Abstract
A novel method for the synthesis of epoxydibenzo[b,f][1,5]diazocines exhibiting a V-shaped molecular architecture is reported. The unique approach is based on unprecedented base-catalyzed, solvent-free autocondensation and cross-condensation of fluorinated o-aminophenones. The structure of the newly synthesized diazocines was confirmed independently by X-ray analysis and chiroptical methods. The rigidity of the diazocine scaffold allowed for the separation of the racemate into single enantiomers that proved to be thermally stable up to 140 °C. Furthermore, the inertness of the diazocine scaffold was demonstrated by performing a series of typical transformations, including transition metal-catalyzed reactions, proceeding without affecting the bis-hemiaminal subunit.
Introduction
Molecules possessing rigid structures with defined curvature constitute the underpinnings of the rapidly growing supramolecular chemistry area. Those small building blocks decorated with appropriate functional groups enable the formation of higher-order structures as a result of self-assembly, which is of use in many areas. The most impressive examples include molecular tweezers,1 capsules,2 or cages.3 Among the molecular building blocks mentioned above, marked interest has been focused on Tröger’s base, a small molecule with a great history (structure A in Scheme 1). Its unique rigid V-shaped structure, confirmed almost 50 years after its first synthesis,4 has spawned an enormous number of applications in many areas, such as molecular recognition,5 metal catalysis, and organometallic6 and medicinal chemistry.7 The basis for widespread application arises from its trivial synthesis, carried out directly from aniline (or its derivatives) and paraformaldehyde (Scheme 1). This clearly underlines that each new, easily accessible building block for the construction of the supramolecular architecture stimulates an enormous development of the field. For these reasons, sustainable, practical methods for the synthesis of bent molecular bricks are still highly desirable.
Scheme 1. Examples of Rigid Molecules with V-Shaped Geometry.
The long-standing interest in Tröger’s base (A) has recently culminated in the efficient synthesis of its heteroatom analogues, epiminodibenzodiazocines, bearing a nitrogen bridge (structure B in Scheme 1).8 In contrast, to the best of our knowledge, no efficient synthetic approach to diazocines, bearing an oxygen bridge, has been developed so far (structure C in Scheme 1). We anticipated that the missing link should be easily provided by the autocondensation of o-aminophenones, providing a new scaffold for supramolecular chemistry. A careful inspection of the literature data revealed that epoxydibenzo[b,f][1,5]diazocines have been isolated as byproducts in the synthesis of heterocycles9 and natural product degradation studies in some cases, albeit in marginal yields.10 Herein, we wish to report an unprecedented solvent-free synthesis of epoxydibenzo[b,f][1,5]diazocines with a well-defined, rigid V-shaped structure from fluorinated o-aminophenones.
Results and Discussion
Our working hypothesis was based on the assumption that fluorinated o-aminophenones could undergo base-catalyzed self-condensation (Table 1). To prove this, we initially screened more than 10 solvents of different polarity in the autocondensation reaction of 1a, catalyzed by N,N,N′,N′-tetramethylguanidine (TMG) (for details, see Scheme S1). The respective diazocine 2a was formed in all cases in good to excellent yield without the concomitant formation of byproducts such as imine or cyclic bisimine (Table 1). Astonishingly, the best results in terms of conversion and yield were achieved under solvent-free conditions, leading to diazocine 2a in 93% yield on a 0.5 mmol scale and finally in 95% yield on a 4.5 mmol scale. It should be mentioned that nitrogen-protected aminophenones 1b–d, including an acidic sulfonamide (1d), failed to react, whereas o-aminobenzaldehyde decomposed completely under solvent-free conditions (for details, see the Supporting Information).
Table 1. Effect of the Solvent in the Formation of Epoxydibenzo[b,f][1,5]diazocine (±)-2a.
| entry | R | solvent | conversion (%)a | yield of 2a (%)b |
|---|---|---|---|---|
| 1 | H (1a) | DMF | 94 | 87 |
| 2 | H (1a) | DMSO | 97 | 84 |
| 3 | H (1a) | Py | 90 | 86 |
| 4 | H (1a) | MeCN | 99 | 87 |
| 5 | H (1a) | TMG | 98 | 18 |
| 6 | H (1a) | (CH2OH)2 | 98 | 73 |
| 7 | H (1a) | i-PrOH | 98 | 72 |
| 8 | H (1a) | n-BuOH | 84 | 29 |
| 9 | H (1a) | water | 97 | 29 |
| 10 | H (1a) | 1,4-dioxane | 74 | 65 |
| 11 | H (1a) | DCE | 37 | 15 |
| 12 | H (1a) | toluene | 61 | 49 |
| 13 | H (1a) | n-heptane | 44 | 24 |
| 14 | H (1a) | – | 99 | 93 (95c) |
| 15 | PMBd (1b) | – | <5 | <5 |
| 16 | Tre (1c) | – | <5 | <5 |
| 17 | Tsf (1d) | – | <5 | <5 |
Conversion based on GC; naphtalene as the internal standard; all reactions conducted on a 0.5 mmol scale.
Yield estimated from the calibration curve.
On a 4.5 mmol scale.
PMB, p-methoxybenzyl.
Tr, trityl.
Ts, p-toluenesulfonyl.
With the optimal conditions secured, the scope of the method was explored. First, a group of trifluoromethyl aminophenones bearing electron-withdrawing and electron-donating groups in the para position to the nitrogen were investigated (Scheme 2). Generally, the formation of epoxydibenzo[b,f][1,5]diazocines proceeded in high yields, and the presence of halogen atoms (including fluorine) 2c, alkyl ester 2h, dimethylamine 2g, or methoxy function 2f was admirably tolerated. Only incorporation of the trifluoromethyl group that exerts a strong positive σ-inductive effect has delivered diazocine 2i in a low 23% yield. A further increase in the reaction time to 48 h slightly increased the yield to 39% (for details, see the Experimental Section). Uniformly, diazocine 2j carrying a perfluorinated side chain was also isolated in a moderate 41% yield. In contrast, the presence of a methyl group and alkoxy side chains bearing alkene or alkyne moieties afforded smoothly diazocines 2k–n. Notably, the autocondensation of alkene- or alkyne-derived aminophenones had to be performed at a lower temperature (80 °C) to maintain the high yield. The application of enantiomerically pure aminophenones has met with partial success, leading cleanly to diazocine 2q. However, an almost equimolar mixture of diastereomers was detected by 19F NMR. Further studies revealed that less nucleophilic aminopyridine derivatives could also participate in the autocondensation to give diazocines 4a–c.
Scheme 2. 1,1,3,3-Tetramethylguanidine (TMG)-Catalyzed Solvent-Free Autocondensation of Aminophenones.
For 16 h at 120 °C.
For 48 h at 120 °C.
For 16 h at 80 °C.
Next, we examined whether a more challenging, slightly acidic difluoromethyl ketone,11 prone to undergoing enolization and subsequent aldol reaction, could be involved in the TMG-catalyzed autocondensation (TMG; pKa ≈ 15.2 in H2O).12 The incorporation of the CF2H group into organic molecules has received a great deal of attention in medicinal chemistry13 due to its ability to act as a lipophilic hydrogen bond donor modifying permeability, binding affinity, and bioavailability.14 The engagement in weak interactions offers an ideal platform for the construction of higher-order molecular scaffolds.15 To our delight, diazocine 2o was formed in 70% yield under basic conditions without competing side reactions. The unique, rigid V-shaped structure was further evidenced by X-ray analysis (Scheme 2, structure 2o) showing a perpendicular arrangement of the two aromatic rings, similar to Tröger’s base.16
The requirements for new building blocks in supramolecular chemistry include ready access to useful quantities of the compounds. Indeed, the autocondensation proved to be scalable, and there was no need for special equipment. Simply heating 1 g of aminophenones (∼5.0 mmol) in a 4 mL closed vial in the presence of a drop of TMG (20 mol %) cleanly furnished the respective products 2b, 2c, 2e, 2h, 2k, 2l, and 2n without any erosion in yield in comparison to a 0.5 mmol scale. The challenging CF2H-substituted compound also afforded derivative 2o in a high 70% yield, emphasizing the practical aspect of the developed method.
With excellent results in the autocondensation process, further investigations were directed to cross-condensation. A careful choice of aminophenones prompted by the different rates of autocondensation and a disparate polarity under chromatographic conditions enabled the isolation of a series of diazocines 5a–e. The key for successful cross-condensation was mixing aminopyridine 3 with a 2-fold molar excess of aminophenone 1.17 A chiral aminophenone bearing a p-menthyloxy group also afforded diazocine 5e in 56% yield, though as a mixture of diastereomers in a ratio close to 1:1 (Scheme 3). Unfortunately, the sterically encumbered methyl substituent located in the ortho position adjacent to the reactive amino group suppressed cross-condensation (Scheme 3, structure 5f).
Scheme 3. TMG-Catalyzed Solvent-Free Cross-condensation of Aminophenones.
For 48 h at 120 °C.
For 16 h at 120 °C.
The synthetic potential of the diazocine products was demonstrated by a series of well-established transformations proceeding without affecting the diazocine core (Scheme 4). First, we turned our attention to the pyridinium salt structural motif, which proved to be useful in many areas18 such as molecular recognition,19 catalysis,20 and medicinal chemistry.21 Gratifyingly, the treatment of diazocines 5a and 4b with MeI cleanly afforded salts 6a and 6b, respectively, without competitive opening of the oxygen bridge under the action of the strong alkylating agent. Moreover, the bisester function was used to surround the hydrophobic cavity of the V-shaped structure by hydrogen bond donors, useful in molecular recognition. The respective bisester 2h was easily converted into bisamide 8 using achiral or chiral amino alcohols through a TBD-catalyzed protocol. Finally, carbon–carbon multiple bonds could also play a role in further functionalization without affecting the diazocine scaffold in metal-catalyzed reactions. Thus, bisalkene 2l underwent a cross-metathesis reaction (CM) with acrylate, whereas bisalkyne 2k provided bis-1,2,3-triazole 10 in high yield under standard conditions.
Scheme 4. Postformation Modification of the Diazocine Core (TBD = 1,5,7-triazabicyclo[4.4.0]dec-5-ene).
Isolated yield after chromatography.
Isolated yield after precipitation from the reaction mixture.
With regard to future applications, the most appealing feature of these systems is their stability in the enantiomerically pure form. Indeed, our initial experiments enabled the separation of racemic 2a into single enantiomers on a preparative scale (Scheme 5). To assign the absolute configuration of the two enantiomers of 2a, electronic and vibrational circular dichroism (ECD and VCD, respectively) spectra were recorded in acetonitrile and then simulated using quantum chemical methods (DFT and TDDFT). These two chiroptical spectroscopies are very sensitive to any stereochemical changes of the chiral system because they rely on electronic and vibrational transitions spanning the entire UV–vis–mid-IR spectral range. Thus, their complementary combinations provide a holistic view of the properties of any chiral molecules, enabling the conclusive assignment of their absolute configuration in solution and also deeper insight into dynamic stereochemistry.22
Scheme 5. (A) Separation of Racemic (±)-2a, (B) Comparison of Experimental and Calculated ECD and UV Data, and (C) VCD and IR Spectra of Enantiomers of 2a.
The calculations of the ECD and UV spectra were carried out at the CAM-B3LYP/def2-TZVP/PCM/CH3CN level of theory, while the VCD and IR spectra were calculated at the ωB97X-D/6-311+G(d,p)/ PCM/CH3CN level. More experimental and calculation details are provided in the Supporting Information. The inset in part B shows the geometry of the calculated structure of (+)-2a.
The ECD and VCD spectra of (+)-2a and (−)-2a display a perfect mirror-image relationship, confirming the enantiomeric relationship of these two newly synthesized diazocines separated by HPLC, as well as their high optical purity (Scheme 5, part A). The determination of the absolute configuration was based on the comparison of experimental and computed ECD and VCD spectra for an arbitrarily chosen R,R-enantiomer of 2a. A conformational search using the MMFF94s force field within 10 kcal/mol followed by DFT geometry optimizations at the ωB97X-D/6-311+G(d,p)/PCM/CH3CN level of theory revealed only one stable conformation, indicating ipso facto the extremely high rigidity of the diazocine core. Moreover, the high configurational and conformational stability was also proved experimentally using variable-temperature ECD measurements by heating the decalin solutions of 2a to 180 °C (Figure S2). The very close similarity between experimental and calculated ECD and VCD spectra observed in Scheme 5 (parts B and C) led to the conclusion that the absolute configuration of (+)-2a is R,R, with S,S for (−)-2a (Figure S2). It should be noted that the similar Tröger’s base and its derivatives underwent racemization, especially in the presence of Brønsted of Lewis acids, which makes diazocine 2a the superior platform for further derivatization.
Conclusions
In conclusion, we have established a new base-catalyzed, solvent-free condensation of fluorinated o-aminophenones for the construction of epoxydibenzo[b,f][1,5]diazocines. This unprecedented approach offers easily scalable access to a broad range of diazocines bearing a unique V-shaped structure, which was confirmed by X-ray and chiroptical analysis. The rigid molecular architecture allowed the separation of racemic diazocines into single enantiomers and proved their configurational stability by ECD measurements up to 140 °C. The ability to create a hydrophobic cavity by dibezo[b,f][1,5]diazocines, closely resembling Tröger’s base, opens up a plethora of possible applications in the area of supramolecular chemistry, which is now an ongoing subject in our group.
Experimental Section
General Remarks
NMR spectra were recorded in CDCl3, DMSO-d6, or CD3OD solutions (unless indicated otherwise); chemical shifts are quoted on the δ scale, with the solvent signal as the internal standard (CDCl3, 1H NMR δ 7.26, 13C NMR δ 77.00; DMSO-d6, 1H NMR δ 2.50, 13C NMR δ 39.40; CD3OD, 1H NMR δ 3.31, 13C NMR δ 49.00). High-resolution mass spectra (HRMS) were recorded using an EI technique or electrospray ionization (Supporting Information). Column chromatography was performed on Merck silica gel 60 (230–400 mesh) or alumina oxide 90 active basic (0.063–0.200 mm, Merck) using a standard glass column or a CombiFlash EzPrep system. TLC was performed on aluminum sheets, Merck 60F 254, or aluminum oxide. Optical rotations were recorded on a Jasco P-2000 polarimeter. Melting points were determined on a hot-stage apparatus and are uncorrected. Anhydrous solvents were obtained by distillation over CaH2 (DCM) or Na/benzophenone (THF, hexane, and MTBE). Air sensitive reactions were performed in flame-dried glassware under an argon atmosphere. Organic extracts were dried, and solvents were evaporated on a rotary evaporator. Reagents were used as they were purchased unless otherwise indicated. Aminophenones were synthesized starting from o-nitroaldehyde by the addition of the CF3 anion/reduction of NO2/oxidation23 sequence (1b and 1h) or orthometalation protocol (1a,241c–g, 1i–k, 1o–q,251h,261s and 1t,241u,271x,28 and 3a–c(25)), according to the literature procedure (for the structure of o-aminophenones used in this study, see Figure S1).
Synthesis of Aminophenones by the Addition/Reduction/Oxidation Sequence
General Procedure for the Synthesis of Trifluoroethanol Derivatives via the Addition of Ruppert–Prakash Reagent to Aldehydes (GP1)
To a cooled solution of aldehyde (the temperature of the cooling bath was kept in the range from −20 to −10 °C; the exact temperature is given in each case) in anhydrous THF was added TMSCF3 (1.2 equiv). Then a catalytic amount of a solution of TBAF (1 mol %) in THF (1 M) was added dropwise (Caution! In some cases, strong exothermic reaction was observed); the cooling bath was removed, and the resulting mixture was stirred for 16 h at rt (TLC analysis usually indicated the presence of silyl ether). Then a solution of TBAF (usually 0.1 mL/mmol of starting aldehyde) and water (usually 0.1 mL/mmol of starting aldehyde) were added, and the mixture was stirred until silyl ether hydrolysis occurred. The reaction mixture was evaporated and redissolved in EtOAc. The solution was washed with water (twice) and brine (twice), dried over Na2SO4, and evaporated. The residue was chromatographed on silica to give pure trifluoroethanol derivatives.
General Procedure for the Oxidation of Trifluoroethanol Derivatives to o-Aminophenones (GP2)
To a three-necked round-bottom flask was added anhydrous toluene followed sequentially by CuCl (5 mol %) and 1,10-phenanthroline (5 mol %). The black complex was immediately formed, and the resulting suspension was stirred at rt for 10 min. Then diethyl hydrazinodicarboxylate (DEAD-H2, 495.6 mg, 1.08 mmol) was added followed by solid K2CO3 (2.0 equiv), and the mixture was stirred for an additional 5 min. Then alcohol 12 (2.78 g, 11.3 mmol) was added (as a solid in one portion), and the solution was heated at 90 °C (temperature of the oil bath) for 1 h. To secure the maximum conversion, O2 was bubbled through the solution for 1 h (Caution! Special care should be taken due to low flash point of toluene, 4.4 °C). Then the reaction mixture was allowed to cool to rt and filtered through a pad of Celite. The filtrate was concentrated in vacuo and chromatographed on silica to give o-aminophenones (in some cases, fluorinated ketones were further purified by crystallization).
2,2,2-Trifluoro-1-[2-nitro-5-(prop-2-en-1-yloxy)phenyl]ethanol (11)
The title compound was obtained according to GP1 using 2-nitro-5-(prop-2-enyl-1-oxy)benzaldehyde29 (4.76 g, 23.0 mmol), anhydrous THF (50 mL), TMSCF3 (4.1 mL, 27.6 mmol, 1.2 equiv), and TBAF (230 μL, 0.23 mmol, 1 mol %, 1 M in THF). TBAF was added at −10 °C, and the reaction mixture was stirred for 3 h at rt. Then TBAF (2 mL) and water (2 mL) were added. After 16 h, THF was evaporated and the residue was dissolved in EtOAc (50 mL), washed with water (2 × 30 mL) and brine (2 × 30 mL), dried over Na2SO4, and evaporated. The residue was chromatographed on silica (10–15% EtOAc/hexanes) to give an orange oil (5.93 g, 92%): 1H NMR (200 MHz, CDCl3) δ 8.10 (d, J = 9.2 Hz, 1H), 7.46–7.39 (m, 1H), 6.99 (dd, J = 9.2, 2.8 Hz, 1H), 6.33 (q, J = 6.1 Hz, 1H), 6.15–5.90 (m, 1H), 5.54–5.25 (m, 2H), 4.71–4.60 (m, 2H), 3.34 (br s, 1H, OH); 13C{1H} NMR (50 MHz, CDCl3) δ 162.6, 141.1, 132.1 (q, JCF = 1.0 Hz), 131.6, 127.9, 123.8 (q, JCF = 281.2 Hz), 118.9, 115.5, 115.1, 69.5, 66.8 (q, JCF = 32.2 Hz); 19F NMR (376 MHz, CDCl3) δ −77.3; HRMS calcd for C11H9F3NO4 (ESI) m/z [M – H]− 276.0484, found 276.0482.
1-[2-Amino-5-(prop-2-en-1-yloxy)phenyl]-2,2,2-trifluoroethanol (12)
To a solution of nitroalcohol 11 (5.75 g, 20.7 mmol) in THF (80 mL) was added a saturated solution of NH4Cl (80 mL). The resulting biphasic mixture was cooled to 0 °C, and zinc powder (8.14 g, 124.4 mmol, 6.0 equiv, Sigma-Aldrich, 10 μm) was added in a few portions. Then the reaction mixture was vigorously stirred at rt for 24 h, and the resulting suspension was filtered (washing with 1 × 50 mL of EtOAc). Then the aqueous phase was separated, saturated with solid NaCl, and extracted with EtOAc (4 × 50 mL). The combined organic extracts were dried over Na2SO4 and evaporated, and the residue was chromatographed on silica (10–25% EtOAc/hexanes) to give a pure aniline 12 as a light yellow solid (2.62 g). Fractions containing impurities were collected and crystallized from n-heptane to give additional (0.60 g) white solid (3.37 g, overall yield of 66%): mp 111–112 °C (DCM/n-heptane); 1H NMR (400 MHz, CDCl3) δ 6.81 (br s, 3H), 6.09–5.96 (m, 1H), 5.43–5.35 (m, 1H), 5.31–5.25 (m, 1H), 5.00 (q, J = 7.5 Hz, 1H), 4.50–4.45 (m, 2H), 4.34 (br s, 2H); 13C{1H} NMR (100 MHz, CDCl3) δ 153.5, 136.8, 133.4, 125.3 (q, JCF = 281.2 Hz), 123.5, 122.2, 117.8, 116.7, 116.1, 72.4 (q, JCF = 32.0 Hz), 69.5; 19F NMR (376 MHz, CDCl3) δ −77.6; HRMS (ESI) m/z calcd for C11H13F3NO2 [M + H]+ 248.0898, found 248.0895.
1-[2-Amino-5-(prop-2-en-1-yloxy)phenyl]-2,2,2-trifluoroethanone (1l)
The title compound was obtained according to GP2 using toluene (56 mL), CuCl (55.7 mg, 0.56 mmol, 5 mol %), 1,10-phenanthroline (111.5 mg, 0.56 mmol, 5 mol %), DEAD-H2 (495.6 mg, 1.08 mmol), K2CO3 (2.81 g, 22.5 mmol, 2.0 equiv), and alcohol 12 (2.78 g, 11.3 mmol). Then the reaction mixture was allowed to cool to rt and filtered through a pad of Celite (washing with toluene). The filtrate was concentrated in vacuo and chromatographed on silica (20–50% DCM/hexanes) to give an orange solid (2.42 g, 88%): mp 76–77 °C (DCM/n-heptane); 1H NMR (400 MHz, CDCl3) δ 7.23–7.19 (m, 1H), 7.12 (dd, J = 9.1, 2.8 Hz, 1H), 6.73 (d, J = 9.1, 1H), 6.35 (br s, 2H, NH2), 6.09–5.97 (m, 1H, CH=CH2), 5.45–5.37 (m, 1H, CH=CH2), 5.33–5.27 (m, 1H, CH=CH2), 4.52–4.44 (m, 2H, OCH2); 13C{1H} NMR (50 MHz, CDCl3) δ 180.2 (q, JCF = 33.0 Hz), 149.0, 148.9, 133.2, 128.2, 119.0, 118.2, 117.2 (q, JCF = 289.9 Hz), 113.2 (q, JCF = 4.2 Hz), 110.4, 69.9; 19F NMR (376 MHz, CDCl3) δ −69.9; HRMS (ESI) m/z calcd for C11H11F3NO2 [M + H]+ 246.0742, found 246.0736.
5-(But-3-en-1-yloxy)-2-nitrobenzaldehyde (13)
To the solution of 5-hydroxy-2-nitrobenzaldehyde (4.25 g, 25.4 mmol) in anhydrous THF (230 mL), precooled to −10 °C, were added PPh3 (8.00 g, 30.5 mmol) and 3-buten-1-ol (2.55 mL, 30.5 mmol) in one portion. Then diisopropyl azodicarboxylate (6.0 mL, 30.52 mmol) was added dropwise over 5 min; the cooling bath was removed, and stirring was continued at rt for 22 h. The resulting solution was concentrated in vacuo, and the residue was chromatographed on silica (15–30% EtOAc/hexanes, Combi Flash) to give a yellowish oil (3.16 g, 56%): 1H NMR (400 MHz, CDCl3) δ 10.48 (s, 1H, CHO), 8.15 (d, J = 9.0 Hz, 1H), 7.32 (d, J = 2.8 Hz, 1H), 7.14 (dd, J = 9.0, 2.9 Hz, 1H), 5.93–5.81 (m, 1H, CH=CH2), 5.23–5.12 (m, 2H, CH=CH2), 4.16 (t, J = 6.6 Hz, 2H, OCH2), 2.63–2.55 (m, 2H, CH2CH=CH2); 13C{1H} NMR (100 MHz, CDCl3) δ 188.6, 163.5, 142.3, 134.5, 133.4, 127.3, 119.0, 118.0, 113.9, 68.6, 33.3; HRMS (EI) m/z calcd for C11H11NO4 [M]•+ 221.0688, found 221.0694.
1-[5-(But-3-en-1-yloxy)-2-nitrophenyl]-2,2,2-trifluoroethanol (14)
The title compound was obtained according to GP1 using aldehyde 13 (3.13 g, 15.2 mmol), anhydrous THF (50 mL), TMSCF3 (2.5 mL, 17.0 mmol, 1.2 equiv), and TBAF (150 μL, 0.15 mmol, 1 mol %, 1 M in THF). The TBAF solution was added at −10 °C, and the reaction mixture was stirred for 3 h at rt. Then TBAF (1.5 mL, 1 M in THF) and water (1.5 mL) were added to deprotect silyl ether. After 16 h, THF was evaporated and the residue was dissolved in EtOAc (50 mL), washed with water (2 × 30 mL) and brine (2 × 30 mL), dried over Na2SO4, and evaporated. The residue was chromatographed on silica (10–15% EtOAc/hexanes) to give an orange oil (2.79 g, 63%): 1H NMR (400 MHz, CDCl3) δ 8.10 (d, J = 9.2 Hz, 1H), 7.40 (d, J = 2.8 Hz, 1H), 6.97 (dd, J = 9.2, 2.8 Hz, 1H), 6.36–6.28 (m, 1H), 5.95–5.82 (m, 1H), 5.23–5.11 (m, 2H), 4.13 (t, J = 6.6 Hz, 2H), 3.26 (d, J = 5.5 Hz, 1H), 2.62–2.54 (m, 2H); 13C{1H} NMR (100 MHz, CDCl3) δ 163.1, 141.1, 133.5, 132.1, 127.9, 123.0 (q, J = 281.4 Hz), 117.7, 115.2, 114.9, 68.1, 67.0 (q, J = 32.4 Hz), 33.2; 19F NMR (376 MHz, CDCl3) δ −77.3; HRMS (ESI) m/z calcd for C12H11F3NO4 [M – H]− 290.0640, found 290.0631.
1-[2-Amino-5-(but-3-en-1-yloxy)phenyl]-2,2,2-trifluoroethanol (15)
To a solution of nitroalcohol 14 (2.73 g, 9.37 mmol) in THF (20 mL) was added a saturated solution of NH4Cl (20 mL); the mixture was cooled to 0 °C, and zinc powder (3.67 g, 56.2 mmol, 6.0 equiv, Sigma-Aldrich, 10 μm) was added in a few portions. Then the reaction mixture was vigorously stirred at rt for 16 h and filtered through pad of Celite (washing with EtOAc). The resulting mixture was diluted with brine (30 mL) and extracted with EtOAc (3 × 20 mL). The combined organic extracts were washed with brine (2 × 20 mL), dried over Na2SO4, and evaporated. The residue was chromatographed on silica (15–50% EtOAc/hexanes, Combi Flash, 80 g column) to give a light yellow solid (1.92 g, 78%): 1H NMR (400 MHz, CDCl3) δ 6.81–6.72 (m, 3H), 5.95–5.82 (m, 1H), 5.20–5.07 (m, 2H), 4.97 (q, J = 7.5 Hz, 1H), 4.26 (s, 3H), 3.95 (t, J = 6.7 Hz, 2H), 2.56–2.46 (m, 2H); 13C{1H} NMR (100 MHz, CDCl3) δ 153.8, 136.2, 134.4, 125.1 (q, JCF = 283.2 Hz), 123.6 (q, JCF = 1.4 Hz), 122.4, 117.0, 116.4, 115.9, 72.4 (q, JCF = 32.1 Hz), 67.8, 33.6; 19F NMR (376 MHz, CDCl3) δ −77.6; HRMS (ESI) m/z calcd for C12H14F3NO2 [M + H]+ 262.1055, found 262.1042.
1-[2-Amino-5-(but-3-en-1-yloxy)phenyl]-2,2,2-trifluoroethanol (1m)
The title compound was obtained according to GP2 using anhydrous toluene (40 mL), CuCl (36.2 mg, 0.366 mmol, 5 mol %), 1,10-phenanthroline (72.6 mg, 0.366 mmol, 5 mol %), DEAD-H2 (322.0 mg, 1.83 mmol), solid K2CO3 (2.02 g, 14.62 mmol, 2.0 equiv), and alcohol 15 (1.91 g, 7.31 mmol). Then the reaction mixture was allowed to cool to rt, filtered through a pad of Celite, and washed with EtOAc. The filtrate was concentrated in vacuo and chromatographed on silica (10% EtOAc/hexanes, CombiFlash) to give an orange solid (1.62 g, 86%): mp 71–72 °C (DCM/n-heptane); 1H NMR (400 MHz, CDCl3) δ 7.20–7.15 (m, 1H), 7.10 (dd, J = 9.1, 2.8 Hz, 1H), 6.68 (d, J = 9.1 Hz, 1H), 6.23 (br s, 2H), 5.96–5.83 (m, 1H), 5.22–5.08 (m, 2H), 3.96 (t, J = 6.6 Hz, 2H), 2.56–2.48 (m, 2H); 13C{1H} NMR (100 MHz, CDCl3) δ 180.1 (q, JCF = 33.1 Hz), 149.3, 148.7, 134.3, 128.0, 118.9, 117.1 (d, JCF = 291.4 Hz), 117.1, 112.9 (q, JCF = 4.3 Hz), 110.5, 68.2, 33.6; 19F NMR (376 MHz, CDCl3) δ −69.9; HRMS (ESI) m/z calcd for C12H13F3NO2 [M + H]+ 260.0898, found 260.0892.
2-Nitro-5-(pent-4-en-1-yloxy)benzaldehyde (16)
To a solution of 5-hydroxy-2-nitrobenzaldehyde (6.0 g, 35.9 mmol) and Ph3P (11.30 g, 43.1 mmol) in THF (200 mL), cooled to −20 °C, was added 4-penten-1-ol (4.42 mL, 43.1 mmol) in one portion. Then diisopropyl azodicarboxylate (6.0 mL, 30.5 mmol) was added dropwise over 10 min. Stirring was continued for 1 h at −20 °C; the cooling bath was removed, and the resulting solution was stirred at rt for an additional 19 h. Then the reaction mixture was concentrated in vacuo, and the residue was chromatographed on silica (5–10% EtOAc/hexanes, Combi Flash) to give a yellowish oil (4.67 g, 54%): 1H NMR (400 MHz, CDCl3) δ 10.47–10.46 (m, 1H, CHO), 8.13 (dd, J = 9.1, 0.9 Hz, 1H), 7.30 (dd, J = 2.8, 0.8 Hz, 1H), 7.13 (dd, J = 9.0, 2.9 Hz, 1H), 5.89–5.77 (m, 1H, CH=CH2), 5.10–4.99 (m, 2H, CH=CH2), 4.11 (t, J = 6.4 Hz, 2H, OCH2), 2.28–2.21 (m, 2H, CH2CH=CH2), 1.97–1.89 (m, 2H, CH2CH2CH2); 13C{1H} NMR (100 MHz, CDCl3) δ 188.7, 163.7, 142.2, 137.2, 134.5, 127.4, 119.0, 115.9, 113.9, 68.6, 29.9, 28.1; HRMS (ESI) m/z calcd for C12H12NO4 [M – H]− 234.0766, found 234.0763.
2,2,2-Trifluoro-1-[2-nitro-5-(pent-4-en-1-yloxy)phenyl]ethanol (17)
The title compound was obtained according to GP1 using aldehyde 16 (4.52 g, 19.2 mmol), anhydrous THF (50 mL), TMSCF3 (3.4 mL, 23.1 mmol, 1.2 equiv), and TBAF (192 μL, 0.19 mmol, 1 mol %, 1 M in THF). TBAF was added at −10 °C, and the reaction mixture was stirred for 3 h at rt (TLC analysis indicated the absence of substrate). Then a TBAF solution (1.5 mL, 1 M in THF) and water (1.5 mL) were added to deprotect silyl ether. After 16 h, the reaction mixture was evaporated and the residue was dissolved in EtOAc (50 mL), washed with water (2 × 30 mL) and brine (2 × 30 mL), dried over Na2SO4, and evaporated. The residue was used in the next step without further purification.
1-[2-Amino-5-(pent-4-en-1-yloxy)phenyl]-2,2,2-trifluoroethanol (18)
To a solution of nitroalcohol 17 (5.86 g, 19.2 mmol) in THF (40 mL) was added a saturated solution of NH4Cl (40 mL); the mixture was cooled to 0 °C, and zinc powder (7.53 g, 115.2 mmol, 6.0 equiv, Sigma-Aldrich, 10 μm) was added in a few portions. Then the reaction mixture was stirred at rt for 16 h and filtered through a pad of Celite (washing with EtOAc). The resulting mixture was diluted with brine (30 mL) and extracted with EtOAc (3 × 20 mL). The combined organic extracts were washed with brine (2 × 20 mL), dried over Na2SO4, and evaporated. The residue was chromatographed on silica (10–25% EtOAc/hexanes) to give a light yellow solid (4.16 g). Fractions including some impurities were collected and crystallized from n-heptane to obtain additional (0.41 g) aniline 18 (4.57 g, overall yield of 86%): mp >87 °C dec (precipitation from a DCM solution with n-heptane); 1H NMR (400 MHz, CDCl3) δ 6.82–6.75 (m, 3H), 5.91–5.79 (m, 1H, CH=CH2), 5.10–4.95 (m, 2H, HOCHCF3, CH=CH2), 4.22 (br s, 3H, NH2 and OH), 3.92 (t, J = 6.4 Hz, 2H, OCH2), 2.27–2.18 (m, 2H, CH2CH=CH2), 1.90–1.81 (m, 2H, CH2CH2CH2); 13C{1H} NMR (100 MHz, CDCl3) δ 154.1, 137.9, 136.4, 125.5 (JCF = 281.5 Hz), 123.7, 122.5, 116.5, 115.9, 115.3, 72.5 (JCF = 32.0 Hz), 67.9, 30.2, 28.6; 19F NMR (376 MHz, CDCl3) δ −77.6; HRMS (ESI) m/z calcd for C13H17F3NO2 [M + H]+ 276.1211, found 276.1203.
1-[2-Amino-5-(pent-4-en-1-yloxy)phenyl]-2,2,2-trifluoroethanone (1n)
The title compound was obtained according to GP2 using anhydrous toluene (82 mL), CuCl (81.4 mg, 0.82 mmol, 5 mol %), 1,10-phenanthroline (162.9 mg, 0.82 mmol, 5 mol %), DEAD-H2 (724.0 mg, 4.11 mmol), solid K2CO3 (4.54 g, 32.9 mmol, 2.0 equiv), and alcohol 18 (4.53 g, 16.4 mmol). Then the reaction mixture was allowed to cool to rt and filtered through a pad of Celite (washing with toluene). The filtrate was concentrated in vacuo and chromatographed on silica (25–35% DCM/hexanes) to give an orange solid (4.06 g, 90%): mp 78–79 °C (DCM/n-heptane); 1H NMR (400 MHz, CDCl3) δ 7.16 (br s, 1H), 7.13–7.06 (m, 1H), 6.68 (d, J = 9.1 Hz, 1H), 6.23 (br s 2H, NH2), 5.92–5.78 (m, 1H, CH=CH2), 5.11–4.97 (m, 2H, CH=CH2), 3.92 (t, J = 6.4 Hz, 2H, OCH2), 2.28–2.19 (m, 2H, CH2CH=CH2), 1.92–1.82 (m, 2H, CH2CH2CH2); 13C{1H} NMR (100 MHz, CDCl3) δ 180.3 (JCF = 32.8 Hz), 149.6, 148.7, 137.8, 128.1, 119.0, 117.2 (JCF = 289.8 Hz), 115.4, 112.9 (JCF = 4.8 Hz), 110.7, 68.2, 30.2, 28.5; 19F NMR (376 MHz, CDCl3) δ −69.8; HRMS (ESI) m/z calcd for C13H15F3NO2 [M + H]+ 274.1055, found 274.1046.
2-Nitro-5-(prop-2-yn-1-yloxy)benzaldehyde (19)
To a solution of 5-hydroxy-2-nitrobenzaldehyde (8.0 g, 47.9 mmol) in DMF (75 mL) were added K2CO3 (7.28 g, 52.6 mmol), 3-bromo-1-propyne (4.3 mL, 57.4 mmol), and TBAB (154.3 mg, 0.48 mmol, 1 mol %). The resulting suspension was stirred for 22 h, and the solvent was evaporated. The residue was partitioned between a saturated aqueous solution of Na2CO3 (50 mL) and EtOAc (50 mL). The aqueous phases was separated and extracted with EtOAc (4 × 50 mL). The combined organic phases were dried over solid K2CO3 and evaporated. The residue was chromatographed on silica (25% DCM/hexanes) to give a yellow oil (7.29 g, 74%): 1H NMR (400 MHz, CDCl3) δ 10.46 (s, 1H, CHO), 8.16 (d, J = 9.0 Hz, 1H), 7.41 (d, J = 3.0 Hz, 1H), 7.24 (dd, J = 9.0, 2.9 Hz, 1H), 4.84 (d, J = 2.4 Hz, 2H, CH2), 2.59 (t, J = 2.4 Hz. 1H). Spectral data are in agreement with those reported previously.30
2,2,2-Trifluoro-1-[2-nitro-5-(prop-2-yn-1-yloxy)phenyl]ethanol (20)
The title compound was obtained according to GP1 using aldehyde 19 (7.26 g, 35.4 mmol), anhydrous THF (70 mL), TMSCF3 (6.8 mL, 46.04 mmol, 1.3 equiv), and TBAF (354 μL, 0.35 mmol, 1 mol %, 1 M in THF). A TBAF solution was added at −10 °C over 15 min (Caution! A strong exothermic effect was observed), and the reaction mixture was stirred for 17.5 h at rt. Then a TBAF solution (2 mL, 1 M in THF) and water (2 mL) were added. After 4 h, the reaction was quenched with a saturated aqueous solution of NH4Cl (50 mL). The aqueous phase was separated and extracted with EtOAc (3 × 50 mL). The combined extracts were washed with brine (1 × 100 mL), dried over Na2SO4, and evaporated. The residue was chromatographed on silica (50–75% DCM/hexanes) to give 20 as a white solid (3.51 g, 36%): mp 78–79 °C (DCM/n-heptane); 1H NMR (400 MHz, CD3OD) δ 8.09 (d, J = 9.2 Hz, 1H), 7.54 (d, J = 2.8 Hz, 1H), 7.17 (dd, J = 9.3, 2.9 Hz, 1H), 4.89 (d, J = 2.5 Hz, 2H, CH2), 3.02 (t, J = 2.5 Hz, C≡CH); 13C{1H} NMR (100 MHz, CD3OD) δ 162.7, 143.3, 134.5, 128.4, 125.8 (q, JCF = 280.7 Hz), 116.8, 116.4, 78.4, 77.9, 69.0 (q, JCF = 31.8 Hz), 57.3; 19F NMR (376 MHz, CD3OD) δ −79.1; HRMS (ESI) m/z calcd for C11H8F3NO4 [M – H]− 274.0327, found 274.0322.
1-[2-Amino-5-(prop-2-yn-1-yloxy)phenyl]-2,2,2-trifluoroethanol (21)
To a solution of nitroalcohol 20 (3.23 g, 11.7 mmol) in THF (50 mL) was added a saturated solution of NH4Cl (50 mL), and then zinc powder (4.61 g, 70.4 mmol, 6.0 equiv, Sigma-Aldrich, 10 μm) was added in a few portions (a slight increase in the temperature of the reaction mixture was detected). After 18 h, the reaction mixture was filtered through a pad of Celite, and the aqueous phase was separated, saturated with solid NaCl, and extracted with EtOAc (4 × 50 mL). The combined organic extracts were dried over Na2SO4 and evaporated. The residue was filtered through a pad of silica (25% EtOAc/hexanes) and then crystallized form n-heptane to give 21 as a light yellow solid (1.29 g, 45%): mp 82–83 °C (n-heptane); 1H NMR (400 MHz, CDCl3) δ 6.89 (br s, 3H), 5.09–5.01 (m, 1H, CHOH), 4.81 (br s, 3H, NH2 and OH), 4.64 (d, J = 2.4 Hz, 2H, CH2), 2.51 (t, J = 2.4 Hz, 1H, C≡CH); 13C{1H} NMR (100 MHz, CDCl3) δ 152.2, 137.5, 125.2 (JCF = 281.3 Hz), 123.1, 121.8, 117.1, 116.6, 78.7, 75.8, 72.2 (JCF = 32.1 Hz), 56.7; 19F NMR (376 MHz, CDCl3) δ −77.6; HRMS (ESI) m/z calcd for C11H10F3NO2 [M + H]+ 246.0742, found 246.0732.
1-[2-Amino-5-(prop-2-yn-1-yloxy)phenyl]-2,2,2-trifluoroethanone (1k)
The title compound was obtained according to modified GP2 using anhydrous toluene (22 mL), CuCl (21.3 mg, 0.216 mmol, 5 mol %), 1,10-phenanthroline (42.7 mg, 0.216 mmol, 5 mol %), DEAD-H2 (189.8 mg, 1.08 mmol), and solid K2CO3 (2.19 g, 8.62 mmol, 2.0 equiv). Then alcohol 21 (1.06 g, 4.31 mmol) was added (as a solid in one portion), and the solution was heated at 90 °C (temperature of the oil bath) for 40 h. To secure the maximum conversion, O2 was slowly bubbled through the solution for 40 h. Then the reaction mixture was allowed to cool to rt and filtered through a pad of Celite (washing with toluene). The filtrate was concentrated in vacuo, and the residue was chromatographed on silica (15% DCM/hexanes) to give an orange solid (348.7 mg, 33%): mp 56–58 °C (DCM/n-heptane); 1H NMR (400 MHz, CDCl3) δ 7.37–7.32 (m, 1H), 7.16 (dd, J = 9.1, 2.8 Hz, 1H), 6.75–6.67 (m, 1H), 6.29 (br s, 2H, NH2), 4.64 (d, J = 2.4 Hz, 2H, CH2), 2.53 (t, 2.4 Hz, 1H, C≡CH); 13C{1H} NMR (100 MHz, CDCl3) δ 180.3 (q, JCF = 33.2 Hz), 149.3, 147.9, 128.3, 119.0, 117.2 (q, JCF = 289.9 Hz), 114.6 (q, JCF = 4.2 Hz), 110.5, 78.4, 76.1, 57.2; 19F NMR (376 MHz, CDCl3) δ −69.9; HRMS (ESI) m/z calcd for C11H9F3NO2 [M + H]+ 244.0585, found 244.0574.
2-{[(1R,2S,5R)-5-Methyl-2-(propan-2-yl)cyclohexyl]oxy}aniline (23)
To a solution of (1R,2S,5R)-5-methyl-2-(propan-2-yl)cyclohexyl 2-nitrophenyl ether3122 (5.65 g, 20.4 mmol) in EtOAc (100 mL) was added 10% Pd/C (1.05 g, 1.0 mmol, 5 mol %), and the resulting suspension was shaken in a Parr apparatus under an atmosphere of H2 (3 bar; Caution! Exothermic reaction!). After 3 h, the reduction was complete (as judged by TLC), and a gentle stream of argon was passed through the solution for 5 min. The resulting suspension was filtered through a pad of Celite, and the solvent was evaporated to give a colorless oil (5.05 g). Crude 2-{[(1R,2S,5R)-5-methyl-2-(propan-2-yl)cyclohexyl]oxy}aniline 23 was used in the next step without further purification.
2,2-Dimethyl-N-(2-{[(1R,2S,5R)-5-methyl-2-(propan-2-yl)cyclohexyl]oxy}phenyl)propanamide (24)
To a solution of aniline 23 (5.05 g, 20.4 mmol) and Et3N (3.3 mL, 24.6 mmol, 1.2 equiv) in anhydrous DCM (100 mL), cooled to 0 °C, was added dropwise PivCl (2.8 mL, 22.5 mmol, 1.1. equiv); the cooling bath was removed, and the reaction mixture was stirred at rt. After 16 h, the reaction mixture was washed with 10% aqueous citric acid (2 × 30 mL) and a saturated solution of Na2CO3 (2 × 30 mL), dried over MgSO4, and evaporated. The residue was chromatographed on silica (5% EtOAc/hexanes) to give a colorless oil (5.47 g, 91%): [α]D23 = −91.0 (CHCl3, c = 1.0); 1H NMR (400 MHz, CDCl3) δ 8.41 (dd, J = 7.9, 1.7 Hz, 1H), 8.26 (br s, 1H), 7.02–6.85 (m, 3H), 4.18–4.08 (m, 1H), 2.27–2.12 (m, 2H), 1.83–1.70 (m, 2H), 1.57–1.40 (m, 2H), 1.38–1.09 (m, 2H) overlapping 1.31 (s, 9H), 1.02–0.88 (m, 1H) overlapping 0.95 (d, J = 7.1 Hz, 3H) and 0.92 (d, J = 6.6 Hz, 3H), 0.80 (d, J = 6.9 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3) δ 176.2, 146.4, 129.0, 123.2, 120.8, 119.6, 111.9, 78.4, 48.6, 40.6, 40.0, 34.4, 31.4, 27.6, 26.5, 23.8, 22.0, 20.7, 16.9; HRMS (ESI) m/z calcd for C21H33N2ONa [M + H]+ 354.2409, found 354.2403.
1-(2-Amino-3-{[(1R,2S,5R)-5-methyl-2-(propan-2-yl)cyclohexyl]oxyphenyl)-2,2,2-trifluoroethanone (1r)
To a solution of 2,2-dimethyl-N-(2-{[(1R,2S,5R)-5-methyl-2-(propan-2-yl)cyclohexyl]oxy}phenyl)propanamide (24) (5.60 g, 17 mmol) and TMEDA (5.14 mL, 37.4 mmol, 2.2 equiv) in anhydrous THF (40 mL) was added dropwise n-BuLi (16.4 mL, 37.4 mmol, 2.2 equiv, 2.28 M in hexane) with a syringe pump within 40 min. Then the cooling bath was removed, and the reaction mixture was stirred for 4 h at 20 °C. The resulting yellow suspension was cooled again to −30 °C, and CF3CO2Et (2.86 mL, 23.8 mmol, 1.4 equiv) was added. Stirring was continued for 1.5 h at rt with 4 M HCl with dioxane (30 mL) and water (3 mL). The biphasic mixture was heated at 90 °C for 4 h, cooled to rt, and neutralized with a saturated solution of Na2CO3. Then the reaction mixture was extracted with EtOAc (3 × 30 mL), and the combined organic extracts were dried over MgSO4 and evaporated. The residue was chromatographed on silica (2.5% EtOAc/hexanes) to give ketone 1r as a red-brown oil (1.27 g, 22%): [α]D23 = −83.5 (CHCl3, c = 0.4); IR (film) 3506, 3371, 2957, 2928, 2871, 1667, 1619, 1580, 1545, 1454 cm–1; 1H NMR (400 MHz, CDCl3) δ 7.37–7.31 (m, 1H), 6.92 (d, J = 7.6 Hz, 1H) overlapping 6.85 (br s, 2H, NH2), 4.16–4.07 (m, 1H), 2.24–2.10 (m, 2H), 1.82–1.70 (m, 2H), 1.64–1.54 (m, 1H), 1.54–1.40 (m, 1H), 1.38–0.99 (m, 3H), 0.99–0.85 (m, 7H), 0.79 (d, J = 6.96 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3) δ 180.7 (q, JCF= 33.0 Hz), 145.9, 145.2, 121.9 (q, JCF= 4.1 Hz), 117.1 (q, JCF= 289.7 Hz), 115.9, 114.7, 110.6, 78.4, 48.0, 40.0, 34.8, 31.4, 26.3, 23.7, 22.0, 20.7, 16.7; 19F NMR (376 MHz, CDCl3) δ −69.7; HRMS (ESI) m/z calcd for C18H24F3N2O [M + H]+ 344.1837, found 344.1834.
2,2,2-Trifluoro-1-(6-nitro-1,3-benzodioxol-5-yl)ethanol (25)
The title compound was obtained according to GP1 using 6-nitro-1,3-benzodioxole-5-carbaldehyde (26) (2.0 g, 10.3 mmol), anhydrous THF (20 mL), TMSCF3 (2.0 mL, 13.3 mmol, 1.3 equiv), and TBAF (0.1 mL, 0.1 mmol, 1 mol %). A TBAF solution was added at −20 °C, and the reaction mixture was stirred for 4 h at rt. Then the reaction mixture was cooled to 0 °C, and water (1 mL) and a TBAF solution (1 mL, 1 M in THF) were added. After 1 h, the reaction mixture was evaporated, and the residue was dissolved in MTBE (20 mL), washed with water (2 × 20 mL) and brine (2 × 20 mL), dried over Na2SO4, and evaporated. The residue was chromatographed on silica (15% EtOAc/hexanes) to give a yellow oil (2.71 g, 99%): 1H NMR (400 MHz, CDCl3) δ 7.52 (s, 1H), 7.33 (s, 1H), 6.20 (q, J = 6.7 Hz, 2H), 6.17–6.14 (m, 2H), 3.20 (br s, 1H, OH); 19F NMR (376 MHz, CDCl3) δ −77.5. Spectroscopic data are in agreement with those reported previously.32
1-(6-Amino-1,3-benzodioxol-5-yl)-2,2,2-trifluoroethanol (27)
To a solution of nitroalcohol 25 (2.71 g, 10.2 mmol) in reagent-grade EtOAc (100 mL) was added 10% Pd/C (543.9 mg, 0.51 mmol, 5 mol %), and the resulting mixture was shaken in a Parr apparatus under an atmosphere of H2 for 2 h (3 bar; Caution! In some cases, reduction of the nitro group has appeared to be strongly exothermic, and caution should be taken; the temperature of the reaction mixture increased from 17 to 26 °C within 10 min). Then a gentle stream of argon was bubbled through the solution for 5 min. The resulting suspension was filtered through a pad of Celite (washing with 50% EtOAc/hexanes), and solvents were evaporated to give a light-yellow solid (2.21 g, 92%). Crude aminoalcohol 28 was used in the next step without further purification.
1-(6-Amino-1,3-benzodioxol-5-yl)-2,2,2-trifluoroethanone (28)
The title compound was obtained according to GP2 using anhydrous toluene (50 mL), CuCl (464.8 mg, 4.70 mmol, 0.5 equiv), 1,10-phenanthroline (186.1 mg, 1.03 mmol, 0.11 equiv), DEAD-H2 (413.6 mg, 2.35 mmol, 0.25 equiv), solid K2CO3 (2.59 g, 18.78 mmol, 2.0 equiv), and alcohol 27 (2.21 g, 9.39 mmol). Then the reaction mixture was allowed to cool to rt and filtered through a pad of Celite (washing with toluene). The filtrate was concentrated in vacuo and chromatographed on silica (120 g, 10–15% EtOAc/hex) to give an orange solid (1.59 g, 75%): mp 154–155 °C (n-heptane); IR (KBr) 3416, 3317, 3087, 3013, 2919, 1663, 1641, 1576 cm–1; 1H NMR (400 MHz, CDCl3) δ 7.07 (q, J = 2.1 Hz, 1H), 6.70 (br s, 2H), 6.17 (s, 1H), 5.95 (s, 2H); 13C{1H} NMR (100 MHz, CDCl3) δ 177.9 (q, JCF = 32.8 Hz), 155.4, 153.8, 139.6, 117.5 (q, JCF = 289.5 Hz), 106.8, 106.8, 103.9, 101.8, 96.2; 19F NMR (376 MHz, CDCl3) δ −69.7; HRMS (EI) m/z calcd for C9H6F3NO3 [M]•+ 233.0300, found 233.0293.
General Procedure for the Synthesis of Symmetric Epoxydibenzo[b,f][1,5]diazocines (GP3)
Fluoromethylketone 1 (x mmol) and N,N,N′,N′-tetramethylguanidine (TMG, 20 mol %) were placed in a screw-cap 4 mL vial, and the resulting mixture was heated at 120 °C (IKA heating block, temperature of the reference vial filled with silicon oil). Then the reaction mixture was diluted with EtOAc or DCM (10 mL), adsorbed on silica (or aluminum oxide), and chromatographed to give the corresponding dibenzo[b,f][1,5]diazocines 2. The analytical sample was crystallized from a given solvent to measure the melting point.
(6S*,12S*)-2,8-Dichloro-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2a)
The title compound was obtained according to GP3 using ketone 1a (1.0 g, 4.47 mmol) and TMG (111 μL, 0.89 mmol, 20 mol %). The crude product was chromatographed on silica (5% EtOAc/hexanes) to give a colorless solid (911.2 mg, 95%). Autocondensation of 1a was also performed on a 0.5 mmol scale to afford product 2a in 93% yield (99.8 mg) after chromatography during the optimization studies. It should be mentioned that all attempts to use the Dean–Stark apparatus to continuously remove water formed during the condensation have failed (aminophenone 1a sublimed in the condenser). Similarly, when an open round-bottom flask was used as the reaction vessel, aminophenone 1a also easily sublimed: IR (KBr) 3380, 3352, 3070, 3041, 2902, 1776, 1775, 1612, 1493 cm–1; 1H NMR (400 MHz, CDCl3) δ 7.43 (br s, 2H), 7.21 (dd, J = 8.6, 2.3 Hz, 2H), 6.80 (d, J = 8.6 Hz, 2H), 4.89 (br s, 2H, 2 × NH). Spectral data are in agreement with those reported in the literature.9e Representative chromatograms of the formation of diazocine 2a, catalyzed by TMG, are presented below.
(6S*,12S*)-2,8-Dibromo-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2b)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1b (134.0 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (10% EtOAc/hexanes) to give a light yellow oil that solidified upon standing in the refrigerator (120.5 mg, 93%). The reaction was also performed on a 3.7 mmol scale (1.0 g of o-TMFK 1b) to give the product in 89% yield (865.3 mg): IR (KBr) 3374, 3336, 1606, 1490 cm–1; 1H NMR (600 MHz, CD3OD) δ 7.51–7.49 (m, 2H), 7.30 (dd, J = 8.7, 2.2 Hz, 2H), 6.77 (d, J = 8.7 Hz, 2H); 13C{1H} NMR (150 MHz, CD3OD) δ 142.2, 134.1, 128.8 (q, JCF = 3.1 Hz), 124.7 (q, JCF = 281.5 Hz), 122.8, 120.4, 113.0, 83.9 (q, JCF = 32.4 Hz); 19F NMR (376 MHz, CD3OD) δ −80.7; HRMS (ESI) m/z calcd for C16H8N2OBr2F6 [M]•+ 515.8908, found 515.8907.
(6S*,12S*)-2,8-Difluoro-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2c)
The title compound was obtained according to GP3 using ketone 1c (103.6 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (5% EtOAc/hexane) to give a white solid (73.3 mg, 74%). The reaction was also performed on a 4.82 mmol scale (1.0 g of o-TMFK 1c) to give the product in 70% yield (670.3 mg): mp 169–170 °C (n-heptane); 1H NMR (400 MHz, CDCl3) δ 7.23–7.17 (m, 2H), 7.02–6.96 (m, 2H), 6.86 (dd, J = 8.9, 4.9 Hz, 2H), 4.77 (br s, 2H, 2 × NH); 19F NMR (376 MHz, CDCl3) δ −79.3, −118.8. The spectroscopic data are in agreement with those reported previously.25
(6S*,12S*)-6,12-Bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2d)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1d (95.0 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (30–50% toluene/EtOAc) to give a yellow solid (77.6 mg, 86%): mp 140–141 °C (n-heptane at −78 °C); IR (KBr) 3413, 3336, 3081, 3054, 1955, 1922, 1802, 1612, 1586, 1495 cm–1; 1H NMR (400 MHz, CDCl3) δ 7.50–7.42 (m, 2H), 7.24–7.18 (m, 2H), 6.98–6.91 (m, 2H), 6.85–6.79 (m, 2H), 4.90 (s, 2H, 2 × NH); 13C{1H} NMR (100 MHz, CDCl3) δ 140.0, 130.4, 125.4 (q, JCF = 3.0 Hz), 122.8 (q, JCF = 282.4 Hz), 122.0, 120.4, 119.1, 83.2 (q, JCF = 32.0 Hz); 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI-TOF) m/z calcd for C16H9F6N2O [M – H]− 359.0619, found 359.0628.
(6S*,12S*)-2,8-Dimethyl-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2e)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1e (203.2 mg, 0.5 mmol) and TMG (12.5 μL, 20 mol %, 0.1 mmol). The crude product was chromatographed on silica (50% hexane/toluene) to give an off-white solid (72.8 mg, 74%). The reaction was also performed on a 4.8 mmol scale (1.0 g of o-TMFK 1e) to give product 2e in 69% yield (659.3 mg): mp 164–165 °C (n-heptane, −20 °C); IR (KBr) 3361, 3308, 3029, 2930, 2868, 2742, 1621, 1585, 1507 cm–1; 1H NMR (400 MHz, CDCl3) δ 7.25 (br s, 2H), 7.03 (dd, J = 8.2, 1.5 Hz, 2H), 4.74 (d, J = 8.2 Hz, 2H), 4.74 (br s, 2H, NH), 2.25 (s, 6H, 2 × CH3); 13C{1H} NMR (100 MHz, CDCl3) δ 137.5, 131.8, 131.3, 125.7 (q, JCF= 2.9 Hz), 122.9 (q, JCF= 282.4 Hz), 120.5, 119.5, 83.5 (q, JCF= 31.7 Hz), 20.8 (CH3); 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI) m/z calcd for C18H15F6N2O [M + H]+ 389.1089, found 389.1086.
(6S*,12S*)-2,8-Dimethoxy-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2f)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1f (109.6 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (10–15% EtOAc/hexanes) to give a white solid (75.7 mg, 72%): mp 225–226 °C (DCM; by slow evaporation); IR (KBr) 3357, 3305, 3024, 2960, 2845, 1620, 1592, 1508 cm–1; 1H NMR (400 MHz, CDCl3) δ 6.99 (br s, 2H), 6.87–6.79 (m, 4H), 4.58 (br s, 2H, 2 × NH), 3.73 (s, 6H, 2 × OCH3); 13C{1H} NMR (100 MHz, CDCl3) δ 155.3, 133.2, 122.8 (q, JCF = 282.3 Hz), 121.8, 121.7, 117.2, 110.4 (q, JCF = 3.1 Hz), 83.9 (q, JCF = 31.5 Hz), 55.5 (OCH3); 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI) m/z calcd for C18H15F6N2O3 [M + H]+ 421.0987, found 421.0984.
(6S*,12S*)-N,N,N′,N′-Tetramethyl-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine-2,8-diamine (2g)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1g (116.1 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (25% EtOAc/hexanes) to give a yellow oil (58.7 mg, 53%): IR (KBr) 3324, 2888, 2802, 1729, 1621, 1515 cm–1; 1H NMR (400 MHz, CDCl3) δ 6.84 (br s, 2H), 6.81–6.75 (m, 2H), 6.70 (dd, J = 8.7, 2.0 Hz, 2H), 4.51 (br s, 2H, 2 × NH), 2.85 [s, 12H, 2 × N(CH3)2]; 13C{1H} NMR (100 MHz, CDCl3) δ 146.7, 130.5, 123.1 (q, JCF = 282.4 Hz), 122.1, 121.5, 116.1, 109.6, 84.3 (q, JCF = 30.9 Hz), 41.1 [N(CH3)2]; 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI) m/z calcd for C20H20F6N4O [M + H]+ 447.1614, found 447.1610.
Dimethyl (6SR*,12SR*)-6,12-Bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine-2,8-dicarboxylate (2h)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1h (123.6 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The residue was chromatographed on silica (10–15% EtOAc/hexanes) to give a white solid (145.0 mg, 56%). The reaction was also performed on a 4.1 mmol scale (1.0 g of o-TMFK 1h) to give the product in 62% yield (597.2 mg): mp 256–257 °C (n-heptane); IR (KBr) 3350, 3308, 3010, 2958, 2850, 2489, 1716, 1620, 1508, 1491 cm–1; 1H NMR (400 MHz, CD3OD) δ 7.54–7.47 (m, 4H), 7.44 (dd, J = 8.3, 1.8 Hz, 2H), 4.80 (s, 2H), 3.82 [s, 6H, 2 × (CO2CH3)]; 13C{1H} NMR (100 MHz, CD3OD) δ 167.3 (CO2CH3), 143.3, 133.0, 126.5 (q, JCF = 2.9 Hz), 133.0, 124.2 (q, JCF = 281.5 Hz), 121.7, 119.5, 84.3 (q, JCF = 32.4 Hz), 52.7 (CO2CH3); 19F NMR (376 MHz, CD3OD) δ −80.4; HRMS (ESI) m/z calcd for C20H14N2O5F6Na [M + Na]+ 499.0705, found 499.0691.
(6S*,12S*)-2,6,8,12-Tetrakis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2i)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1i (257.1 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (10% EtOAc/hexanes) to give an off-white solid (28.3 mg, 23%). The reaction was conducted on a 0.5 mmol scale for 48 h at 120 °C to afford the product in 39% yield after chromatography (47.8 mg): mp 142–143 °C (n-heptane); IR (KBr) 3413, 1631, 1595, 1524; 1H NMR (400 MHz, CDCl3) δ 7.69 (br s, 2H), 7.50 (dd, J = 8.5, 1.6 Hz, 2H), 6.93 (d, J = 8.6 Hz, 2H), 5.30 (br s, 2H, 2 × NH); 13C{1H} NMR (100 MHz, CDCl3) δ 142.6, 127.8 (q, JCF = 3.5 Hz), 124.1 (q, JCF = 33.1 Hz), 123.7 (q, JCF = 270.0 Hz), 122.3 (q, JCF = 282.5 Hz), 123.0–122.6 [m, F3CCHCCCF3(NH)], 119.4, 118.6, 82.5 (q, JCF = 32.7 Hz); 19F NMR (470 MHz, CDCl3) δ −62.1, −79.1; HRMS (EI) m/z calcd for C18H8F12N2O [M]+• 496.0445, found 496.0455.
(6S*,12S*)-2,8-Dichloro-6,12-bis(heptafluoropropyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2j)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1j (161.8 mg, 0.5 mmol) and TMG (12.5 μL, 20 mol %, 0.1 mmol). The crude product was chromatographed on silica (5% EtOAc/hexane) to give an off-white solid (65.2 mg, 41%): mp 125–127 °C (n-heptane); IR (KBr) 3382, 2932, 1709, 1610, 1490 cm–1; 1H NMR (400 MHz, CDCl3) δ 7.41 (br s, 2H), 7.21 (dd, J = 8.6, 2.3 Hz, 2H), 6.80 (d, J = 8.7 Hz, 2H), 4.97 (br s, 2H, NH); 13C{1H} NMR (100 MHz, CDCl3) δ 138.3, 130.9, 127.3, 125.7–125.9 (m), 121.6, 120.6, 84.6–83.8 (m) (signals of perfluorinated groups have been omitted from the description of the 13C NMR spectrum for the sake of clarity due to complicated multiplicity); 19F NMR (376 MHz, CDCl3) δ −80.8 to −80.9 (m, 3F), −116.5 to −119.3 (m, 2F), −121.6 to −124.3 (m, 2F); HRMS (ESI) m/z calcd for C20H7F14N2O [M – H]− 626.9712, found 626.9726.
(6S*,12S*)-2,8-Bis(prop-2-yn-1-yloxy)-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2k). The title compound was obtained according to GP3 (16 h, 80 °C) using ketone 1k (100 mg, 0.41 mmol) and TMG (10.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (40% DCM/hexanes) to give a white solid (95.0 mg, 99%). The reaction was slightly scaled up using ketone 1k (287.0 g, 1.18 mmol) and TMG (30 μL, 0.236 mmol, 20 mol %) to give a white solid (238.5 mg, 86%): mp 159–160 °C (DCM/n-heptane); 1H NMR (400 MHz CDCl3) δ 7.10 (br s, 2H), 6.98–6.75 (m, 4H), 4.65 (br s, 2H, NH), 4.60 (br s, 4H, OCH2), 2.50 (br s, 2H, CH2CCH); 13C{1H} NMR (100 MHz, CDCl3) δ 153.2, 134.2, 122.9 (q, JCF = 282.6 Hz), 121.8, 121.6, 118.3, 112.3 (q, JCF = 2.7 Hz), 83.9 (q, JCF = 31.4 Hz), 78.3, 75.9, 56.6; 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI) m/z calcd for C22H15F6N2O3 [M + H]+ 469.0987, found 469.0996.
(6S*,12S*)-2,8-Bis(prop-2-yn-1-yloxy)-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2l). The title compound was obtained according to GP3 (16 h, 80 °C) using ketone 1l (122.6 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (15–30% DCM/hexanes) to give a white solid (102.4 mg, 87%). The reaction was conducted on a 4.1 mmol scale using ketone 1l (1.0 g, 4.08 mmol) and TMG (102 μL, 0.82 mmol, 20 mol %) to give 2l as a white solid (918.0 mg, 95%): mp >160 °C dec (DCM/n-heptane); 1H NMR (400 MHz CDCl3) δ 7.04–7.00 (m, 2H), 6.87–6.76 (m, 4H), 6.05–5.94 (m, 2H, CH2CH=CH2), 5.37 (ddd, J = 17.2, 3.2, 1.6 Hz, 2H, CH2CH=CHH’), 5.27 (ddd, J = 10.5, 2.8, 1.4 Hz, 2H, CH2CH=CHH′), 4.57 (m, 2H), 4.47–4.42 (m, 4H, OCH2); 13C{1H} NMR (50 MHz, CDCl3) δ 154.3, 133.5, 133.1, 123.0 (q, JCF = 282.5 Hz), 121.9, 121.7, 118.1, 118.0, 111.6 (q, JCF = 3.2 Hz), 84.0 (q, JCF = 31.5 Hz), 84.9, 84.3, 83.6, 83.0, 69.4; 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI) m/z calcd for C22H19F6N2O3 [M + H]+ 473.1300, found 473.1315.
(6S*,12S*)-2,8-Bis(but-3-en-1-yloxy)-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2m). The title compound was obtained according to GP3 (16 h, 80 °C) using ketone 1m (501.1 mg, 1.92 mmol) and TMG (48.4 μL, 0.39 mmol, 20 mol %). The progress of the reaction was monitored by TLC on alumina. The crude product was chromatographed on aluminum oxide (Brockman activity scale I, 5–10% EtOAc/hexanes) to give a white solid (407.5 mg, 84%): mp 138–139 °C (DCM/n-heptane); 1H NMR (400 MHz, CDCl3) δ 7.00 (br s, 2H), 6.85–6.74 (m, 4H), 5.93–5.79 (m, 2H), 5.19–5.04 (m, 4H), 4.58 (s, 2H), 3.97–3.87 (m, 4H), 2.54–2.43 (m, 4H); 13C{1H} NMR (100 MHz, CDCl3) δ 154.6, 134.3, 133.3, 122.8 (q, J = 282.5 Hz), 121.8, 121.6, 117.8, 117.1, 83.9 (q, J = 31.5 Hz), 67.7, 33.6; 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI) m/z calcd for C24H23F6N2O3 [M + H]+ 501.1613, found 501.1609.
(6S*,12S*)-2,8-Bis(pent-4-en-1-yloxy)-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2n)
The title compound was obtained according to GP3 (16 h, 80 °C) using ketone 1n (1.50 g, 5.49 mmol) and TMG (138 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (20–30% DCM/hexanes) to give an off-white solid (1.30 g, 89%): mp 125–127 °C (DCM/n-heptane); 1H NMR (400 MHz CDCl3) δ 6.99 (s, 2H), 6.85–6.76 (m 4H), 5.90–5.77 (m, 2H, CH2=CHCH2), 5.10–4.96 (m, 4H, CH2=CHCH2), 4.59 (br s, 2H), 3.88 (t, J = 6.3 Hz, 4H, OCH2), 2.25–2.16 (m, 4H, CH2=CHCH2), 1.88–1.79 (m, 4H, CH2CH2CH2); 13C{1H} NMR (100 MHz, CDCl3) δ 154.9, 137.9, 133.3, 123.0 (q, JCF = 282.4 Hz), 122.0, 121.7, 117.8, 115.3, 111.3 (q, JCF = 3.0 Hz), 84.0 (q, JCF = 31.4 Hz), 67.7, 30.2, 28.5; 19F NMR (376 MHz, CDCl3) δ −79.2; HRMS (ESI) m/z calcd for C26H26F6N2O2Na [M + Na]+ 551.1745, found 551.1742.
(6S*,12S*)-2,8-Dichloro-6,12-bis(difluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2o)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1o (102.8 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (10% EtOAc/hexanes) to give an off-white solid (74.4 mg, 76%). The reaction was conducted on a 4.9 mmol scale (1.0 g) to afford 668.1 mg of diazocine 2o (70%): mp 175–177 °C (n-heptane); IR (KBr) 3372, 3320, 3072, 3000, 1897, 1775, 1739, 1608, 1577, 1499 cm–1; 1H NMR (400 MHz, CDCl3) δ 7.40 (br s, 2H), 7.16 (dd, J = 8.6, 2.3 Hz, 2H), 6.74 (d, J = 8.6 Hz, 2H), 6.02 (dd, JHF = 54.9, 54.9 Hz, 2H, 2 × CHF2), 4.88 (br s, 2H, 2 × NH); 13C{1H} NMR (100 MHz, CDCl3) δ 138.9, 130.2, 126.4, 125.6 (dd, JCF = 4.2, 4.2 Hz), 122.3, 119.6, 114.1 (dd, JCF = 254.2, 248.9 Hz), 81.7 (dd, JCF = 25.3, 23.2 Hz); 19F NMR (376 MHz, CDCl3) δ −129.7 (d, J = 290.4 Hz), −131.6 (d, J = 290.5 Hz); HRMS (ESI) m/z calcd for C16H9F4N2OCl2 [M – H]− 391.0028, found 391.0034.
(6S*,12S*)-7,15-Bis(trifluoromethyl)-7,8,15,16-tetrahydro-7,15-epoxydinafto[1,2-b:1′,2′-f][1,5]diazocine (2p)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 1p (119.6 mg, 0.5 mmol) and TMG (12.5 μL, 20 mol %, 0.1 mmol). The crude product was chromatographed on silica (5% EtOAc/hexane; the Rf of 2p is slightly higher than that of substrate 1p) to give a bright yellow solid (50.8 mg, 44%): mp 208–210 °C (n-heptane); IR (KBr) 3321, 3075, 1582, 1514 cm–1; 1H NMR (400 MHz, CDCl3) δ 8.05 (d, J = 8.0 Hz, 2H), 7.74 (d, J = 7.6 Hz, 2H), 7.60 (d, J = 8.7 Hz, 2H), 7.57–7.41 (m, 6H), 5.25 (s, 2H, 2 × NH); 13C{1H} NMR (100 MHz, CDCl3) δ 136.3, 134.0, 128.2, 127.4, 126.5, 125.8, 123.2 (q, JCF = 282.9 Hz), 122.9, 122.2 (q, JCF = 3.0 Hz), 121.1, 116.0, 84.1 (q, JCF = 31.6 Hz); 19F NMR (376 MHz, CDCl3) δ −78.2; HRMS (ESI) m/z calcd for C24H14F6N2ONa [M + Na]+ 483.0908, found 483.0899.
(6S,12S)-2,8-Bis{[(1R,2S,5R)-5-methyl-2-(propan-2-yl)cyclohexyl]oxy}-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2q) and (6R,12R)-2,8-Bis{[(1R,2S,5R)-5-methyl-2-(propan-2-yl)cyclohexyl]oxy}-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (2q′). The title compounds were obtained according to GP3 (16 h, 120 °C) using ketone 1q (200 mg, 0.58 mmol) and TMG (14.6 μL, 0.12 mmol, 20 mol %). The crude product was chromatographed on silica (2% EtOAc/hexane) to give a yellow foam (143.7 mg, 74%). The diastereomeric ratio was estimated on the basis of 19F NMR to be 53:47. The diastereomeric ratio was independently determined by HPLC analysis using a Daicel Chiralpak OD-H column (2% i-PrOH/hexane, flow rate of 1.0 mL/min, λ = 235 nm) to give a similar result (52:48): tR = 5.4 min (major), tR = 9.2 min (minor); IR (KBr) 3333, 2956, 2927, 2871, 1723, 1617, 1581, 1502 cm–1; 1H NMR (400 MHz, CDCl3) δ 6.99 (br s, 2H), 6.86–6.75 (m, 4H), 4.55 and 4.55 (2 × br s, 2H, NH), 3.92–3.81 (m, 2H), 2.25–2.11 (m, 2H), 2.10–1.98 (m, 2H), 1.75–1.65 (m, 4H), 1.50–1.34 (m, 4H), 1.14–0.94 (m, 4H), 0.94–0.82 (m, 14H), 0.75 (d, J = 6.7 Hz, 3H), 0.73 (d, J = 6.7 Hz, 3H); 13C{1H} NMR (100 MHz, CDCl3) δ 153.9 (×2), 133.2, 133.1, 122.9 (q, JCF = 282.6 Hz, ×2), 121.8 (×2), 121.4, 121.3, 119.3, 119.0, 113.5 (q, JCF = 2.7 Hz), 113.2 (q, JCF = 2.8 Hz), 83.8 (q, JCF = 31.5 Hz, ×2), 78.7, 78.6, 48.2 (×2), 40.3, 40.2, 34.5, 34.4, 31.4, 25.9 (×2), 23.6, 23.5, 22.1, 20.8, 16.4, 16.3; 19F NMR (376 MHz, CDCl3) δ −79.2, −79.3; HRMS (EI) m/z calcd for C36H46F6N2O3 [M]+• 668.3413, found 668.3405.
(5S*,11S*)-3,9-Dichloro-5,11-bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxydipyrido[2,3-b:2′,3′-f][1,5]diazocine (4a)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 2a (112.3 mg, 0.5 mmol) and TMG (12.5 μL, 20 mol %, 0.1 mmol). The crude product was chromatographed on silica (5% EtOAc/toluene) to give a light gray solid (105.2 mg, 98%): mp 252–253 °C (n-heptane); IR (KBr) 3190, 3057, 2925, 1869, 1842, 1607, 1577, 1505 cm–1; 1H NMR (400 MHz, CD3OD) δ 8.17 (d, J = 2.3 Hz, 2H), 7.78–7.74 (m, 2H), 4.54 (br s, 2H, 2 × NH); 13C{1H} NMR (100 MHz, CD3OD) δ 153.4, 149.9, 134.6 (q, JCF = 2.8 Hz), 123.9, 123.0 (q, JCF = 281.7 Hz), 117.0, 84.3 (q, JCF = 33.5 Hz); 19F NMR (376 MHz, CD3OD) δ −80.7; HRMS (EI) m/z calcd for C14H6Cl2F6N4O [M]•+ 429.9823, found 429.9817.
(5S*,11S*)-5,11-Bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxydipyrido[2,3-b:2′,3′-f][1,5]diazocine (4b)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 2c (95.0 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (5% MTBE/DCM to 5–10% MeOH/DCM) to give a white solid (59.2 mg, 65%): mp 309–310 °C (EtOH/n-heptane, product not soluble in CHCl3, DCM, n-heptane, MeOH); IR (KBr) 3161, 3105, 3004, 2929, 2880, 1914, 1912, 1609, 1587, 1525 cm–1; 1H NMR (400 MHz, DMSO-d6) δ 8.96 (br s, 2H, 2 × NH), 8.17 (dd, J = 4.8, 1.5 Hz, 2H), 7.73 (d, J = 7.8 Hz, 2H), 6.90 (dd, J = 7.8, 4.8 Hz, 2H); 13C{1H} NMR (100 MHz, DMSO-d6) δ 153.7, 150.3, 134.1, 122.5 (q, JCF = 282.5 Hz), 118.3, 116.1, 114.6, 83.0 (q, JCF = 32.6 Hz); 19F NMR (376 MHz, DMSO-d6) δ −78.3; HRMS (ESI) m/z calcd for C14H9F6N4O [M + H]+ 363.0681, found 363.0670.
(6S*,12S*)-6,12-Bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydipyrido[4,3-b:4′,3′-f][1,5]diazocine (4c)
The title compound was obtained according to GP3 (16 h, 120 °C) using ketone 2b (95.1 mg, 0.5 mmol) and TMG (12.5 μL, 0.1 mmol, 20 mol %). The crude product was chromatographed on silica (6–10% MeOH/DCM) to give an off-white solid (73.8 mg, 85%): mp >325 °C dec; IR (KBr) 3176, 3126, 3094, 3038, 2993, 2928, 2873, 2789, 2839, 2518, 2322, 1938, 1901, 1866, 1617, 1585, 1531 cm–1; 1H NMR (400 MHz, CD3OD) δ 8.37 (s, 2H), 8.16 (d, J = 5.8 Hz, 2H), 6.85 (d, J = 5.8 Hz, 2H); 13C{1H} NMR (100 MHz, CD3OD) δ 150.6, 150.1, 146.4–146.1 (m, 1H), 123.8 (q, JCF = 281.7 Hz), 117.0, 112.4, 82.7 (q, JCF = 33.8 Hz); 19F NMR (376 MHz, CD3OD) δ −80.7; HRMS (ESI) m/z calcd for C14H9F6N4O [M + H]+ 363.0681, found 363.0671.
General Procedure for the Synthesis of Unsymmetrical Epoxydibenzo[b,f][1,5]diazocines (GP4)
Fluoromethylketone pyridine-derived 3a (FMK, x mmol), fluoroketone 1a (2 equiv), and N,N,N′,N′-tetramethylguanidine (TMG, 20 mol %/equiv of aminophenone) were placed in a screw-cap vial (4 mL), and the resulting mixture was heated at 120 °C (IKA heating block, temperature of the reference vial filled with silicon oil). Then the reaction mixture was diluted with EtOAc or DCM (10 mL), adsorbed on silica (or aluminum oxide), and chromatographed to give corresponding epoxydibenzo[b,f][1,5]diazocines 5. The analytical sample was crystallized from the appropriate solvent to measure the melting point. In all cases, in addition to cross-condensation product 5, diazocine 2a was formed and isolated via chromatography. In addition, diazocine 4b (resulting from autocondensation of 3c) and unreacted aminophenone 3c were detected by TLC (4b and 3c were not isolated due to their marginal amounts).
(5S*,11S*)-9-Fluoro-5,11-bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxypyrido[3,2-c][1,5]benzodiazocine (5a)
The title compound was obtained according to GP4 (16 h, 120 °C) using ketone 1a (223.6 mg, 1.0 mmol), ketone 3c (95.1 mg, 0.5 mmol), and TMG (37.5 μL, 0.3 mmol, 20 mol %/equiv of aminophenone). The residue was chromatographed on silica (10–25% EtOAc/hexanes) to give diazocine 2a (127.1 mg) and a mixture of diazocine 5a and pyridine derivative 3c. The resulting mixture was chromatographed on alumina (15% EtOAc/hexanes, activity on Brockman scale III) to give 5a as a white solid (126.4 mg, 64%).
Experiment on a 2.5 mmol Scale
Ketone 1a (1.118 g, 5.0 mmol, 2 equiv), ketone 3c (475.5 mg, 2.5 mmol), and TMG (188 μL, 1.5 mmol) were placed in a 4 mL screw-cap vial and heated at 120 °C for 16 h. Then the reaction mixture was cooled to rt, dissolved in a mixture of MeOH and DCM (10% MeOH/DCM), adsorbed on silica (10 g), and chromatographed (10–25% EtOAc/hexanes) to give a light-yellow solid of 5a (469.3 mg, 48%). In addition, 2a (712.1 mg) was also isolated: mp 227–229 °C (n-heptane); IR (KBr) 3381, 3341, 3196, 3181, 3168, 3097, 3077, 3002, 2948, 1600, 1589, 1512 cm–1; 1H NMR (400 MHz, CD3OD) δ 8.13 (dd, J = 4.9, 1.7 Hz, 1H), 7.79–7.73 (m, 1H), 7.45–7.40 (m, 1H), 7.19 (dd, J = 8.7, 2.4 Hz, 1H), 6.86 (dd, J = 7.8, 4.9 Hz, 1H) overlapping 6.84 (d, J = 8.8 Hz, 1H), 4.55 (br s, 2H, 2 × NH); 13C{1H} NMR (100 MHz, CD3OD) δ 155.3, 150.8, 141.4, 135.5 (q, JCF = 2.2 Hz), 131.4, 126.4, 126.0 (q, JCF = 3.3 Hz), 124.1 (q, JCF = 281.6 Hz), 123.9 (q, JCF = 281.6 Hz), 122.5, 120.5, 117.0, 116.5, 84.3 (q, JCF = 32.6 Hz), 84.2 (q, JCF = 33.0 Hz); 19F NMR (376 MHz, CD3OD) δ −80.6, −80.7; HRMS (ESI) m/z calcd for C15H9ClF6N3O [M + H]+ 396.0338, found 396.0329.
(5S*,11S*)-9-Methyl-5,11-bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxypyrido[3,2-c][1,5]benzodiazocine (5b)
The title compound was obtained according to GP4 (16 h, 120 °C) using ketone 1c (207.1 mg, 1.0 mmol), ketone 3c (95.1 mg, 0.5 mmol), and TMG (37.5 μL, 0.3 mmol, 20 mol %/equiv of aminophenone). The residue was chromatographed on silica (10% EtOAc/hexanes) to give 2c (154.9 mg) and a mixture of diazocine 5b and pyridine derivative 3c. The resulting mixture of 5b and 3c was further chromatographed on alumina (15% EtOAc/hexanes, activity on Brockman scale III) to give 5b as a white solid (66.4 mg, 35%): mp 196–197 °C (n-heptane); IR (KBr) 3356, 3161, 3098, 2994, 2926, 1604, 1588, 1445 cm–1; 1H NMR (400 MHz, CD3OD) δ 8.12 (dd, J = 4.8, 1.4 Hz, 1H), 7.81–7.72 (m, 1H), 7.22–7.15 (m, 1H), 7.03–6.96 (m, 1H), 6.91–6.82 (m, 2H), 4.55 (br s, 1H, NH); 13C{1H} NMR (100 MHz, CD3OD) δ 158.6 (d, JCF= 237.5 Hz), 155.4, 150.8, 138.6 (d, JCF = 1.8 Hz), 135.7 (q, JCF= 2.7 Hz), 124.1 (q, JCF = 281.6 Hz), 123.8 (q, JCF = 281.6 Hz), 122.3 (d, JCF= 6.8 Hz), 121.2 (d, JCF= 7.6 Hz), 118.6 (d, JCF= 23.1 Hz), 117.0, 116.5, 112.5 (dq, JCF= 24.8, 3.0 Hz), 84.7–83.6 (m) overlapping 84.7 (q, JCF= 32.2 Hz); 19F NMR (376 MHz, CD3OD) δ −80.7 (×2), −123.7; HRMS (ESI) m/z calcd for C15H9F7N3O [M + H]+ 380.0634, found 380.0633.
(5S*,11S*)-9-Chloro-1-methyl-5,11-bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxypyrido[3,2-c][1,5]benzodiazocin-1-ium iodide (5c). The title compound was obtained according to GP4 (16 h, 120 °C) using ketone 1e (203.2 mg, 1.0 mmol), ketone 3c (95.1 mg, 0.5 mmol), and TMG (37.5 μL, 0.3 mmol, 20 mol %/equiv of aminophenone). The residue was chromatographed on silica (15–25% EtOAc/hexanes) to give 2e (135.9 mg) and a mixture of diazocine 5c and pyridine derivative 3c. The resulting mixture of 5c and 3c was chromatographed on alumina (30% MTBE/hexanes) to give a colorless solid (82.9 mg, 44%): mp 209–210 °C (n-heptane); IR (KBr) 3374, 3333, 3182, 3098, 3070, 3000, 2936, 1948, 1928, 1601, 1588, 1516 cm–1; 1H NMR (400 MHz, CD3OD) δ 8.09 (dd, J = 4.9, 1.7 Hz, 1H), 7.78–7.73 (m, 1H), 7.24 (br s, 1H), 7.01 (dd, J = 8.3, 1.4 Hz, 1H), 6.81 (dd, J = 7.8, 4.9 Hz, 1H), 6.76 (d, J = 8.2 Hz, 1H), 2.21 (s, 3H, CH3); 13C{1H} NMR (100 MHz, CD3OD) δ 155.6, 150.5, 139.8, 135.5 (q, JCF = 2.8 Hz), 132.0, 131.7, 126.1 (q, JCF = 2.8 Hz), 124.3 (q, JCF = 281.6 Hz), 124.1 (q, JCF = 281.4 Hz), 121.4, 119.5, 117.0, 116.8, 84.7 (q, J = 32.1 Hz), 84.5 (q, J = 32.7 Hz), 20.7 (CH3); 19F NMR (376 MHz, CD3OD) δ −80.4, −80.7; HRMS (EI) m/z calcd for C16H11F6N3O [M]•+ 376.0875, found 376.0876.
(5S*,11S*)-9-Chloro-5-(difluoromethyl)-11-(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxypyrido[3,2-c][1,5]benzodiazocine (5d). The title compound was obtained according to GP4 (16 h, 120 °C) using ketone 1a (223.6 mg, 1.0 mmol), ketone 3d (86.1 mg, 0.5 mmol), and TMG (37.5 μL, 0.3 mmol, 20 mol %/equiv of aminophenone). The crude product was chromatographed on silica (10% EtOAc/hexanes) to give 2a (136.6 mg) and a mixture of diazocine 5d and pyridine derivative 3d (a complicated mixture of products was detected by TLC). The resulting mixture was chromatographed on reversed phase silica (20% H2O/MeOH, RP-18) to give a colorless solid (22.7 mg, 12%) mp 229–230 °C (n-heptane); IR (KBr) 3412, 3383, 3336, 3202, 3090, 2992, 2943, 2850, 1899, 1601, 1565, 1508 cm–1; 1H NMR (400 MHz, CD3OD) δ 8.10 (br s, 1H), 7.77 (d, J = 7.8 Hz, 1H), 7.44–7.40 (m, 1H), 7.15 (dd, J = 8.7, 2.4 Hz, 1H), 6.86–6.79 (m, 2H), 6.23 (dd, JHF= 54.7, 54.3 Hz, 1H, CHF2); 13C{1H} NMR (100 MHz, CD3OD) δ 155.5, 150.1, 141.7, 135.8 (dd, JCF= 4.3, 2.5 Hz), 131.2, 126.0 (q, JCF = 3.0 Hz), 126.0, 124.1 (q, JCF= 281.6 Hz), 122.8, 120.4, 117.9–117.6 (m), 117.1–116.7 (m), 115.7 (dd, JCF = 248.4, 247.0 Hz), 84.4–83.2 (m); 19F NMR (376 MHz, CD3OD) δ −80.5, −132.0 (dd, J = 291.2, 0.0 Hz), −133.6 (dd, J = 291.1, 0.0 Hz); HRMS (ESI) m/z calcd for C15H10ClF5N3O [M + H]+ 378.0433, found 378.0427.
(5S,11S)-9-{[(1R,2S,5R)-5-Methyl-2-(propan-2-yl)cyclohexyl]oxy}-5,11-bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxypyrido[3,2-c][1,5]benzodiazocine (5e) and (5R,11R)-9-{[(1R,2S,5R)-5-Methyl-2-(propan-2-yl)cyclohexyl]oxy}-5,11-bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxypyrido[3,2-c][1,5]benzodiazocine (5e′). The title compounds were obtained according to GP4 (16 h, 120 °C) using ketone 3c (95.1 mg, 0.5 mmol), ketone 1q (343.4 mg, 1.0 mmol), and TMG (37.5 μL, 0.3 mmol, 20 mol %/equiv of aminophenone). The residue was chromatographed on silica (10–15% EtOAc/hexanes) to give 2q (214.4 mg) and an inseparable mixture of 5e and 5e′ as a white solid (145.0 mg, 56%): mp 197–198 °C (n-heptane); IR (KBr) 3309, 3213, 3093, 2960, 2929, 2875, 2395, 1925, 1600, 1504 cm–1; 1H NMR (600 MHz, CD3OD) δ 8.10 (dd, J = 4.9, 1.5 Hz, 1H), 7.77 (d, J = 7.8 Hz, 1H), 7.01 (br s, 1H), 6.86–6.78 (m, 3H), 3.97–3.88 (m, 1H), 2.22–2.11 (m, 1H), 2.10–2.02 (m, 1H), 1.75–1.66 (m, 2H), 1.50–1.38 (m, 2H), 1.17–1.07 (m, 1H), 0.97–0.84 (m, 8H), 0.75 (d, J = 7.0 Hz, 0.5 × 3H), 0.74 (d, J = 7.0 Hz, 0.5 × 3H); 13C{1H} NMR (150 MHz, CD3OD) δ 155.6 (×2), 154.4, 154.2, 150.6, 135.8, 135.4, 135.3, 124.3 (q, JCF = 281.7 Hz), 124.1 (q, JCF = 281.1 Hz), 122.6, 122.5, 121.4, 121.4, 120.3, 120.1, 116.8 (×2), 113.3, 113.2, 85.6–84.5 (m) overlapping 84.3 (q, JCF = 32.3 Hz), 79.6, 49.7, 49.6, 41.6, 41.5, 35.7, 35.6, 32.5 (×2), 27.2, 24.7 (×2), 22.5 (×2), 21.1 (×2) 16.9, 16.8; 19F NMR (376 MHz, CD3OD) δ −80.4, −80.5, −80.7, −80.7; HRMS (ESI) m/z calcd for C25H28F6N3O2F6 [M + H]+ 516.2086, found 516.2081.
Functionalization of Epoxydibenzo[b,f][1,5]diazocines
(5S*,11S*)-9-Chloro-1-methyl-5,11-bis(trifluoromethyl)-5,6,11,12-tetrahydro-5,11-epoxypyrido[3,2-c][1,5]benzodiazocin-1-ium iodide (6a). Epoxydibenzodiazocine 5a (60 mg, 0.152 mmol, 1 equiv), MeCN (2 mL), and MeI (66 μL, 1.06 mmol, 7 equiv) were placed in a screw-cap vial (4 mL), and the resulting mixture was heated at 80 °C for 48 h (IKA heating block, temperature of the reference vial filled with silicon oil). Then the reaction mixture was evaporated. The residue was chromatographed on silica (5–15% EtOAc/hexanes) to give a pale green solid (63.1 mg, 77%): mp >180 °C dec (n-heptane); 1H NMR (600 MHz, CD3OD) δ 7.37 (dd, J = 6.8, 1.7 Hz, 1H), 7.34–7.28 (m, 2H), 7.08 (dd, J = 8.7, 2.4 Hz, 1H), 6.78 (d, J = 8.6 Hz, 1H), 5.97 (t, J = 7.0 Hz, 1H), 3.44 (s, 3H, CH3); 13C{1H} NMR (150 MHz, CD3OD) δ 156.1, 141.7, 140.7, 133.2 (q, JCF = 3.0 Hz), 129.8, 125.9 (q, JCF = 3.4 Hz), 125.2, 125.1 (q, JCF = 280.9 Hz), 124.7, 124.4 (q, JCF = 281.9 Hz), 121.6, 119.6, 103.9, 88.3 (q, JCF = 31.3 Hz), 83.0 (q, JCF = 32. Hz), 39.5; 19F NMR (376 MHz, CD3OD) δ −80.7, −81.1; HRMS (ESI) m/z calcd for C16H11ClF6N3O [M]+ 410.0495, found 410.0483.
(6S*,12S*)-2,8-Dimethyl-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydipyrido[4,3-b:4′,3′-f][1,5]diazocine-2,8-diium Diiodide (6b). To a solution (in a screw-cap 4 mL vial) of epoxydibenzodiazocine 4b (50 mg, 0.138 mmol, 1 equiv) in a mixture of MeCN (2 mL) and MeOH (1 mL) was added MeI (52 μL, 0.828 mmol, 6 equiv), and the resulting mixture was heated at 90 °C for 18 h (IKA heating block, temperature of the reference). Then the reaction mixture was diluted with MeOH and evaporated with silica. The residue was chromatographed on silica (10–20% MeOH/DCM) to give a white solid (89.2 mg, ∼100%). The reaction was conducted on large scale using epoxydibenzodiazocine 4b (150 mg, 0.414 mmol), MeI (0.15 mL, 2.49 mmol, 6.0 equiv), MeCN (6 mL), and MeOH (3 mL). The resulting homogeneous mixture (MeOH was used to solubilize diazocine 4b) was heated at 90 °C for 18 h. Then the reaction mixture was evaporated, redissolved in a minimal volume of MeOH (3 mL), and precipitated with Et2O (6 mL). The resulting solid was filtered, washed with Et2O (3 × 2 mL), and dried in vacuo to give an off-white solid (230.4 mg, 86%): mp >330 °C (MeOH/Et2O); 1H NMR (400 MHz, CD3OD) δ 7.84 (s, 2H), 7.72 (d, J = 7.4 Hz, 2H), 6.67 (d, J = 7.4 Hz, 2H), 3.82 (s, 6H, 2 × CH3); 13C{1H} NMR (125 MHz, CD3OD) δ 157.0, 142.3, 138.6, 124.2 (q, JCF = 281.4 Hz), 117.0, 114.1, 84.7 (q, JCF = 33.0 Hz), 45.2; 19F NMR (376 MHz, CD3OD) δ −80.5; HRMS (ESI) m/z calcd for C16H13F6N4O [M – H]+ 391.0995, found 391.0995.
(6S*,12S*)-N,N′-Bis(2-hydroxyethyl)-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine-2,8-dicarboxamide (8a). To a solution of epoxydibenzo[b,f][1,5]diazocine 2h (50.0 mg, 0.105 mmol) in THF (1.5 mL) were added ethanolamine 7a (19.2 mg, 0.315 mmol) and TBD (8.8 mg, 0.063 mmol, 60 mol %). The resulting reaction mixture was heated at 60 °C for 21 h, quenched with a saturated solution of NH4Cl (10 mL), and extracted with EtOAc (6 × 15 mL). The combined organic extracts were dried over Na2SO4 and evaporated. The residue was chromatographed on silica (10–15% MeOH/DCM) to give a white solid (33.6 mg, 60%): mp >119 °C dec (MeOH/DCM); 1H NMR (400 MHz, CD3OD) δ 7.49 (d, J = 8.2 Hz, 2H), 7.29–7.21 (m, 4H), 3.66 (t, J = 5.8 Hz, 4H, CH2), 3.44 (t, J = 5.8 Hz, 4H, CH2); 13C{1H} NMR (100 MHz, CD3OD) δ 169.8 (CO2R), 143.4, 137.7, 128.5, 126.5 (q, JCF = 2.1 Hz), 124.3 (q, JCF = 281.5 Hz), 123.9, 119.5, 117.6, 84.3 (q, JCF = 32.2 Hz), 61.5, 43.5; 19F NMR (376 MHz, CD3OD) δ −80.5; HRMS (ESI) m/z calcd for C22H20F6N4O5Na [M + Na]+ 557.1236, found 557.1232.
(6S,12S)-N,N′-Bis[(2S)-1-hydroxy-3-methylbutan-2-yl]-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine-2,8-dicarboxamide (8b) and (6R,12R)-N,N′-Bis[(2S)-1-hydroxy-3-methylbutan-2-yl]-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine-2,8-dicarboxamide (8b′). To a solution of epoxydibenzo[b,f][1,5]diazocine 2h (100.0 mg, 0.210 mmol) in THF (2 mL) were added (S)-(+)-2-amino-3-methyl-1-butanol (l-Valinol, 7b) (64.9 mg, 0.630 mmol) and TBD (17.5 mg, 0.126 mmol, 60 mol %), and the resulting reaction mixture was heated at 60 °C for 48 h. Then the reaction was quenched with solid NH4Cl (33.7 g, 0.630 mmol), and the mixture diluted with MeOH (15 mL) and adsorbed on silica. The residue was chromatographed on silica (5–10% MeOH/DCM) to give a white solid (90.4 mg, 70%). The diastereomeric ratio was estimated on the basis of 19F NMR to be 55:45: 1H NMR (400 MHz, CD3OD) δ 7.49 (d, J = 8.4 Hz, 2H), 7.28–7.19 (m, 4H), 3.90–3.79 (m, 2H), 3.71–3.56 (m, 4H), 2.00–1.85 (m, 2H), 1.04–0.82 [m, 12H, HOCH2CHCH(CH3)2]; 13C{1H} NMR (100 MHz, CD3OD) δ 170.1 (CO2R), 143.4, 138.3, 138.2, 126.5, 124.3 (q, JCF = 282.0 Hz), 123.8, 119.7, 119.6, 117.6 (×2), 84.4 (q, JCF = 32.0 Hz), 63.1, 58.8, 30.2 (×2), 20.1, 19.2; 19F NMR (376 MHz, CD3OD) δ −80.4, −80.5; HRMS (ESI) m/z calcd for C28H32F6N4O5Na [M + Na]+ 641.2175, found 641.2184.
Diethyl (2E,2′E)-4,4′-{[(6S*,12S*)-6,12-Bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine-2,8-diyl]bis(oxy)}bisbut-2-enoate (9). The flask was charged with epoxydibenzodiazocine 2l (55.0 mg, 0.116 mmol, 1 equiv), a second-generation Grubbs catalyst (4.9 mg, 0.006 mmol, 5 mol %), and anhydrous DCE (2.3 mL), and a gentle stream of argon was bubbled for 30 min. Then ethyl acrylate (76 μL, 0.699 mmol, 6 equiv) was added, and the reaction mixture was refluxed for 5 h (TLC analysis indicated a partial consumption of the substrate). Then another portion of a second-generation Grubbs catalyst (4.9 mg, 0.006 mmol, 5 mol %) was added, and the mixture was heated at reflux for 19 h. The solvent was evaporated, and the residue was chromatographed on silica (10–15% EtOAc/hexanes) to give a violet oil (35.1 mg, 49%): 1H NMR (400 MHz, CDCl3) δ 7.04–6.96 [m, 4H, (2 × ArH, 2 × CH=CHCO2Et)], 6.84–6.79 (m, 4H), 6.14 (dt, J = 15.8, 2.0 Hz, 2H, CH=CHCO2Et), 4.62–4.54 (m, 4H, ArOCH2), 4.20 (q, J = 7.1 Hz, 4H, OCH2CH3), 1.29 (t, J = 7.1 Hz, 6H, OCH2CH3); 13C{1H} NMR (100 MHz, CDCl3) δ 166.1, 153.8, 142.1, 134.0, 122.9 (q, JCF = 282.6 Hz), 122.0, 121.8, 121.5, 118.0, 111.8 (q, JCF = 3.1 Hz), 83.9 (q, JCF = 31.3 Hz), 67.1, 60.7, 14.3; 19F NMR (376 MHz, CDCl3) δ −79.1; HRMS (ESI) m/z calcd for C28H26F6N2O7Na [M + Na]+ 639.1542, found 639.1536.
(6S*,12S*)-2,8-Bis[(1-benzyl-1H-1,2,3-triazol-4-yl)methoxy]-6,12-bis(trifluoromethyl)-5,6,11,12-tetrahydro-6,12-epoxydibenzo[b,f][1,5]diazocine (10). The contents of a 4 mL screw-cap vial charged with epoxydibenzodiazocine 2k (50.0 mg, 0.182 mmol, 1 equiv), benzyl azide (120.9 mg, 0.908 mmol, 5 equiv), THF (2.5 mL), CuI (2.1 mg, 0.011 mmol, 6 mol %), and Et3N (127 μL, 0.908 mmol, 5 equiv) were stirred for 22 h at rt. The solvent was evaporated in vacuo. The residue was chromatographed on silica (30–40% EtOAc/toluene) to give a white solid (80.2 mg, 60%): mp >162 °C dec (MeOH/DCM); 1H NMR (400 MHz, CD3OD) δ 7.87 (s, 2H), 7.35–7.17 (m, 10H), 7.03 (br s, 2H), 6.81 (dd, J = 8.9, 2.7 Hz, 2H), 6.74 (d, J = 8.8 Hz, 2H), 5.50 (s, 4H, CH2), 4.99 (s, 4H, CH2); 13C{1H} NMR (100 MHz, CD3OD) δ 153.8, 145.3, 137.1, 136.5, 130.0, 129.6, 129.0, 125.2, 124.4 (JCF = 282.1 Hz), 122.4, 120.5, 119.0, 112.7 (JCF = 3.2 Hz), 84.8 (JCF = 31.6 Hz), 63.0, 54.9; 19F NMR (376 MHz, CD3OD) δ −80.4; HRMS (ESI) m/z calcd for C36H28F6N8O3Na [M + Na]+ 757.2078, found 757.2086.
Configuration Assignment and Stability Investigations
ECD spectra at room temperature were measured in acetonitrile using a Jasco J-815 spectrometer in the range of 180–400 nm (c = 2.9 × 10–4 M) in quartz cells with a path length of 0.1 or 1 cm. The following measurement parameters were used: a scanning speed of 100 nm/min, a step size of 0.2 nm, a bandwidth of 1 nm, a response time of 0.5 s, and an accumulation of three scans. The spectra were background-corrected using acetonitrile.
ECD spectra at variable temperatures were measured in decalin (c = 2.75 × 10–4 M) using a Jasco J-715 spectrometer equipped with a dedicated variable-temperature transmission cell holder from Specac. The spectra of (+)- and (−)-2a were recorded from 190 to 400 nm in a quartz cell with a path length of 0.1 cm. Baseline correction was achieved by subtracting the spectrum of decalin obtained under the same conditions. All spectra were normalized to Δε (cubic decimeters per mole per centimeter) using volume correction for decalin.
VCD spectra of enantiomers (+)- and (−)-2a were recorded simultaneously with IR spectra by a ChiralIR-2X instrument from BioTools (Jupiter, FL) at a resolution of 4 cm–1 in the range of 2000–900 cm–1 using CD3CN as a solvent. A solution with a concentration ∼0.2 M was measured in a BaF2 cell with a path length of 100 μm. Spectra were recorded for approximately 3 h to improve the signal-to-noise ratio. Baseline correction was achieved by subtracting the spectrum of a solvent recorded under the same conditions.
Computational Details
A conformational search was carried out at the molecular mechanics level using the MMFF94s force field within 10 kcal/mol for (+)-2a. Next, the found structure was submitted for DFT optimization using Gaussian16[1] at the ωB97X-D/6-311+G(d,p) level of theory applying PCM for CH3CN.
The same level of theory was used for VCD and IR simulations. The VCD simulated spectrum was converted to Lorentzian bands with an 8 cm–1 half-width and was scaled by 0.982 (the best scaling factor, giving the best agreement between the experimental and simulated spectra).
TDDFT calculations of the final ECD spectrum were carried out using the CAM-B3LYP functionals with the def2-TZVP basis set and PCM model for CH3CN. The calculations at the B3LYP/def2-TZVP/PCM and ωB97X-D/def2-TZVP/PCM levels yielded consistent results. Rotatory strengths were calculated using both length and velocity representations. The differences between the length and velocity of the calculated values of the rotatory strengths were <3%, and for this reason, only the velocity representations (Rvel) were taken into account. The UV and ECD spectra are simulated by overlapping Gaussian functions for 40 electronic transitions using bands with a 0.3 eV exponential half-width and red-shifted by 13 nm (UV correction).
Acknowledgments
X-ray diffraction data were collected at the synchrotron at PETRA III (Hamburg, Germany). M.G. thanks the Wroclaw Centre for Networking and Supercomputing (WCSS) for computational support.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.1c00884.
Copies of NMR spectra for all new compounds, chromatograms for separation of enantiomers of (+)-2a, Cartesian coordinates for (+)-2a, and detailed optimization studies of the condensation reaction leading to (+)-2a and (+)-2l (PDF)
Accession Codes
CCDC 2055622 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
Financial support for this work was provided by the Polish National Science Centre (Grants SONATA BIS 2017/26/E/ST5/00510 and MINIATURA 2017/01/X/ST5/01384).
The authors declare no competing financial interest.
Supplementary Material
References
- a Klärner F.-G.; Schrader T. Aromatic Interactions by Molecular Tweezers and Clips in Chemical and Biological Systems. Acc. Chem. Res. 2013, 46, 967–978. 10.1021/ar300061c. [DOI] [PubMed] [Google Scholar]; b Ibáñez S.; Poyatos M.; Peris E. N-Heterocyclic Carbenes: A Door Open to Supramolecular Organometallic Chemistry. Acc. Chem. Res. 2020, 53, 1401–1413. 10.1021/acs.accounts.0c00312. [DOI] [PubMed] [Google Scholar]; c Han Y.; Tian Y.; Li Z.; Wang F. Donor–acceptor-type supramolecular polymers on the basis of preorganized molecular tweezers/guest complexation. Chem. Soc. Rev. 2018, 47, 5165–5176. 10.1039/C7CS00802C. [DOI] [PubMed] [Google Scholar]; d Schrader T.; Bitan G.; Klärner F.-G. Molecular tweezers for lysine and arginine – powerful inhibitors of pathologic protein aggregation. Chem. Commun. 2016, 52, 11318–11334. 10.1039/C6CC04640A. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Krykun S.; Dekhtiarenko M.; Canevet D.; Carré V.; Aubriet F.; Levillain E.; Allain M.; Voitenko Z.; Sallé M.; Goeb S. Metalla-Assembled Electron-Rich Tweezers: Redox-Controlled Guest Release Through Supramolecular Dimerization. Angew. Chem., Int. Ed. 2020, 59, 716–720. 10.1002/anie.201912016. [DOI] [PubMed] [Google Scholar]; f Knezevic M.; Heilmann M.; Piccini G. M.; Tiefenbacher K. Overriding Intrinsic Reactivity in Aliphatic C–H Oxidation: Preferential C3/C4 Oxidation of Aliphatic Ammonium Substrates. Angew. Chem., Int. Ed. 2020, 59, 12387–12391. 10.1002/anie.202004242. [DOI] [PubMed] [Google Scholar]; g Lindqvist M.; Borre K.; Axenov K.; Kótai B.; Nieger M.; Leskelä M.; Pápai I.; Repo T. Chiral Molecular Tweezers: Synthesis and Reactivity in Asymmetric Hydrogenation. J. Am. Chem. Soc. 2015, 137, 4038–4041. 10.1021/ja512658m. [DOI] [PubMed] [Google Scholar]; h Takeda M.; Hiroto S.; Yokoi H.; Lee S.; Kim D.; Shinokubo H. Azabuckybowl-Based Molecular Tweezers as C60 and C70 Receptors. J. Am. Chem. Soc. 2018, 140, 6336–6342. 10.1021/jacs.8b02327. [DOI] [PubMed] [Google Scholar]; i Doistau B.; Benda L.; Cantin J.-L.; Chamoreau L.-M.; Ruiz E.; Marvaud V.; Hasenknopf B.; Vives G. Six States Switching of Redox-Active Molecular Tweezers by Three Orthogonal Stimuli. J. Am. Chem. Soc. 2017, 139, 9213–9220. 10.1021/jacs.7b02945. [DOI] [PubMed] [Google Scholar]; j Mastalerz M. Single-Handed Towards Nanosized Organic Molecules. Angew. Chem., Int. Ed. 2016, 55, 45–47. 10.1002/anie.201509420. [DOI] [PubMed] [Google Scholar]; k Bier D.; Rose R.; Bravo-Rodriguez K.; Bartel M.; Ramirez-Anguita J. M.; Dutt S.; Wilch C.; Klärner F.-G.; Sanchez-Garcia E.; Schrader T.; Ottmann C. Molecular tweezers modulate 14–3-3 protein–protein interactions. Nat. Chem. 2013, 5, 234–239. 10.1038/nchem.1570. [DOI] [PubMed] [Google Scholar]
- a Dalgarno S. J.; Power N. P.; Atwood J. L. Metallo-supramolecular capsules. Coord. Chem. Rev. 2008, 252, 825–841. 10.1016/j.ccr.2007.10.010. [DOI] [Google Scholar]; b Catti L.; Zhang Q.; Tiefenbacher K. Advantages of Catalysis in Self-Assembled Molecular Capsules. Chem. - Eur. J. 2016, 22, 9060–9066. 10.1002/chem.201600726. [DOI] [PubMed] [Google Scholar]; c Ferrand Y.; Huc I. Designing Helical Molecular Capsules Based on Folded Aromatic Amide Oligomers. Acc. Chem. Res. 2018, 51, 970–977. 10.1021/acs.accounts.8b00075. [DOI] [PubMed] [Google Scholar]; d Jędrzejewska H.; Szumna A. Making a Right or Left Choice: Chiral Self-Sorting as a Tool for the Formation of Discrete Complex Structures. Chem. Rev. 2017, 117, 4863–4899. 10.1021/acs.chemrev.6b00745. [DOI] [PubMed] [Google Scholar]; e Szumna A. Inherently chiral concave molecules—from synthesis to applications. Chem. Soc. Rev. 2010, 39, 4274–4285. 10.1039/b919527k. [DOI] [PubMed] [Google Scholar]; f La Manna P.; Talotta C.; Floresta G.; De Rosa M.; Soriente A.; Rescifina A.; Gaeta C.; Neri P. Mild Friedel–Crafts Reactions inside a Hexameric Resorcinarene Capsule: C–Cl Bond Activation through Hydrogen Bonding to Bridging Water Molecules. Angew. Chem., Int. Ed. 2018, 57, 5423–5428. 10.1002/anie.201801642. [DOI] [PubMed] [Google Scholar]; g Riwar L.-J.; Trapp N.; Root K.; Zenobi R.; Diederich F. Supramolecular Capsules: Strong versus Weak Chalcogen Bonding. Angew. Chem., Int. Ed. 2018, 57, 17259–17264. 10.1002/anie.201812095. [DOI] [PubMed] [Google Scholar]; h Dumele O.; Trapp N.; Diederich F. Halogen Bonding Molecular Capsules. Angew. Chem., Int. Ed. 2015, 54, 12339–12344. 10.1002/anie.201502960. [DOI] [PubMed] [Google Scholar]; i Biros S. M.; Rebek J. Jr. Structure and binding properties of water-soluble cavitands and capsules. Chem. Soc. Rev. 2007, 36, 93–104. 10.1039/B508530F. [DOI] [PubMed] [Google Scholar]; j Yoshizawa M.; Catti L. Bent Anthracene Dimers as Versatile Building Blocks for Supramolecular Capsules. Acc. Chem. Res. 2019, 52, 2392–2404. 10.1021/acs.accounts.9b00301. [DOI] [PubMed] [Google Scholar]
- a Nurttila S. S.; Linnebank P. R.; Krachko T.; Reek J. N. H. Supramolecular Approaches To Control Activity and Selectivity in Hydroformylation Catalysis. ACS Catal. 2018, 8, 3469–3488. 10.1021/acscatal.8b00288. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Jans A. C. H.; Caumes X.; Reek J. N. H. Gold Catalysis in (Supra)Molecular Cages to Control Reactivity and Selectivity. ChemCatChem 2019, 11, 287–297. 10.1002/cctc.201801399. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Ahmad N.; Younus H. A.; Chughtai A. H.; Verpoort F. Metal–organic molecular cages: applications of biochemical implications. Chem. Soc. Rev. 2015, 44, 9–25. 10.1039/C4CS00222A. [DOI] [PubMed] [Google Scholar]; d Jin Y.; Yu C.; Denman R. J.; Zhang W. Recent advances in dynamic covalent chemistry. Chem. Soc. Rev. 2013, 42, 6634–6654. 10.1039/c3cs60044k. [DOI] [PubMed] [Google Scholar]; e McConnell A. J.; Wood C. S.; Neelakandan P. P.; Nitschke J. R. Stimuli-Responsive Metal–Ligand Assemblies. Chem. Rev. 2015, 115, 7729–7793. 10.1021/cr500632f. [DOI] [PubMed] [Google Scholar]; f Han M.; Engelhard D. M.; Clever G. H. Self-assembled coordination cages based on banana-shaped ligands. Chem. Soc. Rev. 2014, 43, 1848–1860. 10.1039/C3CS60473J. [DOI] [PubMed] [Google Scholar]; g Amouri H.; Desmarets C.; Moussa J. Confined Nanospaces in Metallocages: Guest Molecules, Weakly Encapsulated Anions, and Catalyst Sequestration. Chem. Rev. 2012, 112, 2015–2041. 10.1021/cr200345v. [DOI] [PubMed] [Google Scholar]; h Holst J. R.; Trewin A.; Cooper A. I. Porous organic molecules. Nat. Chem. 2010, 2, 915–920. 10.1038/nchem.873. [DOI] [PubMed] [Google Scholar]
- a Tröger J. Ueber einige mittelst nascirenden Formaldehydes entstehende Basen. J. Prakt. Chem. 1887, 36, 225–245. 10.1002/prac.18870360123. [DOI] [Google Scholar]; b Spielman M. A. The Structure of Troeger’s Base. J. Am. Chem. Soc. 1935, 57, 583–585. 10.1021/ja01306a060. [DOI] [Google Scholar]
- a Webb T. H.; Suh H.; Wilcox C. S. Chemistry of synthetic receptors and functional group arrays. 16. Enantioselective and diastereoselective molecular recognition of alicyclic substrates in aqueous media by a chiral, resolved synthetic receptor. J. Am. Chem. Soc. 1991, 113, 8554–8555. 10.1021/ja00022a070. [DOI] [Google Scholar]; b Adrian J. C.; Wilcox C. S. Chemistry of synthetic receptors and functional group arrays. 10. Orderly functional group dyads. Recognition of biotin and adenine derivatives by a new synthetic host. J. Am. Chem. Soc. 1989, 111, 8055–8057. 10.1021/ja00202a078. [DOI] [Google Scholar]; c Manjula A.; Nagarajan M. New supramolecular hosts: Synthesis and cation binding studies of novel Tröger’s base-crown ether composites. Tetrahedron 1997, 53, 11859–11868. 10.1016/S0040-4020(97)00759-X. [DOI] [Google Scholar]; d Yuan C.; Zhang Y.; Xi H.; Tao X. An acidic pH fluorescent probe based on Tröger’s base. RSC Adv. 2017, 7, 55577–55581. 10.1039/C7RA11228A. [DOI] [Google Scholar]; e Shanmugaraju S.; Dabadie C.; Byrne K.; Savyasachi A. J.; Umadevi D.; Schmitt W.; Kitchen J. A.; Gunnlaugsson T. A supramolecular Tröger’s base derived coordination zinc polymer for fluorescent sensing of phenolic-nitroaromatic explosives in water. Chem. Sci. 2017, 8, 1535–1546. 10.1039/C6SC04367D. [DOI] [PMC free article] [PubMed] [Google Scholar]; f Delente J. M.; Umadevi D.; Shanmugaraju S.; Kotova O.; Watson G. W.; Gunnlaugsson T. Aggregation induced emission (AIE) active 4-amino-1,8-naphthalimide-Tröger’s base for the selective sensing of chemical explosives in competitive aqueous media. Chem. Commun. 2020, 56, 2562–2565. 10.1039/C9CC08457F. [DOI] [PubMed] [Google Scholar]; g Yuan C.; Li J.; Xi H.; Li Y. A sensitive pyridine-containing turn-off fluorescent probe for pH detection. Mater. Lett. 2019, 236, 9–12. 10.1016/j.matlet.2018.10.060. [DOI] [Google Scholar]; h Aroche D. M. P.; Vargas J. P.; Nogara P. A.; da Silveira Santos F.; da Rocha J. B. T.; Lüdtke D. S.; Rodembusch F. S. Glycoconjugates Based on Supramolecular Tröger’s Base Scaffold: Synthesis, Photophysics, Docking, and BSA Association Study. ACS Omega 2019, 4, 13509–13519. 10.1021/acsomega.9b01857. [DOI] [PMC free article] [PubMed] [Google Scholar]; i Trupp L.; Bruttomesso A. C.; Vardé M.; Eliseeva S. V.; Ramírez J. A.; Petoud S.; Barja B. C. Innovative Multipodal Ligands Derived from Tröger’s Bases for the Sensitization of Lanthanide(III) Luminescence. Chem. - Eur. J. 2020, 26, 16900–16909. 10.1002/chem.202003524. [DOI] [PubMed] [Google Scholar]
- a Goldberg Y.; Alper H. Transition metal complexes of Tröger’s base and their catalytic activity for the hydrosilylation of alkynes. Tetrahedron Lett. 1995, 36, 369–372. 10.1016/0040-4039(94)02272-D. [DOI] [Google Scholar]; b Shen Y.-M.; Zhao M.-X.; Xu J.; Shi Y. An Amine-Promoted Aziridination of Chalcones. Angew. Chem. 2006, 118, 8173–8176. 10.1002/ange.200602586. [DOI] [PubMed] [Google Scholar]; c Harmata M.; Rayanil K.-O.; Barnes C. L. Sequential Alkylation of Tröger’s Base. An Approach to New Chiral Ligands. Supramol. Chem. 2006, 18, 581–586. 10.1080/10610270600853477. [DOI] [Google Scholar]; d Du X.; Sun Y.; Tan B.; Teng Q.; Yao X.; Su C.; Wang W. Tröger’s base-functionalised organic nanoporous polymer for heterogeneous catalysis. Chem. Commun. 2010, 46, 970–972. 10.1039/B920113K. [DOI] [PubMed] [Google Scholar]; e Cuenú F.; Abonia R.; Bolaños A.; Cabrera A. Synthesis, structural elucidation and catalytic activity toward a model Mizoroki–Heck C–C coupling reaction of the pyrazolic Tröger’s base Pd4Cl8(PzTB)2 complex. J. Organomet. Chem. 2011, 696, 1834–1839. 10.1016/j.jorganchem.2011.02.016. [DOI] [Google Scholar]; f Cabrero-Antonino J. R.; García T.; Rubio-Marqués P.; Vidal-Moya J. A.; Leyva-Pérez A.; Al-Deyab S. S.; Al-Resayes S. I.; Díaz U.; Corma A. Synthesis of Organic–Inorganic Hybrid Solids with Copper Complex Framework and Their Catalytic Activity for the S-Arylation and the Azide–Alkyne Cycloaddition Reactions. ACS Catal. 2011, 1, 147–158. 10.1021/cs100086y. [DOI] [Google Scholar]; g Cui Y.; Du J.; Liu Y.; Yu Y.; Wang S.; Pang H.; Liang Z.; Yu J. Design and synthesis of a multifunctional porous N-rich polymer containing s-triazine and Tröger’s base for CO2 adsorption, catalysis and sensing. Polym. Chem. 2018, 9, 2643–2649. 10.1039/C8PY00177D. [DOI] [Google Scholar]; h Yuan R.; Li M.-q.; Zhou H.; Sun Y.-w.; Liang Y.-n.; Xu H.; Wan Y.; Wu H. Four-component 1,4-addition Ugi reaction catalyzed by the Schiff base derived from Tröger’s base and BINOL. Tetrahedron Lett. 2020, 61, 152388. 10.1016/j.tetlet.2020.152388. [DOI] [Google Scholar]
- a Rúnarsson Ö. V.; Artacho J.; Wärnmark K. The 125th Anniversary of the Tröger’s Base Molecule: Synthesis and Applications of Tröger’s Base Analogues. Eur. J. Org. Chem. 2012, 2012, 7015–7041. 10.1002/ejoc.201201249. [DOI] [Google Scholar]; b Yashima E.; Akashi M.; Miyauchi N. Chiral Bis(1,10-phenanthroline) with Tröger’s Base Skeleton. Synthesis and Interaction with DNA. Chem. Lett. 1991, 20, 1017–1020. 10.1246/cl.1991.1017. [DOI] [Google Scholar]; c Tatibouet A.; Demeunynck M.; Andraud C.; Collet A.; Lhomme J. Synthesis and study of an acridine substituted Troger’s base: preferential binding of the (−)-isomer to B-DNA. Chem. Commun. 1999, 161–162. 10.1039/a808822e. [DOI] [Google Scholar]; d Baldeyrou B.; Tardy C.; Bailly C.; Colson P.; Houssier C.; Charmantray F.; Demeunynck M. Synthesis and DNA interaction of a mixed proflavine–phenanthroline Tröger base. Eur. J. Med. Chem. 2002, 37, 315–322. 10.1016/S0223-5234(02)01356-9. [DOI] [PubMed] [Google Scholar]; e Veale E. B.; Gunnlaugsson T. Synthesis, Photophysical, and DNA Binding Studies of Fluorescent Tröger’s Base Derived 4-Amino-1,8-naphthalimide Supramolecular Clefts. J. Org. Chem. 2010, 75, 5513–5525. 10.1021/jo1005697. [DOI] [PubMed] [Google Scholar]; f Gaslonde T.; Léonce S.; Pierré A.; Pfeiffer B.; Michel S.; Tillequin F. Tröger’s bases in the acronycine, benzo[a]acronycine, and benzo[b]acronycine series. Tetrahedron Lett. 2011, 52, 4426–4429. 10.1016/j.tetlet.2011.06.064. [DOI] [Google Scholar]; g Paul A.; Maji B.; Misra S. K.; Jain A. K.; Muniyappa K.; Bhattacharya S. Stabilization and Structural Alteration of the G-Quadruplex DNA Made from the Human Telomeric Repeat Mediated by Tröger’s Base Based Novel Benzimidazole Derivatives. J. Med. Chem. 2012, 55, 7460–7471. 10.1021/jm300442r. [DOI] [PubMed] [Google Scholar]; h Shanmugaraju S.; la Cour Poulsen B.; Arisa T.; Umadevi D.; Dalton H. L.; Hawes C. S.; Estalayo-Adrián S.; Savyasachi A. J.; Watson G. W.; Williams D. C.; Gunnlaugsson T. Synthesis, structural characterisation and antiproliferative activity of a new fluorescent 4-amino-1,8-naphthalimide Tröger’s base–Ru(ii)–curcumin organometallic conjugate. Chem. Commun. 2018, 54, 4120–4123. 10.1039/C8CC01584H. [DOI] [PubMed] [Google Scholar]; i Veale E. B.; Frimannsson D. O.; Lawler M.; Gunnlaugsson T. 4-Amino-1,8-naphthalimide-Based Tröger’s Bases As High Affinity DNA Targeting Fluorescent Supramolecular Scaffolds. Org. Lett. 2009, 11, 4040–4043. 10.1021/ol9013602. [DOI] [PubMed] [Google Scholar]; j Murphy S.; Bright S. A.; Poynton F. E.; McCabe T.; Kitchen J. A.; Veale E. B.; Williams D. C.; Gunnlaugsson T. Synthesis, photophysical and cytotoxicity evaluations of DNA targeting agents based on 3-amino-1,8-naphthalimide derived Tröger’s bases. Org. Biomol. Chem. 2014, 12, 6610–6623. 10.1039/C3OB42213E. [DOI] [PubMed] [Google Scholar]; k Zhao Y.; Chen K.; Yildiz E. A.; Li S.; Hou Y.; Zhang X.; Wang Z.; Zhao J.; Barbon A.; Yaglioglu H. G.; Wu H. Efficient Intersystem Crossing in the Tröger’s Base Derived From 4-Amino-1,8-naphthalimide and Application as a Potent Photodynamic Therapy Reagent. Chem. - Eur. J. 2020, 26, 3591–3599. 10.1002/chem.201905248. [DOI] [PubMed] [Google Scholar]
- a Frank K. E.; Aubé J. Lewis acid-mediated cyclizations of (2′-amino-N′-tert-butoxycarbonyl-benzylidene)-3-alkenylamines. Tetrahedron Lett. 1998, 39, 7239–7242. 10.1016/S0040-4039(98)01596-2. [DOI] [Google Scholar]; b Frank K. E.; Aubé J. Cyclizations of Substituted Benzylidene-3-alkenylamines: Synthesis of the Tricyclic Core of the Martinellines. J. Org. Chem. 2000, 65, 655–666. 10.1021/jo990843c. [DOI] [PubMed] [Google Scholar]; c Leganza A.; Bezze C.; Zonta C.; Fabris F.; De Lucchi O.; Linden A. Synthesis of 1,5-Substituted Iminodibenzo[b,f][1,5]diazocine, an Analogue of Tröger’s Base. Eur. J. Org. Chem. 2006, 2006, 2987–2990. 10.1002/ejoc.200600167. [DOI] [Google Scholar]; d Redshaw C.; Wood P. T.; Elsegood M. R. J. Synthesis and Structure of Pentamethylcyclopentadienyl Tungsten(V) Complexes Containing Functionalized 6,12-Epiiminodibenzo[b,f][1.5]diazocine Ligands. Organometallics 2007, 26, 6501–6504. 10.1021/om700810p. [DOI] [Google Scholar]; e Pettersson B.; Bergman J.; Svensson P. H. Synthetic studies towards 1,5-benzodiazocines. Tetrahedron 2013, 69, 2647–2654. 10.1016/j.tet.2013.01.018. [DOI] [Google Scholar]; f Mao D.; Tang J.; Wang W.; Wu S.; Liu X.; Yu J.; Wang L. Scandium Pentafluorobenzoate-Catalyzed Unexpected Cascade Reaction of 2-Aminobenzaldehydes with Primary Amines: A Process for the Preparation of Ring-Fused Aminals. J. Org. Chem. 2013, 78, 12848–12854. 10.1021/jo402331t. [DOI] [PubMed] [Google Scholar]; g Muthukrishnan I.; Karuppasamy M.; Nagarajan S.; Maheswari C. U.; Pace V.; Menéndez J. C.; Sridharan V. Synthesis of 6,12-Epiminodibenzo[b,f][1,5]diazocines via an Ytterbium Triflate-Catalyzed, AB2 Three-Component Reaction. J. Org. Chem. 2016, 81, 9687–9694. 10.1021/acs.joc.6b01764. [DOI] [PubMed] [Google Scholar]; h Su C.; Tandiana R.; Balapanuru J.; Tang W.; Pareek K.; Nai C. T.; Hayashi T.; Loh K. P. Tandem Catalysis of Amines Using Porous Graphene Oxide. J. Am. Chem. Soc. 2015, 137, 685–690. 10.1021/ja512470t. [DOI] [PubMed] [Google Scholar]; i Nakaike Y.; Yoshida Y.; Yokoyama S.; Ito A.; Nishiwaki N. Synthesis and intramolecular ring transformation of N,N′-dialkylated 2,6,9-triazabicyclo[3.3.1]nonadienes. Org. Biomol. Chem. 2020, 18, 9109–9116. 10.1039/D0OB01950J. [DOI] [PubMed] [Google Scholar]; j Chen Y.; Li S.; Hou S.; Xu J.; Yang Z. Chiral Synthesis of McGeachin-Type Bisaminals. J. Org. Chem. 2020, 85, 3709–3716. 10.1021/acs.joc.9b03362. [DOI] [PubMed] [Google Scholar]
- a Stefanovic G.; Lorenc L.; Mamuzić R. I.; Mihailović M. L. Anhydrobiisatic acid. Tetrahedron 1959, 6, 304–311. 10.1016/0040-4020(59)80009-0. [DOI] [Google Scholar]; b Albert A.; Yamamoto H. Quinazoline studies. Part XII. Action of acid and alkali on quinazoline. J. Chem. Soc. C 1968, 1944–1949. 10.1039/j39680001944. [DOI] [Google Scholar]; c Shidlovskii A. F.; Golubev A. S.; Gusev D. V.; Suponitsky K. Y.; Peregudov A. S.; Chkanikov N. D. A new synthesis of N-substituted o-trifluoroacetylanilines. J. Fluorine Chem. 2012, 143, 272–280. 10.1016/j.jfluchem.2012.07.002. [DOI] [Google Scholar]; d Zhang G.-W.; Meng W.; Ma H.; Nie J.; Zhang W.-Q.; Ma J.-A. Catalytic Enantioselective Alkynylation of Trifluoromethyl Ketones: Pronounced Metal Fluoride Effects and Implications of Zinc-to-Titanium Transmetallation. Angew. Chem., Int. Ed. 2011, 50, 3538–3542. 10.1002/anie.201007341. [DOI] [PubMed] [Google Scholar]; e Griffiths G. J.; Warm A. Proposed Mechanism for the Enantioselective Alkynylation of an Aryl Trifluoromethyl Ketone, the Key Step in the Synthesis of Efavirenz. Org. Process Res. Dev. 2016, 20, 803–813. 10.1021/acs.oprd.6b00058. [DOI] [Google Scholar]
- a Bernauer K. Über die Synthese einer Modellverbindung mit dem Chromophor der Toxiferine. 1. Mitteilung über Indole, Indolenine und Indoline. Helv. Chim. Acta 1963, 46, 197–210. 10.1002/hlca.19630460120. [DOI] [Google Scholar]; b Fritz H.; Rubach G. Synthesen in der Reihe der Indole und Indolalkaloide, VII1) Synthese eines undecacyclischen Di-indolin-diäthers vom Typ dimerer Calebassencurare-Alkaloide. Justus Liebigs Ann. Chem. 1968, 715, 135–145. 10.1002/jlac.19687150115. [DOI] [PubMed] [Google Scholar]
- a Butin K. P.; Kashin A. N.; Beletskaya I. P.; German L. S.; Polishchuk V. R. Acidities of some fluorine substituted C·H acids. J. Organomet. Chem. 1970, 25, 11–16. 10.1016/S0022-328X(00)86198-7. [DOI] [Google Scholar]; b Hine J.; Mahone L. G.; Liotta C. L. alpha.-Fluoro and.alpha.-alkoxy substituents as deactivators in carbanion formation. J. Am. Chem. Soc. 1967, 89, 5911–5920. 10.1021/ja00999a030. [DOI] [Google Scholar]
- Ishikawa T.Superbases for Organic Synthesis: Guanidines, Amidines, Phosphazenes and Related Organocatalysts; John Wiley & Sons: Hoboken, NJ, 2009; p 23. [Google Scholar]
- a Zafrani Y.; Yeffet D.; Sod-Moriah G.; Berliner A.; Amir D.; Marciano D.; Gershonov E.; Saphier S. Difluoromethyl Bioisostere: Examining the “Lipophilic Hydrogen Bond Donor” Concept. J. Med. Chem. 2017, 60, 797–804. 10.1021/acs.jmedchem.6b01691. [DOI] [PubMed] [Google Scholar]; b 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. 10.1021/cr4002879. [DOI] [PubMed] [Google Scholar]
- Meanwell N. A. Fluorine and Fluorinated Motifs in the Design and Application of Bioisosteres for Drug Design. J. Med. Chem. 2018, 61, 5822–5880. 10.1021/acs.jmedchem.7b01788. [DOI] [PubMed] [Google Scholar]
- Sessler C. D.; Rahm M.; Becker S.; Goldberg J. M.; Wang F.; Lippard S. J. CF2H, a Hydrogen Bond Donor. J. Am. Chem. Soc. 2017, 139, 9325–9332. 10.1021/jacs.7b04457. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Larson S. B.; Wilcox C. S. Structure of 5,11-methano-2,8-dimethyl-5,6,11,12-tetrahydrodibenzo[b,f][1,5]diazocine (Tröger’s base) at 163 K. Acta Crystallogr. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1986, 42, 224–227. 10.1107/S0108270186096713. [DOI] [Google Scholar]
- In all cases, diazocine 2a has been detected as a result of homocondenstaion (for details, see the Supporting Information). The respective homocondensataion of pyridine 3 has been suppressed under these conditions, and 3 was formed in marginal yield. In addition, a residual amount of unreacted pyridine 3 was also detected.
- Sowmiah S.; Esperança J. M. S. S.; Rebelo L. P. N.; Afonso C. A. M. Pyridinium salts: from synthesis to reactivity and applications. Org. Chem. Front. 2018, 5, 453–493. 10.1039/C7QO00836H. [DOI] [Google Scholar]
- a Kosiorek S.; Rosa B.; Boinski T.; Butkiewicz H.; Szymański M. P.; Danylyuk O.; Szumna A.; Sashuk V. Pillar[4]pyridinium: a square-shaped molecular box. Chem. Commun. 2017, 53, 13320–13323. 10.1039/C7CC08562A. [DOI] [PubMed] [Google Scholar]; b Guo Q.-H.; Zhou J.; Mao H.; Qiu Y.; Nguyen M. T.; Feng Y.; Liang J.; Shen D.; Li P.; Liu Z.; Wasielewski M. R.; Stoddart J. F. TetrazineBox: A Structurally Transformative Toolbox. J. Am. Chem. Soc. 2020, 142, 5419–5428. 10.1021/jacs.0c01114. [DOI] [PubMed] [Google Scholar]; c Cetin M. M.; Beldjoudi Y.; Roy I.; Anamimoghadam O.; Bae Y. J.; Young R. M.; Krzyaniak M. D.; Stern C. L.; Philp D.; Alsubaie F. M.; Wasielewski M. R.; Stoddart J. F. Combining Intra- and Intermolecular Charge Transfer with Polycationic Cyclophanes To Design 2D Tessellations. J. Am. Chem. Soc. 2019, 141, 18727–18739. 10.1021/jacs.9b07877. [DOI] [PubMed] [Google Scholar]; d Strutt N. L.; Zhang H.; Schneebeli S. T.; Stoddart J. F. Functionalizing Pillar[n]arenes. Acc. Chem. Res. 2014, 47, 2631–2642. 10.1021/ar500177d. [DOI] [PubMed] [Google Scholar]
- a Makosza M. Phase-transfer catalysis. A general green methodology in organic synthesis. Pure Appl. Chem. 2000, 72, 1399. 10.1351/pac200072071399. [DOI] [Google Scholar]; b Shirakawa S.; Maruoka K. Recent Developments in Asymmetric Phase-Transfer Reactions. Angew. Chem., Int. Ed. 2013, 52, 4312–4348. 10.1002/anie.201206835. [DOI] [PubMed] [Google Scholar]; c Qian D.; Sun J. Recent Progress in Asymmetric Ion-Pairing Catalysis with Ammonium Salts. Chem. - Eur. J. 2019, 25, 3740–3751. 10.1002/chem.201803752. [DOI] [PubMed] [Google Scholar]; d Krištofíková D.; Modrocká V.; Mečiarová M.; Šebesta R. Green Asymmetric Organocatalysis. ChemSusChem 2020, 13, 2828–2858. 10.1002/cssc.202000137. [DOI] [PubMed] [Google Scholar]; e Rössler S. L.; Jelier B. J.; Magnier E.; Dagousset G.; Carreira E. M.; Togni A. Pyridinium Salts as Redox-Active Functional Group Transfer Reagents. Angew. Chem., Int. Ed. 2020, 59, 9264–9280. 10.1002/anie.201911660. [DOI] [PubMed] [Google Scholar]
- Matos M. J.; Navo C. D.; Hakala T.; Ferhati X.; Guerreiro A.; Hartmann D.; Bernardim B.; Saar K. L.; Compañón I.; Corzana F.; Knowles T. P. J.; Jiménez-Osés G.; Bernardes G. J. L. Quaternization of Vinyl/Alkynyl Pyridine Enables Ultrafast Cysteine-Selective Protein Modification and Charge Modulation. Angew. Chem., Int. Ed. 2019, 58, 6640–6644. 10.1002/anie.201901405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- a Bruhn T.; Pescitelli G.; Jurinovich S.; Schaumlöffel A.; Witterauf F.; Ahrens J.; Bröring M.; Bringmann G. Axially Chiral BODIPY DYEmers: An Apparent Exception to the Exciton Chirality Rule. Angew. Chem., Int. Ed. 2014, 53, 14592–14595. 10.1002/anie.201408398. [DOI] [PubMed] [Google Scholar]; b Górecki M. A configurational and conformational study of (−)-Oseltamivir using a multi-chiroptical approach. Org. Biomol. Chem. 2015, 13, 2999–3010. 10.1039/C4OB02369B. [DOI] [PubMed] [Google Scholar]; c Rehman N. U.; Hussain H.; Al-Shidhani S.; Avula S. K.; Abbas G.; Anwar M. U.; Górecki M.; Pescitelli G.; Al-Harrasi A. Incensfuran: isolation, X-ray crystal structure and absolute configuration by means of chiroptical studies in solution and solid state. RSC Adv. 2017, 7, 42357–42362. 10.1039/C7RA07351H. [DOI] [Google Scholar]; d Pescitelli G.; Lüdeke S.; Chamayou A.-C.; Marolt M.; Justus V.; Górecki M.; Arrico L.; Di Bari L.; Islam M. A.; Gruber I.; Enamullah M.; Janiak C. Broad-Range Spectral Analysis for Chiral Metal Coordination Compounds: (Chiro)optical Superspectrum of Cobalt(II) Complexes. Inorg. Chem. 2018, 57, 13397–13408. 10.1021/acs.inorgchem.8b01932. [DOI] [PubMed] [Google Scholar]; e Kemper M.; Engelage E.; Merten C. Chiral molecular propellers of triarylborane ammonia adducts. Angew. Chem., Int. Ed. 2021, 60, 2958–2962. 10.1002/anie.202014130. [DOI] [PMC free article] [PubMed] [Google Scholar]; f Olszewska K.; Jastrzebska I.; Łapiński A.; Górecki M.; Santillan R.; Farfán N.; Runka T. Steroidal Molecular Rotors with 1,4-Diethynylphenylene Rotators: Experimental and Theoretical Investigations Toward Seeking Efficient Properties. J. Phys. Chem. B 2020, 124, 9625–9635. 10.1021/acs.jpcb.0c06464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Markó I. E.; Giles P. R.; Tsukazaki M.; Brown S. M.; Urch C. J. Copper-Catalyzed Oxidation of Alcohols to Aldehydes and Ketones: An Efficient, Aerobic Alternative. Science 1996, 274, 2044–2046. 10.1126/science.274.5295.2044. [DOI] [PubMed] [Google Scholar]
- Pierce M. E.; Parsons R. L.; Radesca L. A.; Lo Y. S.; Silverman S.; Moore J. R.; Islam Q.; Choudhury A.; Fortunak J. M. D.; Nguyen D.; Luo C.; Morgan S. J.; Davis W. P.; Confalone P. N.; Chen C.-y.; Tillyer R. D.; Frey L.; Tan L.; Xu F.; Zhao D.; Thompson A. S.; Corley E. G.; Grabowski E. J. J.; Reamer R.; Reider P. J. Practical Asymmetric Synthesis of Efavirenz (DMP 266), an HIV-1 Reverse Transcriptase Inhibitor. J. Org. Chem. 1998, 63, 8536–8543. 10.1021/jo981170l. [DOI] [Google Scholar]
- Czerwiński P.; Michalak M. NHC-Cu(I)-Catalyzed Friedländer-Type Annulation of Fluorinated o-Aminophenones with Alkynes on Water: Competitive Base-Catalyzed Dibenzo[b,f][1,5]diazocine Formation. J. Org. Chem. 2017, 82, 7980–7997. 10.1021/acs.joc.7b01235. [DOI] [PubMed] [Google Scholar]
- Holmquist E. F.; B. Keiding U.; Kold-Christensen R.; Salomón T.; Jørgensen K. A.; Kristensen P.; Poulsen T. B.; Johannsen M. ReactELISA: Monitoring a Carbon Nucleophilic Metabolite by ELISA—a Study of Lipid Metabolism. Anal. Chem. 2017, 89, 5066–5071. 10.1021/acs.analchem.7b00507. [DOI] [PubMed] [Google Scholar]
- Sun Y.-L.; Wei Y.; Shi M. Tunable regiodivergent phosphine-catalyzed [3 + 2] cycloaddition of alkynones and trifluoroacetyl phenylamides. Org. Chem. Front. 2017, 4, 2392–2402. 10.1039/C7QO00512A. [DOI] [Google Scholar]
- a Okada E.; Sakaemura T.; Shimomura N. A Simple Synthetic Method for 3-Trifluoroacetylated 4-Aminoquinolines from 4-Dimethylaminoquinoline by Novel Trifluoroacetylation and N-N Exchange Reactions. Chem. Lett. 2000, 29, 50–51. 10.1246/cl.2000.50. [DOI] [Google Scholar]; b Okada E.; Hatakenaka M.; Sakaemura T.; Shimomura N.; Ashida T. Simple syntheses of 3-trifluoroacetyl-4-quinolylamines, sulfides, and ethers starting from N,N-dimethyl-4-quinolylamine. Heterocycles 2012, 86, 1177–1185. 10.3987/COM-12-S(N)73. [DOI] [Google Scholar]
- Li L.; Deng X.-X.; Li Z.-L.; Du F.-S.; Li Z.-C. Multifunctional Photodegradable Polymers for Reactive Micropatterns. Macromolecules 2014, 47, 4660–4667. 10.1021/ma501019c. [DOI] [Google Scholar]
- Hanana M.; Arcostanzo H.; Das P. K.; Bouget M.; Le Gac S.; Okuno H.; Cornut R.; Jousselme B.; Dorcet V.; Boitrel B.; Campidelli S. Synergic effect on oxygen reduction reaction of strapped iron porphyrins polymerized around carbon nanotubes. New J. Chem. 2018, 42, 19749–19754. 10.1039/C8NJ04516J. [DOI] [Google Scholar]
- Altermann S. M.; Richardson R. D.; Page T. K.; Schmidt R. K.; Holland E.; Mohammed U.; Paradine S. M.; French A. N.; Richter C.; Bahar A. M.; Witulski B.; Wirth T. Catalytic Enantioselective α-Oxysulfonylation of Ketones Mediated by Iodoarenes. Eur. J. Org. Chem. 2008, 2008, 5315–5328. 10.1002/ejoc.200800741. [DOI] [Google Scholar]
- Baumann M.; Baxendale I. R.; Martin L. J.; Ley S. V. Development of fluorination methods using continuous-flow microreactors. Tetrahedron 2009, 65, 6611–6625. 10.1016/j.tet.2009.05.083. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







