioid activity of Tyr1-Ψ[(Z)CF=CH]-Gly2 and Tyr1-Ψ[(S)/(R)-CF3CH-NH]-Gly2 leu-enkephalin fluorinated peptidomimetics, ChemMedChem 12 (2017) 571–576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Altman RA, Sharma KK, Rajewski LG, Toren PC, Baltezor MJ, Pal M, Karad SN, Tyr1-Ψ[(Z)CF=CH]-Gly2 Fluorinated Peptidomimetic Improves Distribution and Metabolism Properties of Leu-Enkephalin, ACS Chem. Neurosci 9 (2018) 1735–1742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Carpino LA, Gao HS, Ti GS, Segev D, Thioxanthene dioxide based amino-protecting groups sensitive to pyridine bases and dipolar aprotic solvents, J. Org. Chem 54 (1989) 5887–5897. [Google Scholar]
  • 22.Hintermann L, Comprehensive organic name reactions and reagents. By Zerong Wang, Angew. Chem. Int. Ed 49 (2010) 2659–2660. [Google Scholar]
  • 23.Kolb M, Barth J, Neises B, Synthesis of fluorinated α-amino ketones part I: α-benzamidoalkyl mono- di- and trifluoromethyl ketones, Tetrahedron Lett 27 (1986) 1579–1582. [Google Scholar]
  • 24.Godfrey AG, Brooks DA, Hay LA, Peters M, McCarthy JR, Mitchell D, Application of the dakin–west reaction for the synthesis of oxazole-containing dual PPARα/γ agonists, J. Org. Chem. 68 (2003) 2623–2632. [DOI] [PubMed] [Google Scholar]
  • 25.Kawase M, Saito S, Kurihara T, Convenient synthesis of alpha-trifluoromethylated acyloins from alpha-hydroxy or alpha-amino acids, Chem. Pharm. Bull. (Tokyo) 48 (2000) 1338–1343. [DOI] [PubMed] [Google Scholar]
  • 26.Wiedemann J, Heiner T, Mloston G, Prakash GKS, Olah GA, Direct preparation of trifluoromethyl tetones from carboxylic esters: Trifluoromethylation with (trifluoromethyl)trimethylsilane, Angew. Chem. Int. Ed 37 (1998) 820–821. [DOI] [PubMed] [Google Scholar]
  • 27.Rudzinski DM, Kelly CB, Leadbeater NE, A Weinreb amide approach to the synthesis of trifluoromethylketones, Chem. Commun 48 (2012) 9610–9612. [DOI] [PubMed] [Google Scholar]
  • 28.Richter JM, Whitefield BW, Maimone TJ, Lin DW, Castroviejo MP, Baran PS, Scope and mechanism of direct indole and pyrrole couplings adjacent to carbonyl compounds: Total synthesis of acremoauxin A and oxazinin 3, J. Am. Chem. Soc 129 (2007) 12857–12869. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Ziegler FE, Klein SI, Pati UK, Wang TF, Acyclic diastereoselection as a synthetic route to quassinoids: a Claisen rearrangement based strategy for bruceantin, J. Am. Chem. Soc 107 (1985) 2730–2737. [Google Scholar]
  • 30.Zhao M, Li J, Song Z, Desmond R, Tschaen DM, Grabowski EJJ, Reider PJ, A novel chromium trioxide catalyzed oxidation of primary alcohols to the carboxylic acids, Tetrahedron Lett 39 (1998) 5323–5326. [Google Scholar]
  • 31.Kunishima M, Kawachi C, Hioki K, Terao K, Tani S, Formation of carboxamides by direct condensation of carboxylic acids and amines in alcohols using a new alcohol- and water-soluble condensing agent: DMT-MM, Tetrahedron. 57 (2001) 1551–1558. [Google Scholar]
  • 32.Li H, Jiang X, Ye Y.-h., Fan C, Romoff T, Goodman M , 3-(Diethoxyphosphoryloxy)-1,2,3- benzotriazin-4(3H)-one (DEPBT): A new coupling reagent with remarkable resistance to racemization, Org. Lett 1 (1999) 91–94. [DOI] [PubMed] [Google Scholar]
  • 33.Hu L, Xu S, Zhao Z, Yang Y, Peng Z, Yang M, Wang C, Zhao J, Ynamides as Racemization-Free Coupling Reagents for Amide and Peptide Synthesis, J. Am. Chem. Soc 138 (2016) 13135–13138. [DOI] [PubMed] [Google Scholar]
  • 34.Cody WL, Greene's protective groups in organic synthesis. Fourth Edition By Wuts Peter G. M. and Greene Theodora W.. John Wiley & Sons, Inc., Hoboken, NJ: 2006., J. Med. Chem 50 (2007). [Google Scholar]
  • 35.DeShong P, Protecting Groups, 3rd Edition By Kocienski Philip J. (University of Leeds). Georg Thieme Verlag: Stuttgart, New York: 2005., J. Am. Chem. Soc 128 (2006). [Google Scholar]
  • 36.Maryanoff BE, Greco MN, Zhang H-C, Andrade-Gordon P, Kauffman JA, Nicolaou KC, Liu A, Brungs PH, Macrocyclic Peptide Inhibitors of Serine Proteases. Convergent Total Synthesis of Cyclotheonamides A and B via a Late-Stage Primary Amine Intermediate. Study of Thrombin Inhibition under Diverse Conditions, J. Am. Chem. Soc 117 (1995) 1225–1239. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1

RESOURCES