Next Article in Journal
Metabolite Analysis of Alternaria Mycotoxins by LC-MS/MS and Multiple Tools
Next Article in Special Issue
Chiral Separation of Apremilast by Capillary Electrophoresis Using Succinyl-β-Cyclodextrin—Reversal of Enantiomer Elution Order by Cationic Capillary Coating
Previous Article in Journal
Benchmark Study on Phosphorescence Energies of Anthraquinone Compounds: Comparison between TDDFT and UDFT
Previous Article in Special Issue
A Density Functional Study on Ethylene Trimerization and Tetramerization Using Real Sasol Cr-PNP Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Elimination Pathways of 1-Methanesulfonyl-1,2-dihydroquinoline Sulfonamides

Department of Chemistry, Oklahoma State University, Stillwater, OK 74078-3071, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(7), 3256; https://doi.org/10.3390/molecules28073256
Submission received: 18 March 2023 / Revised: 2 April 2023 / Accepted: 3 April 2023 / Published: 5 April 2023
(This article belongs to the Special Issue Molecules in 2023)

Abstract

:
A series of new Morita–Baylis–Hillman acetates were prepared and reacted with methanesulfonamide (K2CO3, DMF, 23 °C) to produce tertiary dihydroquinoline sulfonamides in high yields. Subsequent efforts to eliminate the methylsulfonyl group from these derivatives (K2CO3, DMF, 90 °C) as a route to quinolines were met with mixed results. Although dihydroquinoline sulfonamides prepared from ethyl acrylate and acrylonitrile generally underwent elimination to give excellent yields of quinolines, those generated from 3-buten-2-one failed to undergo elimination and instead decomposed. The failure of these ketone substrates to aromatize presumably derives from the enolizable methyl ketone at C-3. Finally, the attempted aromatization of the acrylate-derived 6,7-difluoro-1,2-dihydroquinoline sulfonamide demonstrated that other interesting processes could occur in preference to the desired elimination.

Graphical Abstract

1. Introduction

We recently reported the synthesis of quinolines, naphthyridines and their dihydro derivatives from Morita–Baylis–Hillman (MBH) acetates [1,2]. In the current work, we hoped to develop another approach to these structures involving the reaction of sulfonamides with these same activated substrates. There is ample precedent showing that sulfonamides can be used in aza–Michael reactions to prepare simple adducts as well as a variety of heterocycles [3,4,5,6,7,8,9,10]. Additionally, we showed in the past that aromatic sulfonamide anions readily undergo SNAr additions to activated aromatic halides [11]. Our plan sought to utilize an aza–Michael-initiated SN2’-SNAr sequence between methanesulfonamide (pKa 17.5 in DMSO [12]) or p-toluenesulfonamide (pKa ca. 16.1 in DMSO [13]) and an MBH acetate to generate a tertiary dihydroquinoline sulfonamide. Subsequent elimination of the sulfonyl group would then yield substituted quinoline products. For the purpose of atom economy [14,15,16], the elimination of a methylsulfonyl (mesyl) group was deemed to be preferable for this reaction.
The original project aimed to add methanesulfonamide to MBH acetates 1 derived from benzaldehydes activated by electron withdrawing groups (EWG) toward SNAr cyclization via Michael-initiated SN2’ intermediate A and Meisenheimer complex B to give tertiary cyclic sulfonamides 2. In the second step, the plan was to eliminate the mesyl group to generate 3,6-disubstituted quinoline derivatives (Scheme 1). Such eliminations were first observed by Kim and co-workers [17] from 1-p-tosyl-1,2-dihydroquinolines derived from p-tosylamido MBH adducts. During our study, however, it was found that this process only proceeded using dihydroquinolines incorporating C-3 esters and nitriles but failed when a C-3 acetyl group was present.
A previous study by other researchers described a similar ring-forming process involving a SN2’-copper catalyzed coupling process using p-toluene- and p-nitrophenylsul-fonamides [18] with MBH acetates activated by an acrylate ester (Scheme 2). This procedure did not require SNAr-activating substituents on the aromatic ring of the MBH acetate, but instead, it coupled the initially added sulfonamide nitrogen to a halogenated carbon at the ortho position. This method yielded a diverse selection of substituted dihydroquinolines that were eliminated (4 equiv. DBU, CH3CN, 80 °C) to give ethyl 3-quinolinecarboxylates. Though this sequence shows a broader scope than the current work, reactions involving metal-promoted couplings can often be difficult to purify.
The formation of sulfonamides as intermediates in the current reaction is also of interest, as this family of compounds constitutes an important class of drug compounds with wide-ranging activities. Several experimental drug compounds are pictured in Figure 1. The tetrahydroquinoline sulfonamide 7 was found to impede the development and progression of several cancers by inhibiting the antagonist activity of myeloid cell leukemia-1 toward pro-apoptotic Bcl-2 proteins [19]. Phenylpyrrolidine sulfonamide 8 in conjunction with Doxorubicin showed antiproliferative activity against breast cancer cells both in vitro and in xenograft models [20]. Indole sulfonamide 9 inhibited the carbonic anhydrase IX isoform, possessed anti-proliferitive potential, and significantly increased apoptosis in breast cancer cells [21]. Coumarin-proline sulfonamide 10 proved to be active against breast cancer cells, and its inhibition of dipeptidal peptidase IV suggested that compound hybrids incorporating this functional triad might have potential in the treatment of type II diabetes [22]. Heteroaryl sulfonamide 11 exhibited neuroprotective effects via the inhibition of human carbonic anhydrase I and II, which may yield a new strategy for the treatment of Alzheimer’s disease [23]. Finally, pyrrolidinediol 12 proved effective in reversing pulmonary edema and increasing arterial oxygenation, lowering the chances of death in congestive heart failure patients through antagonism of the transient receptor potential vallinoid-4 (TRPV4) cation channel [24]. Although most sulfonamides are prepared via reaction of a sulfonyl chloride with a primary or secondary amine or alkylation of a pre-existing sulfonamide [25], the scheme outlined in this report represents a unique variation on double alkylation of a primary sulfonamide.

2. Results and Discussion

The MBH acetates were prepared as previously described [1,2]. The reaction of 2-fluoro-C5-activated benzaldehydes with ethyl acrylate, acrylonitrile, or 3-buten-2-one in the presence of DABCO in dry CH3CN gave the MBH alcohols, which were not isolated. Subsequent treatment of the adducts with acetic anhydride in the presence of catalytic trimethylsilyl trifluoromethanesulfonate [26] gave acetates 1327 for the current study (Table 1). These substrates were spectroscopically pure and used directly without further purification.
Our results are summarized in Table 2. For most substrates, the initial addition of methanesulfonamide (and in one case, of p-toluenesulfonamide) with ester and ketone substrates (1:1:2 methanesulfonamide: MBH acetate: K2CO3, DMF, 23 °C, 2 h) occurred to produce 1-methanesulfonyl-1,2-dihydroquinolines in high yields. Additionally, the p-toluenesulfonyl derivative 28-Ts was prepared in similar yields to demonstrate that aromatic sulfonamides could also be used in the ring closure and elimination protocols. Acrylonitrile substrates 3537, however, failed to undergo this reaction, presumably due to the diminished activation of the SNAr acceptor ring by CN, CF3 and Cl relative to NO2. Attempts to isolate the intermediate Michael adducts were thwarted by the presence of multiple products, and prolonged heating led to eventual decomposition.
Our subsequent attempts to perform elimination reactions to produce quinoline derivatives met with varied results (Table 3). With one exception, eliminations from acrylate-derived substrates were successful in producing the targeted quinolines in high yields (1 equiv. K2CO3, DMF, 90 °C, 1–2 h). Additionally, the nitro-substituted acrylonitrile adduct gave excellent conversion to the heteroaromatic product, whereas other substrates incorporating methyl ketones decomposed under these conditions. We note that attempts to perform the elimination with DBU as reported earlier [18] also failed for these substrates. Apparently, this aromatization becomes problematic when enolizable protons reside on the C-3 substituent of the sulfonyl-bearing ring. Eliminations to form naphthyridines were also unsuccessful. Thus, the limitations of the reaction appear to be reasonably well-defined.
To further explore functionalization of these structures, we wished to evaluate 6,7-difluorodihydroquinoline 32 in a possible extended SNAr reaction [27] through the fused-ring system. Although the carboxylic ester is a weak activator for SNAr reactions, our initial expectation was that the C-7 fluorine might be displaced to give the phenethylamino residue at this position. A similar substitution is a key step in the synthesis of fluoroquinolone antibiotics such as ciprofloxacin [28]. When the 6,7-difluoro substrate was reacted with phenethylamine, however, an unexpected product was observed in 95% yield. 1H NMR indicated that phenethylamine had been incorporated into the product. Additionally, the nitrogen-containing ring had been aromatized, and the methyl singlet of the mesyl in the starting material, though shifted from δ 2.71 in 32 to δ 2.90, was still present. Finally, the lack of 19F-coupled signals in the 13C NMR confirmed that the product contained no fluorine. An X-ray structure was obtained to unequivocally reveal the positions of the various groups. Interestingly, the structure proved to be the 1,3-sulfonyl migrated amine 49 (Figure 2), the C-6 fluorine having been replaced by a phenethylamino group and the C-7 fluorine replaced by a mesyl group. Reference to the literature indicated that 1,3-sulfonyl migrations are common but usually catalyzed by transition metals or Lewis acids [29]. Only one example was found where sulfonyl migration occurred under conditions similar to those reported here, and this involved a base-catalyzed aza–Claisen rearrangement of an N-arylsulfonyl propargyl vinylamine [30]. In the current work, the rearrangement was performed using 1:1:1 phenethylamine: 32: K2CO3 in DMF at 90 °C; no transition metals or Lewis acids were present.
Two mechanisms were considered to rationalize the observed transformation. Mechanism 1 required SNAr addition–elimination through two rings, whereas mechanism 2 avoided this extended SNAr process. In mechanism 1 (Scheme 3), phenethylamine would attack the C-6 position with the facile elimination of the mesyl group to give C. This process is critical to the current transformation, as the N-benzyl analogue of 32 was inert to these same conditions. Deprotonation by base would subsequently aromatize this intermediate to give D. At this point, the mesyl anion could attack at C-7 in a SNAr-type addition to give E. Final elimination of the C-7 fluoride would then be expected to yield the observed product, 49. The weakness of this mechanism is that the ethoxycarbonyl group does not provide sufficient activation for the extended SNAr reaction. Additionally, the phenethylamino group at C-6 would be expected to deactivate the ring toward this addition process.
Mechanism 2 (Scheme 4) begins with the formation of intermediate C as described previously. In the subsequent step, however, the sulfonyl anion would add to C-7 to give the imine stabilized anion F, which could eliminate the C-7 fluoride to give intermediate G. Finally, the base-promoted rearomatization of the system would afford product 49. Although both mechanisms traverse several high energy intermediates, we tend to favor this pathway, as it appears to have better electronic features than mechanism 1. At this point, however, we cannot definitively state which mechanism is operating.

3. Methods and Materials

3.1. General Methods

All reactions were performed under dry N2 in oven-dried glassware. All reagents and solvents were used as received. All wash solutions in work-up procedures were aqueous. Reactions were monitored with thin layer chromatography on Analtech No 21521 silica gel GF plates (Newark, DE, USA). Preparative separations were performed either by using flash chromatography on silica gel (Davisil®, grade 62, 60–200 mesh) containing 0.5% of UV-05 UV-active phosphor (both from Sorbent Technologies, Norcross, GA, USA) slurry packed into quartz columns or by using preparative thin layer chromatography (PTLC) on Analtech No 02,015 silica gel GF plates (Newark, DE, USA). Band elution for all chromatographic separations was monitored using a handheld UV lamp (Fisher Scientific, Pittsburgh, PA, USA). Melting points were obtained using a MEL-TEMP apparatus (Cambridge, MA, USA) and are uncorrected. FT-IR spectra were run as thin films on NaCl disks using a Nicolet iS50 spectrophotometer (Madison, WI, USA). 1H- and 13C-NMR spectra were measured using a Bruker Avance 400 system (Billerica, MA, USA) at 400 MHz and 101 MHz, respectively, using the indicated solvents containing 0.05% (CH3)4Si as the internal standard; coupling constants (J) are given in Hz. Low-resolution mass spectra were obtained using a Hewlett-Packard Model 1800A GCD GC-MS system (Palo Alto, CA, USA). Elemental analyses (±0.4%) were determined for all new compounds by Atlantic Microlabs (Norcross, GA, USA).

3.2. General Procedure for the Preparation of MBH Acetates

To a stirred solution of DABCO (1 equiv.) in CH3CN at room temperature, ethyl acrylate, acrylonitrile or 3-buten-2-one (1 equiv.) was added. The aldehyde was then added, and stirring continued until TLC analysis (10% EtOAc/hexane) indicated that reaction was complete (12–48 h). The crude reaction mixture was poured into water (30 mL) and extracted with EtOAc (3 × 25 mL). The combined organic layers were washed with 1 M HCl (2 × 25 mL), saturated NaHCO3 (30 mL) and saturated NaCl (30 mL) and then dried (Na2SO4). Concentration under vacuum gave the MBH alcohol, which was used without further purification.
The general procedure of Procopiou and co-workers was used [26]. A solution of the MBH alcohol (1 mmol) in CH2Cl2 (2 mL) was treated with acetic anhydride (1.5 mmol) at 0 °C, followed by a solution of TMSOTf (1.0 M in CH2Cl2, 20 µL). The reaction mixture was stirred for 1 h, after which TLC (10% EtOAc/hexane) indicated the reaction was complete. A stoichiometric quantity of methanol (142 µL) was added to quench the excess acetic anhydride, and the crude reaction mixture was washed with saturated NaHCO3 (30 mL). The two phases were separated, and the aqueous layer was further extracted with CH2Cl2 (2 × 25 mL). The combined organic extracts were washed with saturated NaHCO3 (30 mL), water (30 mL) and saturated NaCl (30 mL) and then dried (Na2SO4). The solvent was removed under vacuum to afford the spectroscopically pure MBH acetate, which was used directly for the preparation of 1-methylsulfonyl-1,2-dihydroquinolines. (±)-Ethyl 2-(acetoxy(2-fluoro-5-nitrophenyl)methyl)acrylate (13), (±)-ethyl 2-(acetoxy(5-cyano-2-fluorophenyl)methyl)acrylate (14), (±)-ethyl 2-(acetoxy(2-fluoropyridin-3-yl)methyl)acrylate (18), (±)-2-cyano-1-(2-fluoro-5-nitrophenyl)allyl acetate (19) and (±)-2-cyano-1-(5-cyano-2-fluorophenyl)allyl acetate (20) were prepared by using this procedure and were characterized previously [2].

3.2.1. (±)-Ethyl 2-(acetoxy(2-fluoro-5-(trifluoromethyl)phenyl)methyl)acrylate (15)

Yield: 2.10 g (92%) as a yellow oil; IR: 1756, 1724, 1620 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.64 (dd, J = 6.5, 2.5 Hz, 1H), 7.59 (m, 1H), 7.20 (t, J = 9.0 Hz, 1H), 6.96 (s, 1H), 6.52 (s, 1H), 5.91 (s, 1H), 4.17 (q, J = 7.1 Hz, 2H), 2.14 (s, 3H), 1.24 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 169.1, 146.4, 162.1 (d, J = 256.0 Hz), 137.9, 127.7–127.4 (complex), 127.0, 126.6, 126.6–126.5 (complex), 123.6 (q, J = 272.1 Hz), 119.5, 116.6, 116.3, 66.9, 20.7, 13.8; MS (m/z): 334 (M); Anal. Calcd for C15H14F4O4: C, 53.90; H, 4.22. Found: C, 53.81; H, 4.18.

3.2.2. (±)-Ethyl 2-(acetoxy(5-chloro-2-fluorophenyl)methyl)acrylate (16)

Yield: 2.16 g (93%) as a yellow oil; IR: 1753, 1725, 1637 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.30 (m, 1H), 7.25 (m, 1H), 7.00 (t, J = 8.9 Hz, 1H), 6.87 (s, 1H), 6.48 (s, 1H), 5.85 (s, 1H), 4.17 (q, J = 7.1 Hz, 2H), 2.13 (s, 3H), 1.24 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 169.0, 164.4, 158.8 (d, J = 250.5 Hz), 138.1, 130.0 (d, J = 8.5 Hz), 129.2 (d, J = 3.4 Hz), 128.8 (d, J = 3.6 Hz), 127.1 (d, J = 15.1 Hz), 126.9, 117.1 (d, J = 23.6 Hz), 66.9 (d, J = 2.9 Hz), 61.2, 20.8, 13.9; MS (m/z): 300 (M+); Anal. Calcd for C14H14ClFO4: C, 55.92; H, 4.69. Found: C, 55.84; H, 4.66.

3.2.3. (±)-Ethyl 2-(acetoxy(2,4,5-trifluorophenyl)methyl)acrylate (17)

Yield: 2.13 g (92%) as a yellow oil; IR: 1757, 1728, 1634 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.17 (ddd, J = 15.1, 8.8, 6.5 Hz, 1H), 6.94 (td, J = 10.3, 6.5 Hz, 1H), 6.84 (s, 1H), 6.48 (s, 1H), 5.88 (s, 1H), 4.18 (q, J = 7.1 Hz, 2H), 2.13 (d, J = 1.4 Hz, 3H), 1.25 (td, J = 7.1, 1.4 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 169.0, 164.3, 155.4 (dd, J = 249.4, 2.7 Hz), 150.0 (ddd, J = 252.5, 14.4, 12.2 Hz), 147.2 (dd, J = 245.4, 3.7 Hz), 138.0, 126.6, 121.9 (dt, J = 15.8, 4.7 Hz), 116.7 (ddd, J = 20.2, 4.7, 1.6 Hz), 105.8 (dd, J = 27.9, 20.9 Hz), 66.6 (d, J = 2.0 Hz), 61.2, 20.7, 13.9: MS (m/z): 302 (M+); Anal. Calcd for C14H13F3O4: C, 55.63; H, 4.34. Found: C, 55.68; H, 4.39.

3.2.4. (±)-2-Cyano-1-(2-fluoro-5-(trifluoromethyl)phenyl)allyl Acetate (21)

Yield: 1.09 g (93%) as a white solid, m.p. 35–36 °C; IR: 2229, 1761, 1626 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.79 (dd, J = 6.5, 2.3 Hz, 1H), 7.67 (m, 1H), 7.24 (t, J = 9.0 Hz, 1H), 6.64 (s, 1H), 6.15 (d, J = 1.0 Hz, 1H), 6.12 (t, J = 1.0 Hz, 1H), 2.23 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 168.9, 161.5 (d, J = 255.5 Hz), 133.3, 128.4 (dq, J = 9.6, 3.7 Hz), 127.6 (qd, J = 33.4, 3.7 Hz), 125.5 (apparent pentet, J = 3.9 Hz), 124.5 (d, J = 14.0 Hz), 123.4 (q, J = 272.2 Hz), 121.3, 116.7 (d, J = 22.5 Hz), 115.4, 68.0 (d, J = 3.3 Hz), 20.8; MS (m/z): 287 (M+); Anal. Calcd for C13H9F4NO2: C, 54.36; H, 3.16; N, 4.88. Found: C, 54.41; H, 3.11; N, 4.83.

3.2.5. (±)-1-(5-Chloro-2-fluorophenyl)-2-cyanoallyl Acetate (22)

Yield: 1.14 g (95%) as a white solid, m.p. 45–46 °C; IR: 2231, 1756, 1628 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.48 (dd, J = 6.2, 2.6 Hz, 1H), 7.34 (ddd, J = 8.8, 4.5, 2.6 Hz, 1H), 7.05 (t, J = 9.1 Hz, 1H), 6.57 (s, 1H), 6.13 (d, J = 1.0 Hz, 1H), 6.09 (t, J = 1.0 Hz, 1H), 2.22 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 168.9, 158.2 (d, J = 249.1 Hz), 133.1, 130.9 (d, J = 8.5 Hz), 130.1 (d, J = 3.5 Hz), 127.8 (d, J = 3.1 Hz), 125.0 (d, J = 14.6 Hz), 121.4, 117.3 (d, J = 23.0 Hz), 115.6 68.0 (d, J = 3.3 Hz), 20.8; MS (m/z): 253 (M+); Anal. Calcd for C12H9ClFNO2: C, 56.82; H, 3.58; N, 5.52. Found: C, 56.70; H, 3.53; N, 5.42.

3.2.6. (±)-1-(2-Fluoro-5-nitrophenyl)-2-methylene-3-oxobutyl Acetate (23)

Yield: 2.68 g (96%) as a light yellow solid, m.p. 72–73 °C; IR: 1752, 1679, 1636 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.22–8.18 (complex, 2H), 7.21 (t, J = 9.5 Hz, 1H), 6.93 (s, 1H), 6.32 (d, J = 0.8 Hz, 1H), 6.02 (d, J = 1.4 Hz, 1H), 2.37 (s, 3H), 2.14 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 196.9, 169.0, 163.8 (d, J = 261.8 Hz), 145.3, 144.2 (d, J = 3.0 Hz), 127.7 (d, J = 15.5 Hz), 126.3, 125.9 (d, J = 10.4 Hz), 125.0 (d, J = 5.5 Hz), 116.9 (d, J = 24.5 Hz), 66.3 (d, J = 2.9 Hz), 26.0, 20.8; MS (m/z): 281 (M+); Anal. Calcd for C13H12FNO5: C, 55.52; H, 4.30; N, 4.98. Found: C, 55.46; H, 4.22; N, 4.91.

3.2.7. (±)-1-(5-Cyano-2-fluorophenyl)-2-methylene-3-oxobutyl Acetate (24)

Yield: 1.17 g (98%) as a white solid, m.p. 104–105 °C; IR: 2230, 1754, 1681, 1627 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.64 (dd, J = 7.0, 2.1 Hz, 1H), 7.61 (m, 1H), 7.16 (t, J = 9.0 Hz, 1H), 6.89 (s, 1H), 6.35 (s, 1H), 6.15 (d, J = 1.4 Hz, 1H), 2.36 (s, 3H), 2.13 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 196.8, 169.0, 162.7 (d, J = 260.4 Hz), 145.4, 134.3 (d, J = 9.7 Hz), 133.6 (d, J = 4.9 Hz), 127.8 (d, J = 14.7 Hz), 127.1, 117.9, 117.3 (d, J = 23.4 Hz), 108.6 (d, J = 4.0 Hz), 66.4 (d, J = 3.0 Hz), 26.0, 10.9; HRMS (m/z): 261 (M+); Anal. Calcd for C14H12FNO3: C, 64.36; H, 4.63; N, 5.36. Found: C, 64.29; H, 4.58; N, 5.27.

3.2.8. (±)-1-(2-Fluoro-5-(trifluoromethyl)phenyl)-2-methylene-3-oxobutyl Acetate (25)

Yield: 3.19 g (95%) as a yellow oil; IR: 1749, 1684, 1631 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.57 (overlapping d, J = 5.3 Hz, 1H and m, 1H), 7.18 (t, J = 9.3 Hz, 1H), 6.96 (s, 1H), 6.35 (s, 1H), 6.12 (s, 1H), 2.35 (s, 3H), 2.12 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 196.9, 169.1, 162.1 (dd, J = 255.7, 1.6 Hz), 145.7, 127.5 (m, 2C), 126.9 (d, J = 2.0 Hz), 126.8–126.4 (complex, 2C), 123.6 (q, J = 271.9 Hz), 116.5 (d, J = 23.2 Hz), 66.7 (d, J = 2.9 Hz), 26.0, 20.8; MS (m/z): 304 (M); Anal. Calcd for C14H12F4O3: C, 55.27; H, 3.98. Found: C, 55.21; H, 3.96.

3.2.9. (±)-2-Methylene-3-oxo-1-(2,4,5-trifluorophenyl)butyl Acetate (26)

Yield: 2.98 g (97%) as a white solid, m.p. 82–84 °C; IR: 1754, 1680, 1631 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.14 (ddd, J = 15.2, 8.7, 6.5 Hz, 1H), 6.94 (td, J = 9.7, 6.5 Hz, 1H), 6.84 (s, 1H), 6.33 (s, 1H), 6.12 (s, 1H), 2.35 (s, 3H), 2.11 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 196.9, 169.1, 155.5 (ddd, J = 249.4, 9.5, 2.7 Hz), 150.0 (ddd, J = 252.3, 14.3, 12.2 Hz), 146.7 (ddd, J = 245.1, 12.7, 3.7 Hz), 146.5, 126.7 (d, J = 1.2 Hz), 122.1 (dt, J = 15.6, 4.7 Hz), 117.0 (ddd, J = 20.2, 5.1, 1.5 Hz), 105.9 (dd, J = 27.7, 20.9 Hz), 66.5 (d, J = 2.4 Hz), 26.0, 20.8; MS (m/z): 272 (M+); Anal. Calcd for C13H11F3O3: C, 57.36; H, 4.07. Found: C, 57.31; H, 3.98.

3.2.10. (±)-1-(2-Fluoropyridin-3-yl)-2-methylene-3-oxobutyl Acetate (27)

Yield: 2.16 g (89%) as a yellow oil; IR: 1749, 1679, 1638 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.15 (dm, J = 4.9 Hz, 1H), 7.82 (ddd, J = 9.5, 7.4, 2.0 Hz, 1H), 7.19 (ddd, J = 6.6, 4.9, 1.6 Hz, 1H), 6.81 (s, 1H), 6.37 (s, 1H), 6.19 (s, 1H), 2.35 (s, 3H), 2.12 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 197.0, 169.2, 160.9 (d, J = 241.9 Hz), 147.4 (d, J = 14.8 Hz), 145.1 140.3 (J = 4.4 Hz), 127.3 (d, J = 1.7 Hz), 121.4 (d, J = 4.5 Hz), 120.6 (d, J = 27.5 Hz), 67.7 (d, J = 2.8 Hz), 26.0, 20.8; MS (m/z): 237 (M+); Anal. Calcd for C12H12FNO3: C, 60.76; H, 5.10; N, 5.90. Found: C, 60.69; H, 5.07; N, 5.81.

3.3. General Procedure for the Preparation of 1-Methanesulfonyl-1,2-Dihydroquinolines from MBH Acetates

To a mixture of methane- or p-toluenesulfonamide (1 mmol) and K2CO3 (2 mmol) in anhydrous DMF (1 mL), the MBH acetate (1 mmol) was added in DMF (1 mL). The mixture was stirred at room temperature for 2 h, at which time TLC (20% EtOAc/hexane) indicated the complete consumption of the starting material. The crude reaction mixture was poured into water (15 mL) and extracted with EtOAc (3 × 25 mL). The combined organic layers were washed with 1 M HCl (2 × 25 mL), saturated NaHCO3 (2 × 25 mL) and saturated NaCl (1 × 30 mL) and then dried (Na2SO4). Removal of the solvent under vacuum gave a crude product that was further purified via column chromatography to afford the following products.

3.3.1. Ethyl 1-(Methylsulfonyl)-6-nitro-1,2-dihydroquinoline-3-carboxylate (28-Ms)

Yield: 2.01 g (96%) as a white solid, m.p. 163–164 °C; IR: 1704, 1642, 1529, 1352, 1160 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.24–8.17 (complex, 2H), 7.84 (d, J = 8.8 Hz, 1H), 7.60 (s, 1H), 4.75 (d, J = 1.5 Hz, 2H), 4.35 (q, J = 7.1 Hz, 2H), 2.85 (s, 3H), 1.39 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 163.7, 145.4, 141.6, 132.4, 128.6, 127.6, 125.5, 125.4, 124.0, 61.8, 44.5, 39.2, 14.2; MS (m/z): 326 (M); Anal. Calcd for C13H14N2O6S: C, 47.85; H, 4.32; N, 8.58. Found: C, 47.92; H, 4.34; N, 8.51.

3.3.2. Ethyl 6-Nitro-1-tosyl-1,2-dihydroquinoline-3-carboxylate (28-Ts)

Yield: 2.35 g (91%) as a white solid, m.p. 142–143 °C (lit. [18] m.p. 140–142 °C); IR 1708, 1524, 1349, 1167 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.23 (dd, J = 9.0, 2.6 Hz, 1H), 8.02 (d, J = 2.6 Hz, 1H), 7.93 (d, J = 9.0 Hz, 1H), 7.36 (d, J = 8.3 Hz, 2H), 7.13 (d, J = 8.3 Hz, 2H), 7.07 (s, 1H), 4.75 (s, 2H), 4.26 (q, J = 7.1 Hz, 2H), 2.36 (s, 3H), 1.34 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 163.5, 145.6, 144.7, 141.7, 135.6, 131.5, 129.7, 128.3, 127.8, 127.0, 126.9, 125.0, 123.5, 61.4, 44.3, 21.6, 14.3; MS (m/z): 402 (M+).

3.3.3. Ethyl 6-Cyano-1-(methylsulfonyl)-1,2-dihydroquinoline-3-carbonitrile (29)

Yield: 0.49 g (93%) as a white solid, m.p. 103–104 °C; IR: 2226, 1707, 1645, 1348, 1151 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.79 (d, J = 8.5 Hz, 1H), 7.65 (dd, J = 8.4, 2.0 Hz, 1H), 7.60 (d, J = 2.0 Hz, 1H), 7.51 (s, 1H), 4.72 (d, J = 1.4 Hz, 2H), 4.34 (q, J = 7.1 Hz, 2H), 2.80 (s, 3H), 1.38 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 163.8, 140.0, 133.9, 132.5, 132.2, 128.5, 127.9, 125.8, 117.7, 110.2, 61.8, 44.4, 39.0, 14.2; MS (m/z): 306 (M+); Anal. Calcd for C14H14N2O4S: C, 54.89; H, 4.61; N, 9.14. Found: C, 54.84; H, 4.57; N, 9.03.

3.3.4. Ethyl 1-(Methylsulfonyl)-6-(trifluoromethyl)-1,2-dihydroquinoline-3-carboxylate (30)

Yield: 0.40 g (96%) as a white solid, m.p. 99–101 °C; IR: 1714, 1648, 1339, 1130 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.80 (d, J = 8.5 Hz, 1H), 7.63 (dd, J = 8.5, 2.2 Hz, 1H), 7.58 (two coincident signals, apparent s, 2H), 4.72 (d, J = 1.4 Hz, 2H), 4.33 (q, J = 7.1 Hz, 2H), 2.76 (s, 3H), 1.37 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 164.0, 139.1, 133.0, 128.8 (q, J = 33.1 Hz), 128.0, 127.6, 127.5 (q, J = 3.6 Hz), 126.0, 125.9 (q, J = 3.8 Hz), 123.5 (q, J = 272.2 Hz), 61.6, 44.4, 38.7, 14.2; MS (m/z): 349 (M+); Anal. Calcd for C14H14F3NO4S: C, 48.14; H, 4.04; N, 4.01. Found: C, 48.09; H, 3.99; N, 3.92.

3.3.5. Ethyl 6-Chloro-1-(methylsulfonyl)-1,2-dihydroquinoline-3-carboxylate (31)

Yield: 0.38 g (91%) as a white solid, m.p. 109–111 °C; IR 1711, 1359, 1156 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.61 (d, J = 8.6 Hz, 1H), 7.50 (s, 1H), 7.35 (dd, J = 8.6, 2.4 Hz, 1H), 7.32 (d, J = 2.4 Hz, 1H), 4.67 (d, J = 1.1 Hz, 2H), 4.32 (q, J = 7.1 Hz, 2H), 2.69 (s, 3H), 1.37 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 164.2, 134.5, 133.1, 132.5, 130.6, 128.8, 128.5, 127.9, 127.6, 61.6, 44.5, 38.1, 14.2; MS (m/z): 315. (M+); Anal. Calcd for C13H14ClNO4S: C, 49.45; H, 4.47; N, 4.44. Found: C, 49.41; H, 4.46; N, 4.35.

3.3.6. Ethyl 6,7-Difluoro-1-(methylsulfonyl)-1,2-dihydroquinoline-3-carboxylate (32)

Yield: 0.38 g (93%) as a white solid, m.p. 145–147 °C; IR: 1718, 1659, 1354, 1155 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.55 (dd, J = 11.2, 7.3 Hz, 1H), 7.47 (s, 1H), 7.15 (dd, J = 9.8, 8.2 Hz, 1H), 4.66 (d, J = 1.4 Hz, 2H), 4.35 (q, J = 7.1 Hz, 2H), 2.71 (s, 3H), 1.37 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 164.1, 150.8 (dd, J = 254.3, 13.1 Hz), 149.0 (dd, J = 250.0, 13.3 Hz), 132.6 (dd, J = 9.1, 3.3 Hz), 132.5 (t, J = 2.3 Hz), 127.0 (d, J = 2.6 Hz), 124.2 (dd, J = 6.6, 3.8 Hz), 116.7 (dd, J = 18.5, 1.5 Hz), 116.4 (d, J = 21.3 Hz), 61.6, 44.2, 38.2, 14.2; MS (m/z): 317 (M); Anal. Calcd for C13H13F2NO4S: C, 49.21; H, 4.13; N, 4.41. Found: C, 49.12; H, 4.08; N, 4.33.

3.3.7. Ethyl 1-(Methylsulfonyl)-1,2-dihydro-1,8-naphthyridine-3-carboxylate (33)

Yield: 1.94 g (92%) as a white solid, m.p. 105–106 °C; IR: 1704, 1338, 1158 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.28 (dd, J = 5.0, 1.8 Hz, 1H), 7.50 (dd, J = 7.5, 1.8 Hz, 1H), 7.39 (t, J = 1.6 Hz, 1H), 7.02 (dd, J = 7.5, 5.0 Hz, 1H), 4.84 (d, J = 1.6 Hz, 2H), 4.30 (q, J = 7.1 Hz, 2H), 3.51 (s, 3H), 1.35 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 164.0, 151.2, 148.7, 136.8, 132.5, 126.7, 119.6, 119.5, 61.4, 44.5, 42.3, 14.3; MS (m/z): 282 (M+); Anal. Calcd for C12H14N2O4S: C, 51.05; H, 5.00; N, 9.92. Found: C, 50.94; H, 4.96; N, 9.84.

3.3.8. 1-(Methylsulfonyl)-6-nitro-1,2-dihydroquinoline-3-carbonitrile (34)

Yield: 0.79 g (96%) as a white solid, m.p. 126–127 °C; IR: 2231, 1627, 1529, 1355, 1329 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.47 (d, J = 2.7 Hz, 1H), 8.41 (dd, J = 9.1, 2.7 Hz, 1H), 7.81 (s, 1H), 7.49 (d, J = 9.1 Hz, 1H), 5.11 (d, J = 1.4 Hz, 2H), 2.20 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 170.2, 158.5, 156.8, 138.5, 126.5, 126.4, 123.8, 118.8, 117.9, 60.7, 20.8; MS (m/z): 279 (M+); Anal. Calcd for C11H9N3O4S: C, 47.31; H, 3.25; N, 15.05. Found: C, 47.25; H, 3.23; N, 14.96.

3.3.9. 1-(1-(Methylsulfonyl)-6-nitro-1,2-dihydroquinolin-3-yl)ethan-1-one (38)

Yield: 3.26 g (93%) as a yellow solid, m.p. 195–196 °C; IR: 1677, 1524, 1344, 1169 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.45 (d, J = 2.8 Hz, 1H), 8.25 (dd, J = 9.1, 2.8 Hz, 1H), 7.95 (s, 1H), 7.74 (d, J = 9.1 Hz, 1H), 4.58 (d, J = 1.3 Hz, 2H), 3.15 s, 3H), 2.44 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 196.4, 144.5, 142.7, 135.3, 133.7, 128.0, 125.9, 125.1, 123.9, 43.5, 39.6, 25.7; MS (m/z): 296 (M+); Anal. Calcd for C12H12N2O5S: C, 48.64; H, 4.08; N, 9.45. Found: C, 48.59; H, 4.01; N, 9.37.

3.3.10. 3-Acetyl-1-(methylsulfonyl)-1,2-dihydroquinoline-6-carbonitrile (39)

Yield: 3.26 g (88%) as a light yellow solid, mp 206–207 °C; IR: 2236, 1635, 1338, 1168 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.79 (d, J = 8.4 Hz, 1H), 7.67 (dd, J = 8.4, 1.9 Hz, 1H), 7.64 (d, J = 1.9 Hz, 1H), 7.37 (s, 1H), 4.70 (d, J = 1.3 Hz, 2H), 2.82 (s, 3H), 2.48 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 195.1, 140.5, 135.9, 134.3, 132.7, 132.1, 127.7, 125.4, 117.7, 109.9, 43.4, 39.1, 25.4; MS (m/z): 276 (M+); Anal. Calcd. for C13H12N2O3S: C, 56.51; H, 4.38; N, 10.14. Found: C, 56.48; H, 4.37; N, 10.05.

3.3.11. 1-(1-(Methylsulfonyl)-6-(trifluoromethyl)-1,2-dihydroquinolin-3-yl)ethan-1-one (40)

Yield: 0.53 g (93%) as a white solid, m.p. 183–185 °C; IR: 1660, 1632, 1338, 1163 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.80 (d, J = 8.4 Hz, 1H), 7.65 (dd, J = 10.6, 8.4 Hz, 1H), 7.62 (s, 1H), 7.45 (s, 1H), 4.70 (d, J = 1.3 Hz, 2H), 2.77 (s, 3H), 2.48 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 195.3, 139.6, 135.6, 133.0, 128.6 (q, J = 33.4 Hz), 127.9 (q, J = 3.7 Hz), 127.4, 126.2 (q, J = 3.7 Hz), 125.6 123.5 (q, J = 272.2 Hz), 43.4, 38.8, 25.3; MS (m/z): 319 (M); Anal. Calcd for C13H12F3NO3S: C, 48.90; H, 3.79; N, 4.39. Found: C, 48.87; H, 3.76; N, 4.32.

3.3.12. 1-(6,7-Difluoro-1-(methylsulfonyl)-1,2-dihydroquinolin-3-yl)ethan-1-one (41)

Yield: 0.35 g (84%) as a white solid, m.p. 173–174 °C; IR: 1670, 1646, 1345, 1163 cm−1; 1H NMR (400 MHz, CDCl3): δ 7.56 (dd, J = 11.3, 7.3 Hz, 1H), 7.33 (s, 1H), 7.19 (dd, J = 9.7, 8.2 Hz, 1H), 4.65 (s, 2H), 2.70 (s, 3H), 2.46 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 195.2, 151.1 (dd, J = 255.0, 13.5 Hz), 148.8 (dd, J = 250.2, 13.4 Hz), 134.7 (d, J = 2.8 Hz), 133.2 (dd, J = 9.1, 3.3 Hz), 132.4 (t, J = 2.2 Hz), 124.0 (dd, J = 10.0, 3.6 Hz), 116.9 (dd, J = 19.0, 1.6 Hz), 115.9 (d, J = 21.4 Hz), 43.2, 38.3, 25.3; MS (m/z): 287 (M+); Anal. Calcd for C12H11F2NO3S: C, 50.17; H, 3.86; N, 4.88. Found: C, 50.02; H, 3.81; N, 4.78.

3.3.13. 1-(1-(Methylsulfonyl)-1,2-dihydro-1,8-naphthyridin-3-yl)ethan-1-one (42)

Yield: 0.37 g (87%) as a white solid, m.p. 187–189 °C; IR: 1666, 1347, 1154 cm−1; 1H NMR (400 MHz, CDCl3): δ 8.31 (dd, J = 5.0, 1.8 Hz, 1H), 7.55 (dd, J = 7.5, 1.8 Hz, 1H), 7.25 (s, 1H), 7.04 (dd, J = 7.5, 5.0 Hz, 1H), 4.79 (d, J = 1.4 Hz, 2H), 3.50 (s, 3H), 2.43 (s, 3H); 13C NMR (101 MHz, CDCl3): δ 195.0, 151.4, 149.2, 137.1, 134.4, 132.7, 119.54, 119.49, 43.8, 42.2, 25.3; MS (m/z): 252 (M+); Anal. Calcd for C11H12N2O3S: C, 52.37; H, 4.79; N, 11.10. Found: C, 52.28; H, 4.77; N, 11.01.

3.4. General Procedure for Elimination of the Methylsulfonyl Group to Give Quinolines

To a solution of the dihydroquinoline (1 mmol) in DMF (2 mL) under N2, K2CO3 (1 mmol) was added, and the reaction was stirred at 80–90 °C for 2 h. At this time, TLC (20% EtOAc/hexane) indicated the complete elimination of the methane- or p-toluenesulfonyl group. The solution was poured into water (15 mL) and extracted with EtOAc (3 × 25 mL). The organic layer was washed with 1 M HCl (2 × 25 mL), saturated NaHCO3 (2 × 25 mL) and saturated NaCl (30 mL) and then dried (Na2SO4). Removal of the solvent under vacuum gave a crude product that was further purified via column chromatography to afford the quinoline product.

3.4.1. Ethyl 6-Nitroquinoline-3-carboxylate (43)

Yield: 2.11 g (83%) as a white solid, m.p. 186–187 °C (lit. [31] m.p. 186–187 °C); IR: 1715, 1520, 1354 cm−1; 1H NMR (400 MHz, CDCl3): δ 9.62 (d, J = 2.2 Hz, 1H), 9.03 (d, J = 2.2 Hz, 1H), 8.92 (d, J = 2.5 H, 1H), 8.59 (dd, J = 9.2, 2.5 Hz, 1H), 8.32 (d, J = 9.2 Hz, 1H), 4.53 (q, J = 7.1 Hz, 2H), 1.49 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 164.4, 153.4, 151.6, 146.1, 140.1, 131.5, 125.9, 125.6, 125.1, 125.0, 62.1, 14.3; MS (m/z): 246 (M+). Elimination of the p-toluenesulfonyl group afforded this same quinoline in 81% yield.

3.4.2. Ethyl 6-Cyanoquinoline-3-carboxylate (44)

Yield: 0.29 g (86%) as a white solid, m.p. 138–139 °C; IR: 2229, 1713 cm−1; 1H NMR (400 MHz, CDCl3): δ 9.58 (d, J = 2.1 Hz, 1H), 8.90 (d, J = 2.1 Hz, 1H), 8.36 (d, J = 1.8 Hz, 1H), 8.26 (d, J = 8.7 Hz, 1H), 7.97 (dd, J = 8.7, 1.8 Hz, 1H), 4.52 (q, J = 7.1 Hz, 2H), 1.49 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 164.5, 152.9, 150.5, 138.7, 135.2, 132.2, 131.2, 126.3, 124.9, 118.0, 111.4, 62.0, 14.3; MS (m/z): 226 (M+); Anal. Calcd for C13H10N2O2: C, 69.02; H, 4.46; N, 12.38. Found: C, 68.95; H, 4.41; N, 12.27.

3.4.3. Ethyl 6-(Trifluoromethyl)quinoline-3-carboxylate (45)

Yield: 0.27 g (87%) as a white solid, m.p. 94–96 °C; IR: 1718 cm−1; 1H NMR (400 MHz, CDCl3): δ 9.56 (d, J = 2.2 Hz, 1H), 8.94 (d, J = 2.2 Hz, 1H), 8.28 (d, J = 9.0 Hz, 1H), 8.27 (s, 1H), 8.00 (dd, J = 9.0, 2.2 Hz, 1H), 4.51 (q, J = 7.1 Hz, 2H), 1.48 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 164.8, 152.2, 150.7, 139.3, 130.8, 129.4 (q, J = 33.0 Hz), 127.4 (q, J = 3.0 Hz), 127.0 (q, J = 4.4 Hz), 126.0, 124.5, 123.7 (q, J = 272.4 Hz), 61.9, 14.3; MS (m/z): 269 (M); Anal. Calcd for C13H10F3NO2: C, 58.00; H, 3.74; N, 5.20. Found: C, 57.94; H, 3.72; N, 5.07.

3.4.4. Ethyl 6-Chloroquinoline-3-carboxylate (46)

Yield: 0.27 g (96%) as a white solid, m.p. 136–138 °C; IR 1717 cm−1; 1H NMR (400 MHz, CDCl3): δ 9.56 (d, J = 1.9 Hz, 1H), 8.94 (s, 1H), 8.29 (d, J = 9.0 Hz, 1H), 8.27 (s, 1H), 8.00 (d, J = 9.0 Hz, 1H), 4.51 (q, J = 7.1 Hz, 2H), 1.48 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 165.0, 150.3, 148.2, 137.6, 133.3, 132.6, 131.1, 127.6, 127.5, 124.1, 61.7, 14.3; HRMS (m/z): 235 (M); Anal. Calcd for C12H10ClNO2: C, 61.16; H, 4.28; N, 5.94. Found: C, 61.05; H, 4.24; N, 5.87.

3.4.5. 6-Nitroquinoline-3-carbonitrile (48)

Yield: 0.72 g (95%) as a light yellow solid, m.p. 202–203 °C; IR: 2331, 1528, 1329 cm−1; 1H NMR (400 MHz, CDCl3): δ 9.22 (d, J = 2.0 Hz, 1H), 8.88 (d, J = 2.5 Hz, 1H), 8.75 (dd, J = 2.0, 1.0 Hz, 1H), 8.66 (ddd, J = 9.3, 2.5, 1.0 Hz, 1H), 8.36 (d, J = 9.3 Hz, 1H); 13C NMR (101 MHz, CDCl3): δ 152.8, 150.5, 146.7, 142.9, 132.0, 126.0, 125.2, 124.7, 116.0, 108.9; MS (m/z): 199.0382 (M+); Anal. Calcd for C10H5N3O2: C, 60.31; H, 2.53; N, 21.10. Found: C, 60.35; H, 2.59; N, 21.12.

3.5. Base Promoted Sulfone Migration in 32 to Give Ethyl 7-(Methylsulfonyl)-6-(phenethyl-amino)quinoline-3-carboxylate (49)

To a mixture of 0.10 g (0.11 mL, 0.85 mmol) of phenethylamine and 0.12 g (0.85 mmol) of K2CO3 in anhydrous DMF (2 mL) at 23 °C, 0.27 g (0.85 mmol) of 32 was added. The reaction was stirred for 1 h at 23 °C and for 12 h at 90 °C, after which TLC (20% EtOAc/hexane) indicated that the reactants were consumed. The reaction was cooled, added to water (20 mL) and extracted with ethyl acetate (3 × 15 mL). The combined organic extracts were washed with water (25 mL) and saturated NaCl (25 mL) and then dried (Na2SO4). Concentration under vacuum and purification via PTLC afforded 0.32 g (95%) of 49 as a yellow solid, m.p. 153–154 °C; IR: 3385, 1721, 1283, 1140 cm−1; 1H NMR (400 MHz, CDCl3): δ 9.22 (d, J = 2.1 Hz, 1H), 8.65 (s, 1H), 8.63 (dd, J = 2.1, 1.0 Hz, 1H), 7.37–7.23 (complex, 5H), 7.00 (s, 1H), 6.15 (br t, J = 5.0 Hz, 1H), 4.48 (q, J = 7.1 Hz, 2H), 3.55 (dt, J = 6.9, 5.0 Hz, 2H), 3.08 (t, J = 6.9 Hz, 2H), 2.90 (s, 3H), 1.46 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, CDCl3): δ 165.1, 147.5, 143.4, 142.3, 138.7, 136.4, 132.9, 131.8, 130.5, 128.80, 128.76, 126.8, 125.7, 106.9, 61.7, 44.9, 42.1, 34.9, 14.3; MS (m/z): 398 (M+); Anal. Calcd for C21H22N2O4S: C, 63.30; H, 5.57; N, 7.03. Found: C, 63.26; H, 5.55; N, 6.99. A crystal (Supplementary Materials) of 49 for X-ray structure determination was obtained from EtOH and was filed at the CCDC under deposition number 2246638.

4. Conclusions

A series of new MBH acetates were prepared and reacted with methanesulfonamide to produce tertiary dihydroquinoline sulfonamides in high yields. One example using p-toluenesulfonamide was also performed to further demonstrate the scope of the reaction. Both ester- and ketone-derived MBH acetates gave excellent yields of the cyclic sulfonamides, whereas several of the acrylonitrile-derived substrates failed to undergo the final cyclization. Further efforts to eliminate the methylsulfonyl group from these derivatives as a route to quinolines met with varying results. Although dihydroquinoline sulfonamides derived from ethyl acrylate (with one exception) and acrylonitrile (NO2-activated only) generally underwent elimination to give excellent yields of quinolines, those derived from 3-buten-2-one failed to undergo the elimination and decomposed. This problem presumably derives from the enolizable methyl ketone at C-3 of the dihydroquinolines. Further experiments on the acrylate-derived dihydroquinoline 32 bearing fluorines at C-6 and C-7 indicated that other reaction processes could occur in preference to the desired elimination. Treatment of this dihydroquinoline with phenethylamine gave 1,3-sulfonyl migrated amine 49 in 95% yield. This rearrangement is most often promoted by metals and acids, but it occurred under base conditions in the current work. Two mechanisms are proposed. Although the exact mechanism cannot be definitively assigned, the favored pathway involves (1) an attack at C-6 to displace the N-1 mesyl anion; (2) an attack of this species at C-7 to give the imine-stabilized anion; (3) the elimination of fluoride from C-7; (4) the base-promoted rearomatization of the system with the elimination of fluoride from C-6. The alternative mechanism involving a SNAr addition–elimination through both rings is disfavored due to the weak activation of the system by the ester and the presence of an electron-donating amine ortho at the site of nucleophilic attack. Further work is in progress to elucidate the details of this interesting process.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules28073256/s1, Copies of 1H-NMR and 13C-NMR spectra for all compounds and tables of crystal data for compound 49 (CCDC 2246638) [32,33,34,35,36].

Author Contributions

Project conception, project administration, formal analysis and writing the manuscript text, R.A.B.; investigation, methodology, analysis and writing the experimental section, E.A. and K.F.; reviewing and editing, R.A.B., E.A. and K.F. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support for this work was obtained from the Oklahoma State University Foundation and the College of Arts and Sciences at Oklahoma State University. The authors are indebted to the OSU College of Arts and Sciences for funds to purchase several departmental instruments, including an FT-IR and a 400 MHz NMR unit for the statewide NMR facility. The NMR facility was initially established with support from the NSF (BIR-9512269), the Oklahoma State Regents for Higher Education, the W.M. Keck Foundation and Conoco, Inc. Finally, the authors thank the NSF (CHE-1726630) and the University of Oklahoma for funds to purchase the X-ray equipment and computers used to obtain the crystal structure in this study.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

E.A. and K.F. wish to thank the OSU Foundation for K. Darrell Berlin Fellowships in Summer 2020 and 2021, respectively. The authors also thank Douglas R. Powell at the University of Oklahoma Chemical Crystallography Laboratory for determining the structure of compound 49.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Annor-Gyamfi, J.K.; Ametsetor, E.; Meraz, K.; Bunce, R.A. Dihydroquinolines, dihydronaphthyridines and quinolones by domino reactions of Morita-Baylis-Hillman acetates. Molecules 2021, 26, 890. [Google Scholar] [CrossRef] [PubMed]
  2. Annor-Gyamfi, J.K.; Ametsetor, E.; Meraz, K.; Bunce, R.A. Naphthalenes and Quinolines by Domino Reactions of Morita–Baylis–Hillman Acetates. Molecules 2020, 25, 5168. [Google Scholar] [CrossRef]
  3. Imanzadeh, G.H.; Zare, A.; Kalafi-Hezhad, A.; Hasaninejad, A.; Moosavi Zare, A.R.; Parhami, A. Microwave-assisted Michael addition of sulfonamides to α,β-unsaturated esters: A rapid entry to protected β-amino acid synthesis. J. Iran. Chem. Soc. 2007, 4, 467–475. [Google Scholar] [CrossRef]
  4. Zare, A.; Hasaninejad, A.; Khalafi-Nezhad, A.; Moosavi Zare, A.R.; Parhami, A. An efficient solventless method for the synthesis of N,N-dialkyl sulfonamide derivatives. J. Iran. Chem. Soc. 2008, 5, 617–622. [Google Scholar] [CrossRef]
  5. Wang, D.; Wei, Y.; Shi, M. Synthesis of highly functionalized aminoindolizines by titanium(IV) chloride mediated cycloisomerization and phosphine-catalyzed aza-Michael addition reactions. Asian J. Org. Chem. 2013, 2, 480–485. [Google Scholar] [CrossRef]
  6. Huang, F.-P.; Dong, Z.-W.; Zhang, H.-R.; Hui, X.-P. Tandem aza-Michael-aldol reactions: One-pot synthesis of functionalized piperidine derivatives. J. Heterocycl. Chem. 2013, 51, 532–538. [Google Scholar] [CrossRef]
  7. Douraki, S.M.; Massah, A.R. The zeolite ZSM-5-SO3H catalyzed aza-Michael addition of amines and sulfonamides to electron-deficient alkenes under solvent-free conditions. Indian J. Chem. Sect. B Org. Chem. Med. Chem. 2015, 54, 1346–1349. [Google Scholar] [CrossRef]
  8. Luo, H.; Yan, X.; Chen, L.; Li, Y.; Liu, N.; Yin, G. Enantioselective catalytic domino aza-Michael-Henry reactions: One-pot asymmetric synthesis of 3-nitro-1,2-dihydroquinolines via iminium activation. Eur. J. Org. Chem. 2016, 2016, 1702–1707. [Google Scholar] [CrossRef]
  9. Zhai, X.-D.; Yang, Z.-D.; Luo, Z.; Xu, H.-T. Asymmetric catalyzed intramolecular aza-Michael reaction mediated by quinine-derived primary amines. Chin. Chem. Lett. 2017, 28, 1793–1797. [Google Scholar] [CrossRef]
  10. Roy, T.K.; Parhi, B.; Ghorai, P. Chinchonamine squaramide catalyzed asymmetric aza-Michael reaction: Dihydroisoquinolines and tetrahydropyridines. Angew. Chem. Int. Ed. 2018, 57, 9397–9401. [Google Scholar] [CrossRef] [PubMed]
  11. Bunce, R.A.; Smith, C.L.; Knight, C.L. N-(Nitrophenyl) benzamide and benzenesulfonamide derivatives by nucleophilic aromatic substitution. Org. Prep. Proced. Int. 2004, 36, 482–486. [Google Scholar] [CrossRef]
  12. Bordwell, F.G.; Algrim, D. Nitrogen acids. 1. Carboxamides and sulfonamides. J. Org. Chem. 1976, 41, 2507–2508. [Google Scholar] [CrossRef]
  13. Bordwell, F.G.; Fried, H.E.; Hughes, D.L.; Lynch, T.Y.; Satish, A.V.; Whang, Y.E. Acidities of carboxamides, hydroxamic acids, carbohydrazides, benzenesulfonamides, and benzenesulfonohydrazides in DMSO solution. J. Org. Chem. 1990, 55, 3330–3336. [Google Scholar] [CrossRef]
  14. Trost, B.M. The atom economy–A search for synthetic efficiency. Science 1991, 254, 1471–1477. [Google Scholar] [CrossRef] [PubMed]
  15. Trost, B.M. Atom economy—A challenge for organic synthesis: Homogeneous catalysis leads the way. Angew. Chem. Int. Ed. 1995, 90, 259–281. [Google Scholar] [CrossRef]
  16. Trost, B.M. Atom economy: A challenge for enhanced synthetic efficiency. In Handbook of Green Chemistry; Wiley Online Library: Hoboken, NJ, USA, 2012; pp. 1–33. [Google Scholar] [CrossRef]
  17. Kim, J.N.; Lee, H.J.; Lee, K.Y.; Kim, H.S. Synthesis of 3-quinolinecarboxylic acid esters from the Baylis–Hillman adducts of 2-halobenzaldehyde N-tosylimines. Tetrahedron Lett. 2001, 42, 3737–3740. [Google Scholar] [CrossRef]
  18. Niu, Q.; Mao, H.; Yuan, G.; Gao, J.; Liu, H.; Tu, Y.; Wang, X.; Lv, X. Copper-catalyzed domino SN2/Coupling reaction: A versatile and facile synthesis of cyclic compounds from Baylis-Hillman acetates. Adv. Synth. Catal. 2013, 355, 1185–1192. [Google Scholar] [CrossRef]
  19. Chen, L.; Wilder, P.T.; Drennen, B.; Tran, J.; Roth, B.M.; Chesko, K.; Shapiro, P.; Fletcher, S. Structure-based design of 3-carboxy-substituted 1,2,3,4-tetrahydroquinolines as inhibitors of myeloid cell leukemia-1 (Mcl-1). Org. Biomol. Chem. 2015, 14, 5505–5510. [Google Scholar] [CrossRef]
  20. Okolotowicz, K.J.; Dwyer, M.; Ryan, D.; Cheng, J.; Cashman, E.A.; Moore, S.; Mercola, M.; Cashman, J.R. Novel tertiary sulfonamides as potent anti-cancer agents. Bioorganic Med. Chem. 2018, 26, 4441–4451. [Google Scholar] [CrossRef]
  21. Peerzada, M.N.; Khan, P.; Ahmad, K.; Hassan, M.I.; Azam, A. Synthesis, characterization and biological evaluation of tertiary sulfonamide derivatives of pyridyl-indole based heteroaryl chalcone as potential carbonic anhydrase IX inhibitors and anticancer agents. Eur. J. Med. Chem. 2018, 155, 13–23. [Google Scholar] [CrossRef]
  22. Durgapal, S.D.; Soman, S.S. Evaluation of novel coumarin-proline sulfonamide hybrids as anticancer and antidiabetic agents. Synth. Commun. 2019, 49, 2869–2883. [Google Scholar] [CrossRef]
  23. Taslimi, P.; Sujayev, A.; Mamedova, S.; Kalın, P.; Gulçin, P.; Sadeghian, N.; Beydemir, S.; Kufrevioglu, O.I.; Alwasel, S.H.; Farzaliyev, V.; et al. Synthesis and bioactivity of several new hetaryl sulfonamides. J. Enzym. Inhib. Med. Chem. 2017, 32, 137–145. [Google Scholar] [CrossRef] [PubMed]
  24. Pero, J.E.; Matthews, J.M.; Behm, D.J.; Brnardic, E.J.; Brooks, C.; Budzik, B.W.; Costell, M.H.; Donatelli, C.A.; Eisennagel, S.H.; Erhard, K.; et al. Design and optimization of sulfone pyrrolidine sulfonamide antagonists of transient receptor potential vanilloid-4 with in vivo activity in a pulmonary edema model. J. Med. Chem. 2018, 61, 11209–11220. [Google Scholar] [CrossRef] [PubMed]
  25. Mondal, S. Sulfonamide synthesis under green conditions. Synth. Commun. 2021, 51, 1023–1044. [Google Scholar] [CrossRef]
  26. Procopiou, P.A.; Baugh, S.P.D.; Flack, S.S.; Inglis, G.G.A. An extremely powerful acylation reaction of alcohols with acid anhydrides catalyzed by trimethylsilyl trifluoromethanesulfonate. J. Org. Chem. 1998, 63, 2342–2347. [Google Scholar] [CrossRef]
  27. Gnanasekaran, K.K.; Yoon, J.; Bunce, R.A. Nucleophilic additions to polarized vinylarenes. Tetrahedron Lett. 2016, 57, 3190–3193. [Google Scholar] [CrossRef] [Green Version]
  28. Akhtar, R.; Yousaf, M.; Naqvi, S.A.R.; Irfan, M.; Zahoor, A.F.; Hussain, A.I.; Chatha, S.A.S. Synthesis of ciprofloxacin-based compounds: A review. Synth. Commun. 2016, 46, 1849–1879. [Google Scholar] [CrossRef]
  29. Flynn, A.J.; Ford, A.; Maguire, A.R. Synthetic and mechanistic aspects of sulfonyl migrations. Org. Biomol. Chem. 2020, 18, 2549–2610. [Google Scholar] [CrossRef]
  30. Xin, X.; Wang, D.; Li, X.; Wan, B. Highly regioselective migration of the sulfonyl group: Easy access to functionalized pyrroles. Angew. Chem. Int. Ed. 2012, 51, 1693–1697. [Google Scholar] [CrossRef]
  31. Tummatorn, J.; Thongsornkleeb, C.; Ruchirawat, S.; Gettongsong, T. Synthesis of 2,4-unsubstituted quinoline-3-carboxylic acid ethyl esters from arylmethyl azides via a domino process. Org. Biomol. Chem. 2013, 11, 1463–1467. [Google Scholar] [CrossRef]
  32. Bruker. APEX3, version 2018.7-2; Bruker Inc.: Madison, WI, USA, 2018.
  33. Bruker. SAINT, version 8.38A; Bruker Inc.: Madison, WI, USA, 2016.
  34. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Cryst. 2015, 48, 3–10. [Google Scholar] [CrossRef] [Green Version]
  35. Sheldrick, G.M. SHELXT-Integrated Space-Group and Crystal-Structure Determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  36. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Planned synthesis of tertiary cyclic sulfonamides and quinolines.
Scheme 1. Planned synthesis of tertiary cyclic sulfonamides and quinolines.
Molecules 28 03256 sch001
Scheme 2. Wang group approach to tertiary cyclic sulfonamides and quinolines.
Scheme 2. Wang group approach to tertiary cyclic sulfonamides and quinolines.
Molecules 28 03256 sch002
Figure 1. Experimental drugs incorporating tertiary cyclic sulfonamides (highlighted in blue).
Figure 1. Experimental drugs incorporating tertiary cyclic sulfonamides (highlighted in blue).
Molecules 28 03256 g001
Figure 2. X-ray structure for product 49 (CCDC 2246638) from reaction of 32 with phenethylamine.
Figure 2. X-ray structure for product 49 (CCDC 2246638) from reaction of 32 with phenethylamine.
Molecules 28 03256 g002
Scheme 3. SNAr pathway (mechanism 1) for sulfonyl migration in 32 to give 49.
Scheme 3. SNAr pathway (mechanism 1) for sulfonyl migration in 32 to give 49.
Molecules 28 03256 sch003
Scheme 4. Sulfonyl migration (mechanism 2) to give 49 from 32 that avoids SNAr addition–elimination.
Scheme 4. Sulfonyl migration (mechanism 2) to give 49 from 32 that avoids SNAr addition–elimination.
Molecules 28 03256 sch004
Table 1. Synthesis of MBH acetates.
Table 1. Synthesis of MBH acetates.
Molecules 28 03256 i001
XYZPdtYield (%)
CO2Et2-F-5-NO2CH1398 a
CO2Et2-F-5-CNCH1498 a
CO2Et2-F-5-CF3CH1592
CO2Et2-F-5-ClCH1693
CO2Et2,4,5-tri-FCH1793
CO2Et2-FN1896 a
CN2-F-5-NO2CH1998 a
CN2-F-5-CNCH2098 a
CN2-F-5-CF3CH2193
CN2-F-5-ClCH2295
COCH32-F-5-NO2CH2396
COCH32-F-5-CNCH2498
COCH32-F-5-CF3CH2595
COCH32,4,5-tri-FCH2697
COCH32-FN2789
a Previously reported [2].
Table 2. Formation of tertiary dihydroquinoline sulfonamides.
Table 2. Formation of tertiary dihydroquinoline sulfonamides.
Molecules 28 03256 i002
SubstratePdtXYZYield (%)
13 (X = CO2Et; Y = 2-F-5-NO2; Z = CH)28-MsCO2Et6-NO2CH96
13 (X = CO2Et; Y = 2-F-5-NO2; Z = CH)28-TsCO2Et6-NO2CH91
14 (X = CO2Et; Y = 2-F-5-CN; Z = CH)29CO2Et6-CNCH93
15 (X = CO2Et; Y = 2-F-5-CF3; Z = CH)30CO2Et6-CF3CH96
16 (X = CO2Et; Y = 2-F-5-Cl; Z = CH)31CO2Et6-ClCH91
17 (X = CO2Et; Y = 2,4,5-tri-F; Z = CH)32CO2Et6,7-di-FCH93
18 (X = CO2Et; Y = 2-F; Z = N)33CO2EtHN92
19 (X = CN; Y = 2-F-5-NO2; Z = CH)34CN6-NO2CH96
20 (X = CN; Y = 2-F-5-CN; Z = CH)35CN6-CNCH0 a
21 (X = CN; Y = 2-F-5-CF3; Z = CH)36CN6-CF3CH0 a
22 (X = CN; Y = 2-F-5-Cl; Z = CH)37CN6-ClCH0 a
23 (X = COMe; Y = 2-F-5-NO2; Z = CH)38COMe6-NO2CH93
24 (X = COMe; Y = 2-F-5-CN; Z = CH)39COMeCNCH88
25 (X = COMe; Y = 2-F-5-CF3; Z = CH)40COMeCF3CH93
26 (X = COMe; Y = 2,4,5-tri-F; Z = CH)41COMe6,7-di-FCH84
27 (X = COMe; Y = 2-F; Z = CH)42COMeHN87
a Following initial addition to the side chain, ring closure failed to occur.
Table 3. Elimination results.
Table 3. Elimination results.
Molecules 28 03256 i003
SubstrateXYZPdtYield (%)
28-MsCO2Et6-NO2CH4383
28-TsCO2Et6-NO2CH4385
29CO2Et6-CNCH4486
30CO2Et6-CF3CH4587
31CO2Et6-ClCH4696
32CO2Et6,7-di-FCH470 a
34CN6-NO2CH4895
a See text below regarding the formation of compound 49.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ametsetor, E.; Fobi, K.; Bunce, R.A. Synthesis and Elimination Pathways of 1-Methanesulfonyl-1,2-dihydroquinoline Sulfonamides. Molecules 2023, 28, 3256. https://doi.org/10.3390/molecules28073256

AMA Style

Ametsetor E, Fobi K, Bunce RA. Synthesis and Elimination Pathways of 1-Methanesulfonyl-1,2-dihydroquinoline Sulfonamides. Molecules. 2023; 28(7):3256. https://doi.org/10.3390/molecules28073256

Chicago/Turabian Style

Ametsetor, Ebenezer, Kwabena Fobi, and Richard A. Bunce. 2023. "Synthesis and Elimination Pathways of 1-Methanesulfonyl-1,2-dihydroquinoline Sulfonamides" Molecules 28, no. 7: 3256. https://doi.org/10.3390/molecules28073256

Article Metrics

Back to TopTop