Next Article in Journal
Hibiscus acetosella: An Unconventional Alternative Edible Flower Rich in Bioactive Compounds
Next Article in Special Issue
Three-Step Synthesis of the Antiepileptic Drug Candidate Pynegabine
Previous Article in Journal
Characterization of Polyphenols from Chenopodium botrys after Fractionation with Different Solvents and Study of Their In Vitro Biological Activity
Previous Article in Special Issue
Synthetic Studies on Tetracyclic Diquinane Lycopodium Alkaloids Magellanine, Magellaninone and Paniculatine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Rapid Construction of a Chloromethyl-Substituted Duocarmycin-like Prodrug

by
Christoffer Bengtsson
* and
Ylva Gravenfors
Drug Discovery & Development Platform, Science for Life Laboratory, Department of Organic Chemistry, Stockholm University, Tomtebodavägen 23a, 17165 Solna, Sweden
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(12), 4818; https://doi.org/10.3390/molecules28124818
Submission received: 25 May 2023 / Revised: 15 June 2023 / Accepted: 15 June 2023 / Published: 16 June 2023
(This article belongs to the Special Issue Design and Synthesis of Bioactive Organic Molecules)

Abstract

:
The construction of duocarmycin-like compounds is often associated with lengthy synthetic routes. Presented herein is the development of a short and convenient synthesis of a type of duocarmycin prodrug. The 1,2,3,6-tetrahydropyrrolo[3,2-e]indole-containing core is here constructed from commercially available Boc-5-bromoindole in four steps and 23% overall yield, utilizing a Buchwald–Hartwig amination followed by a sodium hydride-induced regioselective bromination. In addition, protocols for selective mono- and di-halogenations of positions 3 and 4 were also developed, which could be useful for further exploration of this scaffold.

1. Introduction

Duocarmycin A (1) and SA (2) are prominent members of the duocarmycin family that possess extreme cytotoxic properties (Figure 1) [1,2,3]. They were isolated from the Streptomyces sp. in Japan in 1988 and 1990, respectively [4,5]; in the early 1990s, their structures were confirmed by synthesis [6,7,8]. Since then, duocarmycin and its analogs have attracted a lot of attention among synthetic and medicinal chemists, owing to their structural complexity and interesting biological properties. Their mode of action is site-specific DNA alkylation, and their strongly alkylating properties can be attributed to the strained cyclopropane moiety (Figure 1). Unfortunately, the cytotoxicity is not only devoted to the cancer cells; therefore, a variety of duocarmycin analogs [1,2,3], prodrugs [9,10,11,12,13,14,15,16,17,18,19], and even antibody–drug conjugates [20] have been developed in the pursuit for more selective cancer treatments. In a medicinal chemistry project working with prodrugs that, upon site-selective CYP2W1 oxidation, form the phenolic counterpart and render the compound harmful [14,17] (3, Figure 1), we needed access to the chloromethyl-substituted 1,2,3,6-tetrahydropyrrolo[3,2-e]indole core 10 (Figure 2).
The existing synthetic pathways are elaborative and/or give the wrong substitution pattern (Figure 2). Furthermore, in our early attempts to use Boc-5-nitroindole 9 as starting material, we faced several problems, such as over-reduction when reducing the nitro group (i.e., the generation of indoline), the generation of complex mixtures when performing the halogenation reaction on the aniline, and problems with controlling the mono-Boc protection of the aniline.
Figure 2. Previous versus new routes from commercial starting materials [14,16,21,22].
Figure 2. Previous versus new routes from commercial starting materials [14,16,21,22].
Molecules 28 04818 g002
In our approach, we envisioned that the desired di-Boc-protected 5-aminoindole intermediate 12 (Figure 3) could be synthesized from commercially available Boc-5-bromoindole 11 via a Buchwald–Hartwig amination with tBu-carbamate followed by a regioselective bromination. This strategy would considerably shorten the route and also overcome the problems related to the nitro reduction and mono-Boc protection of the aniline nitrogen; vide supra.

2. Results and Discussion

The Pd(OAc)2/XPhos-catalyzed Buchwald–Hartwig amination of Boc-5-bromoindole (11) with tBu-carbamate performed well, and compound 13 could be isolated in 78% yield (Scheme 1). Performing the subsequent halogenation under acidic conditions (i.e., NXS/TsOH) on the Boc-protected aniline gave the wrong regioisomer, although with complete selectivity, and the 3-bromo (14) and 3-iodo (15) products could be isolated in 74% and 71% yields, respectively, using the two different halogen sources. We envisioned that the deprotonation of the Boc-protected aniline with NaH prior to the halogenation might render the aromatic ring sufficiently electron-rich to direct the halogenation to the right position (see Supporting Information). Gratifyingly, that strategy gave the desired 4-bromo analog 12 in 65% yield with complete regioselectivity. All attempts to introduce iodine in this position failed, even when using a more electrophilic I+ source (i.e., N-Iodosaccharin [23]), other solvents, or elevated temperatures.
To our delight, further halogenation of 12 to give 3-iodo-4-bromo compound 16 went smoothly under acidic conditions (NIS/TsOH) in 71% yield. To conclude the synthesis towards the duocarmycin-type prodrug, compound 12 smoothly underwent allylation with 1,3-dichloropropene to give 17 [14] in 82% yield, followed by a tris(trimethylsilyl)silane (TTMSS)/azaisobutyronitrile (AIBN)-induced radical 5-exo-trig cyclization according to published procedures to furnish compound 10 [14] in 56% yield (Scheme 2). After Boc deprotection and subsequent EDC/NaHCO3 amide coupling with 5-fluoroindole-2-carboxylic acid, the desired prodrug rac18 [17] was isolated in 65% yield over two steps. In addition, the enantiomers were separated by chiral supercritical fluid chromatography (SFC) to give (+)—18 and (−)—18 with ee ≥ 99%.

3. Materials and Methods

General Methods: All solvents and reagents were used as received from commercial suppliers. N-Bromosuccinimide (NBS) was recrystallized from hot water and dried under vacuum for 24 h and then stored under cold and dark conditions. Sodium hydride was used as 60% dispersion in mineral oil. Column chromatography was employed on normal-phase silica gel (230–400 mesh, 60 Å; the eluents are given in brackets). 1H- and 13C-NMR spectra were recorded on a 400 MHz spectrometer at 298 K and calibrated using the residual peak of the solvent as an internal standard [CDCl3 (CHCl3 δH 7.26 ppm, CDCl3 δC 77.16 ppm)]. HRMS was performed using a microTOF instrument with electrospray ionization (ESI), and sodium formate was used as a calibration chemical. Optical rotations were measured on a polarimeter at 589 nm (D line of sodium) and 20 °C. Chiral chromatography was performed on supercritical fluid chromatography equipment, using mixtures of MeOH and supercritical CO2 as eluents.
Di-tert-butyl 1-(chloromethyl)-1,2-dihydropyrrolo[3,2-e]indole-3,6-dicarboxylate (10): tert-Butyl-4-bromo-5-((tert-butoxycarbonyl)(3-chloroallyl)amino)-1H-indole-1-carboxylate 17 (600 mg, 1.24 mmol) was dissolved in dry toluene (40 mL), and the solution was degassed for 1 h (by bubbling N2 gas through the solution under stirring). Azobisisobutyronitrile (AIBN) (49 mg, 0.30 mmol) and tris(trimethylsilyl)silane (TTMSS) (0.41 mL, 1.34 mmol) were added, and the reaction was heated to 90 °C (with a preheated oil bath) in a sealed tube for 5 h. The solvent was evaporated, and the crude material was dissolved in MeOH (12 mL) and stirred at rt for 10 min. The solvent was evaporated, and the crude product was purified by column chromatography on silica gel (hexanes:EtOAc 95:5) to give compound 10 as a colorless oil (280 mg, 56%). The spectral data agreed with the published data [14].
tert-Butyl 4-bromo-5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate (12): tert-Butyl 5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate 13 (200 mg, 0.60 mmol) was dissolved in dry DMF (2 mL) and cooled to 0 °C with an ice bath. NaH (60 mg, 60% in mineral oil, 1.5 mmol) was added, followed by NBS (129 mg, 0.72 mmol); the ice bath was removed, and the reaction was stirred for 30 min. The reaction mixture was poured onto saturated NaHCO3 (aq) and extracted with EtOAc. The organic phase was dried (Na2SO4), filtered, and concentrated. The crude material was purified by column chromatography on silica gel (hexanes:EtOAc 95:5) to give compound 12 as a colorless foam (160 mg, 65%). The spectral data agreed with the published data [14].
tert-Butyl 5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate (13): N-Boc-5-bromoindole 11 (1.5 g, 5.06 mmol), tert-butyl carbamate (712 mg, 6.08 mmol), Pd(OAc)2 (57 mg, 0.25 mmol), XPhos (241 mg, 0.50 mmol), and Cs2CO3 (2.31 g, 7.09 mmol) were mixed in dry 1,4-dioxane (45 mL), and the vessel was flushed with N2 gas, sealed, and heated to 90 °C for 20 h. The reaction mixture was diluted with EtOAc, filtered through Celite, and concentrated. The crude material was purified by column chromatography on silica gel (hexanes:EtOAc 95:5) to give compound 13 as a colorless foam (1.32 g, 78%). 1H-NMR (CDCl3, 400 MHz) δ 8.01 (brd, J = 8.0 Hz, 1H), 7.75 (brs, 1H), 7.55 (brd, J = 4.0 Hz, 1H), 7.14 (dd, J = 8.0, 4.0 Hz, 1H), 6.70 (brs, 1H, NH), 6.48 (dd, J = 3.7, 0.8 Hz, 1H), 1.65 (s, 9H), 1.52 (s, 9H); 13C-NMR (CDCl3, 100 MHz) δ 153.3, 149.8, 133.7, 131.5, 131.1, 126.6, 116.4, 115.3, 110.9, 107.4, 83.6, 80.3, 28.5 (3C), 28.3 (3C); HRMS (ESI/TOF) m/z: [M + Na]+ Calcd for C18H24N2O4Na 355.1634; Found 355.1633.
tert-Butyl 3-bromo-5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate (14): tert-Butyl 5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate 13 (200 mg, 0.60 mmol) was dissolved in DMF (2 mL), NBS (118 mg, 0.66 mmol) and TsOH·H2O (23 mg, 0.12 mmol) were added, and the reaction was stirred at rt for 10 min. The reaction mixture was poured onto saturated NaHCO3 (aq) and extracted with EtOAc. The organic phase was dried (Na2SO4), filtered, and concentrated. The crude material was purified by column chromatography on silica gel (hexanes:EtOAc 95:5) to give compound 14 as a colorless foam (183 mg, 74%). 1H-NMR (CDCl3, 400 MHz) δ 8.02 (brd, J = 8.0 Hz, 1H), 7.65 (brs, 1H), 7.60 (brs, 1H), 7.24 (brd, J = 8.0 Hz, 1H), 6.67 (brs, 1H, NH), 1.65 (s, 9H), 1.54 (s, 9H); 13C-NMR (CDCl3, 100 MHz) δ 153.1, 148.9, 134.5, 130.9, 130.0, 125.5, 117.5, 115.6, 109.2, 97.9, 84.4, 80.6, 28.5 (3C), 28.3 (3C); HRMS (ESI/TOF) m/z: [M + Na]+ Calcd for C18H23BrN2O4Na 433.0739; Found 433.0755.
tert-Butyl 5-((tert-butoxycarbonyl)amino)-3-iodo-1H-indole-1-carboxylate (15): tert-Butyl 5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate 13 (1.3 g, 3.91 mmol) was dissolved in DMF (14 mL), NIS (1.06 g, 4.71 mmol) and TsOH·H2O (149 mg, 0.78 mmol) were added, and the reaction was stirred at rt for 15 h. The reaction mixture was poured onto saturated NaHCO3 (aq) and extracted with EtOAc. The organic phase was washed with 10 wt% Na2S2O5 (aq), dried (Na2SO4), filtered, and concentrated. The crude material was purified by column chromatography on silica gel (hexanes:EtOAc 95:5) to give compound 15 as a colorless foam (1.28 g, 71%). 1H-NMR (CDCl3, 400 MHz) δ 8.00 (brd, J = 8.0 Hz, 1H), 7.69 (brs, 1H), 7.53–7.46 (m, 1H), 7.27 (brd, J = 8.0 Hz, 1H), 6.73 (brs, 1H, NH), 1.65 (s, 9H), 1.54 (s, 9H); 13C-NMR (CDCl3, 100 MHz) δ 153.1, 148.7, 134.6, 132.7, 131.1, 130.8, 117.5, 115.5, 111.3, 84.3, 80.6, 65.4, 28.5 (3C), 28.2 (3C); HRMS (ESI/TOF) m/z: [M + Na]+ Calcd for C18H23IN2O4Na 481.0601; Found 481.0595.
tert-Butyl 4-bromo-5-((tert-butoxycarbonyl)amino)-3-iodo-1H-indole-1-carboxylate (16): tert-Butyl 4-bromo-5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate 12 (140 mg, 0.34 mmol) was dissolved in DMF (1.4 mL), NIS (114 mg, 0.51 mmol) and TsOH·H2O (16 mg, 0.08 mmol) were added, and the reaction was stirred at rt for 16 h. The reaction mixture was poured onto saturated NaHCO3 (aq) and extracted with EtOAc. The organic phase was washed with 10 wt% Na2S2O5 (aq), dried (Na2SO4), filtered, and concentrated. The crude material was purified by column chromatography on silica gel (hexanes:EtOAc 95:5) to give compound 16 as a colorless foam (130 mg, 71%). 1H-NMR (CDCl3, 400 MHz) δ 8.11 (m, 2H), 7.77 (s, 1H), 7.08 (brs, 1H, NH), 1.65 (s, 9H), 1.54 (s, 9H); 13C-NMR (CDCl3, 100 MHz) δ 153.0, 148.2, 134.0, 132.7, 131.6, 126.5, 118.5, 114.5, 105.3, 85.0, 81.1, 61.2, 28.5 (3C), 28.2 (3C); HRMS (ESI/TOF) m/z: [M + Na]+ Calcd for C18H22BrIN2O4Na 558.9706; Found 558.9700.
tert-Butyl-4-bromo-5-((tert-butoxycarbonyl)(3-chloroallyl)amino)-1H-indole-1-carboxylate (17): tert-Butyl 4-bromo-5-((tert-butoxycarbonyl)amino)-1H-indole-1-carboxylate 12 (650 mg, 1.58 mmol) was dissolved in dry DMF (12 mL) and cooled to 0 °C, NaH (190 mg, 60% in mineral oil, 4.74 mmol) was added, and the reaction was stirred at 0 °C for 5 min. 1,3-Dichloropropene was added, the ice bath was removed, and the reaction was stirred at rt for 1 h. The reaction mixture was poured onto saturated NaHCO3 (aq) and extracted with EtOAc. The organic phase was dried (Na2SO4), filtered, and concentrated. The crude material was purified by column chromatography on silica gel (hexanes:EtOAc 95:5) to give compound 17 as a colorless oil (630 mg, 82%). The spectral data agreed with the published data [14].
(1-(chloromethyl)-1,6-dihydropyrrolo[3,2-e]indol-3(2H)-yl)(5-fluoro-1H-indol-2-yl)methanone (18): Di-tert-butyl 1-(chloromethyl)-1,2-dihydropyrrolo[3,2-e]indole-3,6-dicarboxylate 10 (280 mg, 0.69 mmol) was dissolved in 4 M HCl in 1,4-dioxane (15 mL, 60 mmol), and the reaction was stirred at rt for 22 h. The solvent was evaporated, and the crude material was coevaporated from EtOAc two times. The crude material, together with 5-fluoro-1H-indole-2-carboxylic acid 19 (148 mg, 0.83 mmol), N-ethyl-N′-(3-dimethylaminopropyl)carbodiimide hydrochloride (EDC) (396 mg, 2.07 mmol), and NaHCO3 (289 mg, 3.45 mmol), were mixed in dry DMF (10 mL), and the reaction was stirred at rt for 5 h. The reaction mixture was poured onto saturated NaHCO3 (aq) and extracted with EtOAc. The organic phase was dried (Na2SO4), filtered, and concentrated. The crude material was purified by column chromatography on silica gel (hexanes:EtOAc 60:40 to 50:50) to give compound 18 (253 mg, 65%) as an off-white solid. The spectral data agreed with the published results [17]. The racemic product was separated by chiral supercritical fluid chromatography (SFC) to give (+)—18, [α]D (c = 1.0, acetone) +17 and (−)—18, [α]D (c = 1.0, acetone) -17, both with ee ≥ 99% (for chromatographic conditions and chromatograms, see Supporting Information).

4. Conclusions

In conclusion, we developed a four-step route to the desired chloromethyl-substituted 1,2,3,6-tetrahydropyrrolo[3,2-e]indole core 10, utilizing an unconventional NaH promoted site-selective bromination of Boc-protected amino indole 13 as the key step. Additionally, 3-iodo-4-bromo indole 16 constitutes an interesting starting point for further diversification. Closely related 3-iodo-4-bromo-indoles have been used in Pd-catalyzed cross-couplings such as the Mizoroki-Heck [24,25,26], Negishi [27], and Suzuki-Miyaura [28,29] reactions in various natural products and heterocyclic syntheses. Finally, the racemate of compound 18 was separated with chiral supercritical fluid chromatography for further investigation of this interesting prodrug.

Supplementary Materials

Supporting information with 1H-NMR and 13C-NMR of all new compounds can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28124818/s1.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Magnus Ingelman-Sundberg at the Department of Physiology and Pharmacology at the Karolinska Institute, Stockholm, Sweden, for inspiring discussions around this interesting project.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Boger, D.L.; Johnson, D.S. CC-1065 and the Duocarmycins: Understanding their Biological Function through Mechanistic Studies. Angew. Chem. Int. Ed. Engl. 1996, 35, 1438–1474. [Google Scholar] [CrossRef]
  2. Boger, D.L.; Boyce, C.W.; Garbaccio, R.M.; Goldberg, J.A. CC-1065 and the Duocarmycins: Synthetic Studies. Chem. Rev. 1997, 97, 787–828. [Google Scholar] [CrossRef] [PubMed]
  3. Tercel, M.; Gieseg, M.A.; Denny, W.A.; Wilson, W.R. Synthesis and Cytotoxicity of Amino-seco-DSA: An Amino Analogue of the DNA Alkylating Agent Duocarmycin SA. J. Org. Chem. 1999, 64, 5946–5953. [Google Scholar] [CrossRef]
  4. Takahashi, I.; Takahashi, K.-I.; Ichimura, M.; Morimoto, M.; Asano, K.; Kawamoto, I.; Tomita, F.; Nakano, H. Duocarmycin A, a new antitumor antibiotic from Streptomyces. J. Antibiot. 1988, 41, 1915–1917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Ichimura, M.; Ogawa, T.; Takahashi, K.-I.; Kobayashi, E.; Kawamoto, I.; Yasuzawa, T.; Takahashi, I.; Nakano, H. Duocarmycin SA, a new antitumor antibiotic from Streptomyces sp. J. Antibiot. 1990, 43, 1037–1038. [Google Scholar] [CrossRef] [PubMed]
  6. Fukuda, Y.; Nakatani, K.; Ito, Y.; Terashima, S. First total synthesis of dl-duocarmycin A. Tetrahedron Lett. 1990, 31, 6699–6702. [Google Scholar] [CrossRef]
  7. Boger, D.L.; McKie, J.A.; Nishi, T.; Ogiku, T. Enantioselective Total Synthesis of (+)-Duocarmycin A, epi-(+)-Duocarmycin A, and Their Unnatural Enantiomers. J. Am. Chem. Soc. 1996, 118, 2301–2302. [Google Scholar] [CrossRef]
  8. Boger, D.L.; Machiya, K. Total synthesis of (+)-duocarmycin SA. J. Am. Chem. Soc. 1992, 114, 10056–10058. [Google Scholar] [CrossRef]
  9. Tietze, L.F.; Schuster, H.J.; Schmuck, K.; Schuberth, I.; Alves, F. Duocarmycin-based prodrugs for cancer prodrug monotherapy. Bioorg. Med. Chem. 2008, 16, 6312–6318. [Google Scholar] [CrossRef]
  10. Li, L.-S.; Sinha, S.C. Studies toward the duocarmycin prodrugs for the antibody prodrug therapy approach. Tetrahedron Lett. 2009, 50, 2932–2935. [Google Scholar] [CrossRef] [Green Version]
  11. Schuster, H.J.; Krewer, B.; Von Hof, J.M.; Schmuck, K.; Schuberth, I.; Alves, F.; Tietze, L.F. Synthesis of the first spacer containing prodrug of a duocarmycin analogue and determination of its biological activity. Org. Biomol. Chem. 2010, 8, 1833–1842. [Google Scholar] [CrossRef]
  12. Lajiness, J.P.; Robertson, W.M.; Dunwiddie, I.; Broward, M.A.; Vielhauer, G.A.; Weir, S.J.; Boger, D.L. Design, Synthesis, and Evaluation of Duocarmycin O-Amino Phenol Prodrugs Subject to Tunable Reductive Activation. J. Med. Chem. 2010, 53, 7731–7738. [Google Scholar] [CrossRef] [Green Version]
  13. Tietze, L.E.; Schmuck, K.; Schuster, H.J.; Müller, M.; Schuberth, I. Synthesis and Biological Evaluation of Prodrugs Based on the Natural Antibiotic Duocarmycin for Use in ADEPT and PMT. Chem. Eur. J. 2011, 17, 1922–1929. [Google Scholar] [CrossRef]
  14. Pors, K.; Loadman, P.M.; Shnyder, S.D.; Sutherland, M.; Sheldrake, H.M.; Guino, M.; Kiakos, K.; Hartley, J.A.; Searcey, M.; Patterson, L.H. Modification of the duocarmycin pharmacophore enables CYP1A1 targeting for biological activity. Chem. Commun. 2011, 47, 12062–12064. [Google Scholar] [CrossRef]
  15. Wolfe, A.L.; Duncan, K.K.; Parelkar, N.K.; Weir, S.J.; Vielhauer, G.A.; Boger, D.L. A Novel, Unusually Efficacious Duocarmycin Carbamate Prodrug That Releases No Residual Byproduct. J. Med. Chem. 2012, 55, 5878–5886. [Google Scholar] [CrossRef] [Green Version]
  16. Stevenson, R.J.; Denny, W.A.; Tercel, M.; Pruijn, F.B.; Ashoorzadeh, A. Nitro seco Analogues of the Duocarmycins Containing Sulfonate Leaving Groups as Hypoxia-Activated Prodrugs for Cancer Therapy. J. Med. Chem. 2012, 55, 2780–2802. [Google Scholar] [CrossRef]
  17. Sheldrake, H.M.; Travica, S.; Johansson, I.; Loadman, P.M.; Sutherland, M.; Elsalem, L.; Illingworth, N.; Cresswell, A.J.; Reuillon, T.; Shnyder, S.D.; et al. Re-engineering of the Duocarmycin Structural Architecture Enables Bioprecursor Development Targeting CYP1A1 and CYP2W1 for Biological Activity. J. Med. Chem. 2013, 56, 6273–6277. [Google Scholar] [CrossRef]
  18. Uematsu, M.; Brody, D.M.; Boger, D.L. A five-membered lactone prodrug of CBI-based analogs of the duocarmycins. Tetrahedron Lett. 2015, 56, 3101–3104. [Google Scholar] [CrossRef] [Green Version]
  19. Giddens, A.C.; Lee, H.H.; Lu, G.-L.; Miller, C.K.; Guo, J.; Phillips, G.D.L.; Pillow, T.H.; Tercel, M. Analogues of DNA minor groove cross-linking agents incorporating aminoCBI, an amino derivative of the duocarmycins: Synthesis, cytotoxicity, and potential as payloads for antibody–drug conjugates. Bioorg. Med. Chem. 2016, 24, 6075–6081. [Google Scholar] [CrossRef]
  20. Menderes, G.; Bonazzoli, E.; Bellone, S.; Black, J.; Altweger, G.; Masserdotti, A.; Pettinella, F.; Zammataro, L.; Buza, N.; Hui, P.; et al. SYD985, a novel duocarmycin-based HER2-targeting antibody-drug conjugate, shows promising antitumor activity in epithelial ovarian carcinoma with HER2/Neu expression. Gynecol. Oncol. 2017, 146, 179–186. [Google Scholar] [CrossRef]
  21. Forbes, I.T.; Ham, P.; Booth, D.H.; Martin, R.T.; Thompson, M.; Baxter, G.S.; Blackburn, T.P.; Glen, A.; Kennett, G.A.; Wood, M.D. 5-Methyl-1-(3-pyridylcarbamoyl)-1,2,3,5-tetrahydropyrrolo[2,3-f]indole: A Novel 5-HT2C/5-HT2B Receptor Antagonist with Improved Affinity, Selectivity, and Oral Activity. J. Med. Chem. 1995, 38, 2524–2530. [Google Scholar] [CrossRef] [PubMed]
  22. Ganton, M.D.; Kerr, M.A. A Domino Amidation Route to Indolines and Indoles: Rapid Syntheses of Anhydrolycorinone, Hippadine, Oxoassoanine, and Pratosine. Org. Lett. 2005, 7, 4777–4779. [Google Scholar] [CrossRef] [PubMed]
  23. Dolenc, D. N-Iodosaccharin—A New Reagent for Iodination of Alkenes and Activated Aromatics. Synlett 2000, 4, 544–546. [Google Scholar]
  24. Harrington, P.J.; Hegedus, L.S. Palladium-catalyzed reactions in the synthesis of 3- and 4-substituted indoles. Approaches to ergot alkaloids. J. Org. Chem. 1984, 49, 2657–2662. [Google Scholar] [CrossRef]
  25. Harrington, P.J.; Hegedus, L.S.; McDaniel, K.F. Palladium-catalyzed reactions in the synthesis of 3- and 4-substituted indoles. 2. Total synthesis of the N-acetyl methyl ester of (+/−)-clavicipitic acids. J. Am. Chem. Soc. 1987, 109, 4335–4338. [Google Scholar] [CrossRef]
  26. Hegedus, L.S.; Toro, J.L.; Miles, W.H.; Harrington, P.J. Palladium-catalyzed reactions in the synthesis of 3- and 4-substituted indoles. 3. Total synthesis of (+/−)-aurantioclavine. J. Org. Chem. 1987, 52, 3319–3322. [Google Scholar] [CrossRef]
  27. Hegedus, L.S.; Sestrick, M.R.; Michaelson, E.T.; Harrington, P.J. Palladium-catalyzed reactions in the synthesis of 3- and 4-substituted indoles. 4. J. Org. Chem. 1989, 54, 4141–4146. [Google Scholar] [CrossRef]
  28. Hellal, M.; Singh, S.; Cuny, G.D. Synthesis of Tetracyclic Indoles via Intramolecular α-Arylation of Ketones. J. Org. Chem. 2012, 77, 4123–4130. [Google Scholar] [CrossRef]
  29. Chen, K.X.; Vibulbhan, B.; Yang, W.; Sannigrahi, M.; Velazquez, F.; Chan, T.-Y.; Venkatraman, S.; Anilkumar, G.N.; Zeng, Q.; Bennet, F.; et al. Structure–Activity Relationship (SAR) Development and Discovery of Potent Indole-Based Inhibitors of the Hepatitis C Virus (HCV) NS5B Polymerase. J. Med. Chem. 2012, 55, 754–765. [Google Scholar] [CrossRef]
Figure 1. Structures of duocarmycin A, SA, and the duocarmycin prodrug with its activation by site-selective CYP2W1 oxidation.
Figure 1. Structures of duocarmycin A, SA, and the duocarmycin prodrug with its activation by site-selective CYP2W1 oxidation.
Molecules 28 04818 g001
Figure 3. Retrosynthetic analysis.
Figure 3. Retrosynthetic analysis.
Molecules 28 04818 g003
Scheme 1. Buchwald–Hartwig amination and subsequent regioselective halogenations.
Scheme 1. Buchwald–Hartwig amination and subsequent regioselective halogenations.
Molecules 28 04818 sch001
Scheme 2. Synthesis of the duocarmycin-type prodrug, * denotes the chiral center.
Scheme 2. Synthesis of the duocarmycin-type prodrug, * denotes the chiral center.
Molecules 28 04818 sch002
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

Bengtsson, C.; Gravenfors, Y. Rapid Construction of a Chloromethyl-Substituted Duocarmycin-like Prodrug. Molecules 2023, 28, 4818. https://doi.org/10.3390/molecules28124818

AMA Style

Bengtsson C, Gravenfors Y. Rapid Construction of a Chloromethyl-Substituted Duocarmycin-like Prodrug. Molecules. 2023; 28(12):4818. https://doi.org/10.3390/molecules28124818

Chicago/Turabian Style

Bengtsson, Christoffer, and Ylva Gravenfors. 2023. "Rapid Construction of a Chloromethyl-Substituted Duocarmycin-like Prodrug" Molecules 28, no. 12: 4818. https://doi.org/10.3390/molecules28124818

Article Metrics

Back to TopTop