Next Article in Journal
Mechanism-based Enzyme Inactivators of Phytosterol Biosynthesis
Previous Article in Journal / Special Issue
New Imide 5-HT1A Receptor Ligands – Modification of Terminal Fragment Geometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biaryl Product Formation from Cross-coupling in Palladiumcatalyzed Borylation of a Boc Protected Aminobromoquinoline Compound

Department of Chemistry, Georgia State University, 33 Gilmer St. S.E., Atlanta, GA 30303 USA
*
Author to whom correspondence should be addressed.
Molecules 2004, 9(3), 178-184; https://doi.org/10.3390/90300178
Submission received: 18 February 2004 / Accepted: 20 February 2004 / Published: 28 February 2004
(This article belongs to the Special Issue Biologically Relevant Heterocyclic Compounds)

Abstract

:
The palladium catalyzed borylation of a Boc protected aminobromoquinoline compound with bis(pinacolato)diboron yielded a biaryl compound, resulting from cross coupling, as the major product, instead of the intended boronate, even though no strong base was used. Such results indicate that under certain conditions and with certain substrates, cross coupling can be a major problem during borylation, leading to unintended consequences.

Introduction

Boronic acids compounds have generated much attention in both chemistry and biology in the last decade because of their importance in organic chemistry [1,2,3,4,5,6,7], medicinal chemistry [8], and sensor development [9,10,11]. Boronic acids have been widely used for saccharide sensor design because of their unique high affinity and reversible complexation with diols, which are commonly found on saccharides [12,13]. Our lab has been engaged in the search for fluorescent sensors for saccharides for various applications [8,9,14,15,16,17,18,19,20,21,22]. Along this line, we are especially interested in the design and synthesis of fluorescent boronic acids that respond to the binding of saccharides by large changes in fluorescent intensities and/or wavelengths, and are water-soluble [18,22,23]. The ultimate goal is to use such reporter compounds for the construction of fluorescent sensors for cell surface saccharides for biological applications [8,17,19,20]. One specific example is 8-quinoline boronic acid (8-QBA, 1), which shows a 40-fold fluorescence intensity change upon binding to a saccharide [18]. However, to incorporate this reporter compound into a diboronic acid sensor, we need to have a way to functionalize it so that various coupling reactions can be used to tether 8-QBA to other groups for improved specificity and affinity. Therefore, we were interested in the synthesis of 2, which has a protected amino group that can be used for further functionalization. Herein, we report the formation of a somewhat unexpected high level of cross-coupling product during the synthesis of 2 using a palladium catalyst, which can cause problems in the synthesis of arylboronic acids.
Molecules 09 00178 i001

Results and Discussion

Generally, there are two types of reactions used for the synthesis of arylboronic acids or arylboronates. One is the transmetalation between an arylmetal and a boron halide or alkoxide [24]; the other is the recently developed PdCl2(dppf)-catalyzed borylation of aryl halides [25,26], triflates [27] or diazonium salts [28] with a tetra(alkoxy)diboron [25,27] or dialkoxyborane reagent [26]. The first method was used to prepare 8-QBA in 1959 [29], but is not compatible with certain sensitive functional groups such as nitro, amide, amine, and hydroxy. The latter coupling reaction can tolerate various functional groups and has been widely applied to the synthesis of arylboronic esters in one step [30]. Therefore, the palladium-catalyzed cross-coupling reaction was chosen for the preparation of monoboronic acid 2 (Scheme 1).
Scheme 1.
Scheme 1.
Molecules 09 00178 g001
The synthesis of compound 2 started from 3, that was prepared by reducing 4-bromo-3-nitrotoluene with stannous chloride [31]. A Skraup reaction was carried out using a modified procedure to give quinoline compound 4 [32]. Bromination of the methyl group using NBS in refluxing benzene under irradiation with a tungsten light gave 5 in 70% yield. The amination of 5 with methylamine in THF gave 6. This was followed by the protection of the amino group by the reaction with di-tert-butyldicarbonate [(Boc)2O] in MeOH in the presence of triethylamine (TEA) to give 7.
The intended borylation to give 9 was carried out by following literature procedures using PdCl2(dppf) as the catalyst in the presence of KOAc [25]. Unexpectedly, the major product (65%) obtained from this palladium-catalyzed reaction was the biaryl product 8, not the intended boronic acid ester 9. The structure of compound 8 was characterized by 1H-NMR and mass spectrometry. Biaryl byproducts in palladium-mediated borylations usually result from the further Suzuki coupling reaction of the arylboronate product with the starting material, bromoarene, in the presence of a base such as K3PO4 and K2CO3. Generally speaking, potassium acetate is not a strong enough base to promote the Suzuki cross-coupling reaction to a significant extent in palladium-catalyzed borylation [25]. Our results indicate that under certain conditions and with certain substrates, such cross coupling reactions can be the major reaction, leading to unintended consequences. In addition, it has been reported that the 8-qunioline boronate can be synthesized by using palladium-catalyzed borylation with arene triflates [26,27]. This suggests that the corresponding 8-quinoline triflate analog of 7 could be used as new substrate to perform palladium catalyzed borylation in the future in order to avoid this dimerization.

Conclusions

Suzuki cross coupling can be the major side reaction in PdCl2(dppf) -mediated borylation of haloarenes. This is true even if only a weak base such as KOAc is used. Further studies are needed to better understand the scope that different factors affect this reaction, and how the borylation yield can be improved.

Experimental

General

Commercially available reagents were used without additional purification unless otherwise indicated. Analytical thin-layer chromatography (TLC) was performed with plastic-backed TLC silica gel 60 F hard layer plates. Flash chromatography was performed with silica gel (flash, 32–63 µm). Mass spectrometry (MS) analyses were performed by the Mass Spectrometry Facility at Georgia State University on a Finnigan electrospray/ion trap mass spectrometer. 1H-NMR spectra were recorded at 300 MHz on a Varian Gemini instrument.

8-Bromo-5-methylquinoline (4)

A solution of 2-bromo-5-methylaniline (3, 6.8 g, 36.5 mmol), glycerol (6.7 g, 72.5 mmol), nitrobenzene (4.5 g, 36 mmol) in 75% H2SO4 (20 mL) was heated at 150 °C for 3 h. The solution was neutralized with NaOH after cooling to room temperature, and then extracted with EtOAc (3 × 120 mL). The combined organic layers was washed with saturated brine, and then dried with MgSO4. After removing solvents under reduced pressure, the crude product was purified by flash column chromatography (hexane/EtOAc, 20:1) to give 4 (5.38 g, 67%): 1H-NMR (CD3OD) δ 8.93 (m, 1H, Ar-H), 8.56 (m, 1H, Ar-H), 7.98 (m, 1H, Ar-H), 7.63 (m, 1H, Ar-H), 7.34 (m, 1H, Ar-H), 2.69 (s, 3H, Ar-CH3); ESI-MS m/z 222/224 (C10H9BrN, M+1).

8-Bromo-5-bromomethylquinoline (5).

A mixture of 8-bromo-5-methylquinoline (4, 5 g, 22.7 mmol) and N-bromosuccinimide (4.9 g, 27.5 mmol) in benzene (55 mL) was refluxed under irradiation with a tungsten light for 12 h. The solution was filtered, and the solvent was evaporated under reduced pressure. The residue was purified by flash column chromatography (hexane/EtOAc, 20:1) to give 5 (4.8 g, 70%); 1H-NMR (CD3OD) δ 8.98 (m, 1H, Ar-H), 8.73 (m, 1H, Ar-H), 8.06 (m, 1H, Ar-H) 7.70 (m, 1H, Ar-H), 7.60 (m, 1H, Ar-H), 5.01 (s, 2H, Ar-CH2-Br); ESI-MS m/z 300/302/304 (C10H8Br2N, M+1).

(8-Bromoquinolin-5-yl-methyl)methylamine (6).

To a solution of 8-bromo-5-bromomethylquinoline (5, 0.39 g, 1.3 mmol) in THF (15 mL) was added 40% methylamine aqueous solution (7 mL). The resulting mixture was stirred at RT for 12 h under nitrogen, and then the organic solvent was removed under reduced pressure. The residue was extracted with DCM (3 × 10 mL) and then washed with saturated brine (30 mL). After drying with MgSO4, evaporation of solvent provided compound 6 (0.31 g, 95%): 1H-NMR (CDCl3) δ 9.05 (m, 1H, Ar-H), 8.56 (m, 1H, Ar-H), 7.99 (m, 1H, Ar-H), 7.49 (m, 1H, Ar-H), 7.38 (m, 1H, Ar-H), 4.16 (s, 2H, Ar-CH2-N), 2.53 (s, 3H, N-CH3); ESI-MS m/z 251/253 (C11H12BrN2 , M+1).

Tert-butyl N-(8-bromoquinolin-5-yl-methyl)-N-methylcarbamate (7).

A mixture of (Boc)2O (1.0 g, 4.61 mmol) in MeOH (5 mL) was added slowly to a solution containing (8-bromo-quinolin-5-yl-methyl)methylamine (6, 0.96 g, 3.84 mmol) and triethylamine (1 mL) in MeOH (20 mL). The resulting solution was further stirred for 12 h, then the methanol was evaporated and the residue was extracted with EtOAc (3 × 10mL). The combined organic extracts were washed with saturated brine (30 mL), dried over MgSO4, and concentrated and dried in vacuo to give compound 7 (1.27 g, 95%): 1H-NMR (CD3OD) δ 8.86(m, 1H, Ar-H), 8.57 (m, 1H, Ar-H), 7.99 (m, 1H, Ar-H), 7.52 (m, 1H, Ar-H), 7.25 (m, 1H, Ar-H), 4.82 (s, 2H, Ar-CH2-N), 2.71(s, 3H, N-CH3), 1.39(s, 9H, -C(CH3)3); ESI-MS m/z 351/353 (C16H20BrN2O2, M+1).

5,5’-Bis[[(tert-butoxycarbonyl)methylamino]methyl]-8,8’-biquinoline (8).

A mixture of PdCl2(dppf) (20 mg, 0.025 mmol), potassium acetate (0.25 g, 2.55 mmol), bis(pinacolato)diboron (0.23 g, 0.94 mmol) and tert-butyl (8-bromoquinolin-5-yl-methyl)-methyl-carbamate (7, 0.30 g, 0.86 mmol) was added to a flask in a glove box under anhydrous condition. After addition of anhydrous DMSO (10 mL) the mixture was stirred at 80 0C for 16 h. The reaction solution was cooled to room temperature and poured into ice-water. The mixture was extracted with ethyl acetate and the combined organic layers was washed with saturated brine, dried over MgSO4, and concentrated in vacuo. The residue was purified by flash column chromatography (hexane/ethyl acetate, 1:1) to give 8 (0.15 g, 65%): 1H-NMR (DMSO-d6) δ 8.68 (m, 4H, Ar-H), 7.70 (m, 2H, Ar-H), 7.52 (m, 4H, Ar-H), 5.00 (s, 2H, Ar-CH2-N), 2.88 (s, 6H, N-CH3), 1.49 (s, 18H, -C(CH3)3); ESI-MS m/z 543 (C32H39N4O4, M+1).

Acknowledgements

Financial support from the National Institutes of Health (NO1-CO-27184 and CA88343), the Georgia Cancer Coalition through a Distinguished Cancer Scientist Award, and the Georgia Research Alliance through an Eminent Scholar endowment is gratefully acknowledged. We also acknowledge Frontier Scientific Corporation for providing certain boronic acid samples in some of our related studies.

References

  1. Martin, R. F.; Haigh, A.; Monger, C.; Pardee, M.; Whittaker, A. D.; Kelly, D. P.; Allen, B. J. Progress in Neutron Capture Therapy for Cancer; AllenB., J., Moore, D. E., Harrington, B. V., Eds.; Plenum Press: New York, 1992; p. 357. [Google Scholar]
  2. Miyaura, N.; Suzuki, A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95, 2457. [Google Scholar]
  3. Ishihara, K.; Yamamoto, H. Arylboron Compounds as Acid Catalysts in Organic Synthetic Transformations. Eur. J. Org. Chem. 1999, 527–538. [Google Scholar]
  4. Latta, R. P.; Springsteen, G.; Wang, B. Development of an Arylboronic Acid-based Solid-Phase Amidation Catalyst. Synthesis 2001, 1611–1613. [Google Scholar]
  5. Petasis, N. A.; Zavialov, I. A. A New and Practical Synthesis of α-Amino Acids from Alkenyl Boronic Acids. J. Am. Chem. Soc. 1997, 119, 445–446. [Google Scholar]
  6. Yu, H.; Wang, B. Phenylboronic Acids Facilitated Selective Reduction of Aldehydes by Tributyltin Hydride. Synth. Commun. 2001, 31, 163–169. [Google Scholar]
  7. Ferrier, R. J. Carbohydrate Boronates. Adv. Cabohydr. Chem. Biochem. 1978, 35, 31. [Google Scholar]
  8. Yang, W.; Gao, X.; Wang, B. Boronic Acid Compounds as Potential Pharmaceutical Agents. Med. Res. Rev. 2003, 23, 346–368. [Google Scholar]
  9. Wang, W.; Gao, X.; Wang, B. Boronic Acid-based Sensors for Carbohydrates. Curr. Org. Chem. 2002, 6, 1285–1317. [Google Scholar]
  10. James, T. D.; Shinkai, S. Artificial Receptors as Chemosensors for Carbohydrates. Top. Curr. Chem. 2002, 218, 159–200. [Google Scholar]
  11. Shinkai, S.; Takeuchi, M. Molecular Design of Artificial Sugar Sensing Systems. Trend in Anal. Chem. 1996, 15, 188–193. [Google Scholar]
  12. Lorand, J. P.; Edwards, J. O. Polyol Complexes and Structure of the Benzeneboronate Ion. J. Org. Chem. 1959, 24, 769. [Google Scholar]
  13. Springsteen, G.; Wang, B. A Detailed Examination of Boronic Acid-Diol Complexation. Tetrahedron 2002, 58, 5291–5300. [Google Scholar]
  14. Wang, W.; Gao, S.; Wang, B. Building Fluorescent Sensors by Template Polymerization: The Preparation of a Fluorescent Sensor for D-Fructose. Org. Lett. 1999, 1, 1209–1212. [Google Scholar]
  15. Liao, Y.; Wang, W.; Wang, B. Building Fluorescent Sensors by Template Polymerization: The Preparation of a Fluorescent Sensor for L-Tryptophan. Bioorg. Chem. 1999, 27, 463–476. [Google Scholar]
  16. Gao, S.; Wang, W.; Wang, B. Building Fluorescent Sensors for Carbohydrates Using Template-directed Polymerizations. Bioorg. Chem. 2001, 29, 308–320. [Google Scholar]
  17. Yang, W.; Gao, S.; Gao, X.; Karnati, V. R.; Ni, W.; Wang, B.; Hooks, W. B.; Carson, J.; Weston, B. Diboronic Acids as Fluorescent Probes for Cells Expressing Sialyl Lewis X. Bioorg. Med. Chem. Lett. 2002, 12, 2175–2177. [Google Scholar]
  18. Yang, W.; Springsteen, G.; Yan, J.; Deeter, S.; Wang, B. A Novel Type of Fluorescent Boronic Acid that Shows Large Fluorescence Intensity Changes upon Binding with a Diol in Aqueous Solution at Physiological pH. Bioorg. Med. Chem. Lett. 2003, 13, 1019–1022. [Google Scholar]
  19. Karnati, V.; Gao, X.; Gao, S.; Yang, W.; Sabapathy, S.; Ni, W.; Wang, B. A Selective Fluorescent Sensor for Glucose. Bioorg. Med. Chem. Lett. 2002, 12, 3373–3377. [Google Scholar]
  20. Yang, W.; Fan, H.; Gao, S.; Gao, X.; Ni, W.; Karnati, V.; Hooks, W. B.; Carson, J.; Weston, B.; Wang, B. The First Fluorescent Diboronic Acid Sensor Specific for Hepatocellular Carcinoma Cells Expressing Sialyl Lewis X. Chem. Biol. 2004. manuscript accepted. [Google Scholar]
  21. Yang, W.; Yan, J.; Fang, H.; Wang, B. The First Fluorescent Sensor for D-Glucarate Based on the Cooperative Action of Boronic Acid and Guanidinium Groups. Chem. Commun. 2003, 792–793. [Google Scholar]
  22. Gao, X.; Zhang, Y.; Wang, B. New Boronic Acid Fluorescent Reporter Compounds II. A Naphthalene-based Sensor Functional at Physiological pH. Org. Lett. 2003, 5, 4615–4618. [Google Scholar]
  23. Ni, W.; Fang, H.; Springsteen, G.; Wang, B. Design of Boronic Acid Spectroscopic Reporter Compounds by Taking Advantage of the pKa-Lowering Effect of Diol-binding: Nitrophenol-based Color Reporters for Diols. J. Org. Chem. 2004. manuscript accepted. [Google Scholar]
  24. Miyaura, N.; Maruoka, K. Synthesis of Organometallic Compounds; Komiya, S., Ed.; Wiley: New York, 1997. [Google Scholar]
  25. Ishiyama, T.; Murata, M.; Miyaura, N. Palladium(0)-Catalyzed Cross-Coupling Reaction of Alkoxydiboron with Haloarenes: A Direct Procedure for Arylboronic Esters. J. Org. Chem. 1995, 60, 7508–7510. [Google Scholar]
  26. Murata, M.; Oyama, T.; Watanabe, S.; Masuda, Y. Palladium-catalyzed Borylation of Aryl Halides or Triflates with Dialkoxyborane: A Novel and Facile Synthetic Route to Arylboronates. J. Org. Chem. 2000, 65, 164–168. [Google Scholar]
  27. Ishiyama, T.; Itoh, Y.; Kitano, T.; Miyaura, N. Synthesis of Arylboronates via the Palladium(0)-catalyzed Cross-coupling Reaction of Tetra(alkoxo)diborons with Aryltriflates. Tetrahedron Lett. 1997, 38, 3447–3450. [Google Scholar]
  28. Willis, D. M.; Strongin, R. M. Palladium-catalyzed Borylation of Aryldiazonium Tetrafluoroborate Salts. A New Synthesis of Arylboronic Esters. Tetrahedron Lett. 2000, 41, 8683–8686. [Google Scholar]
  29. Letsinger, R. L.; Dandegaonker, S. H. Organoboron Compounds. IX. 8-Quinoline boronic Acid, its Preparation and Influence on Reactions of Chlorohydrins. J. Am. Chem. Soc. 1959, 81, 498–501. [Google Scholar]
  30. Ishiyama, T.; Miyaura, N. Chemistry of Group 13 Element-transition Metal Linkage-the Platinum and Palladium-catalyzed Reactions of (Alkoxo)diborons. J. Organomet. Chem. 2000, 611, 392–402. [Google Scholar]
  31. Bellamy, F. D.; Ou, K. Selective Reduction of Aromatic Nitro Compounds with Stannous Chloride in non Acidic and non Aqueous Medium. Tetrahedron Lett. 1984, 25, 839–842. [Google Scholar]
  32. Elderfield, R. C.; Krueger, G. L. Synthesis of Bz-Polymethoxy-8-Aminoquinolines and some Derivatives Thereof. J. Org. Chem. 1952, 17, 358–370. [Google Scholar]
  • Samples Availability: Available from the authors.

Share and Cite

MDPI and ACS Style

Fang, H.; Yan, J.; Wang, B. Biaryl Product Formation from Cross-coupling in Palladiumcatalyzed Borylation of a Boc Protected Aminobromoquinoline Compound. Molecules 2004, 9, 178-184. https://doi.org/10.3390/90300178

AMA Style

Fang H, Yan J, Wang B. Biaryl Product Formation from Cross-coupling in Palladiumcatalyzed Borylation of a Boc Protected Aminobromoquinoline Compound. Molecules. 2004; 9(3):178-184. https://doi.org/10.3390/90300178

Chicago/Turabian Style

Fang, Hao, Jun Yan, and Binghe Wang. 2004. "Biaryl Product Formation from Cross-coupling in Palladiumcatalyzed Borylation of a Boc Protected Aminobromoquinoline Compound" Molecules 9, no. 3: 178-184. https://doi.org/10.3390/90300178

Article Metrics

Back to TopTop