Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 1266 2023-01-09 13:55:46 |
2 update references and layout -2 word(s) 1264 2023-01-10 02:28:19 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Silva, E.M.D.;  Vidal, H.D.A.;  Januário, M.A.P.;  Corrêa, A.G. Asymmetric Synthesis of BINOL Derivatives. Encyclopedia. Available online: https://encyclopedia.pub/entry/39919 (accessed on 04 July 2024).
Silva EMD,  Vidal HDA,  Januário MAP,  Corrêa AG. Asymmetric Synthesis of BINOL Derivatives. Encyclopedia. Available at: https://encyclopedia.pub/entry/39919. Accessed July 04, 2024.
Silva, Everton Machado Da, Hérika Danielle Almeida Vidal, Marcelo Augusto Pereira Januário, Arlene Gonçalves Corrêa. "Asymmetric Synthesis of BINOL Derivatives" Encyclopedia, https://encyclopedia.pub/entry/39919 (accessed July 04, 2024).
Silva, E.M.D.,  Vidal, H.D.A.,  Januário, M.A.P., & Corrêa, A.G. (2023, January 09). Asymmetric Synthesis of BINOL Derivatives. In Encyclopedia. https://encyclopedia.pub/entry/39919
Silva, Everton Machado Da, et al. "Asymmetric Synthesis of BINOL Derivatives." Encyclopedia. Web. 09 January, 2023.
Asymmetric Synthesis of BINOL Derivatives
Edit

The chirality resulting from restricted rotation around a single bond is called atropisomerism (axial chirality). This phenomenon was first described by Christie and Kenner in 1922 when investigating the biaryl 6,6’-dinitro-2,2’-diphenic acid, and the term “atropisomer”, derived from the Greek where “a” means “not” and “tropes” means “turn”, was created by Kuhn. Atropisomers belong to the class of axially chiral compounds; however, here, the enantiomers exist due to restricted rotation around a single bond. Axial chirality has also been considered as an important structural element of many natural products and bioactive compounds, whose enantiomers generally exhibit different pharmacological activities and metabolic processes in vivo and in vitro.

BINOL chirality transfer metal-mediated enantioselective coupling organocatalysis

1. Metal-Mediated Oxidative Enantioselective Coupling

Among the available methods for the synthesis of optically active BINOLs, one of the most explored is the oxidative dimerization of 2-naphthols mediated by complexes of Cu [1][2][3][4][5][6][7][8] Fe [9][10][11], V [12][13][14][15][16][17], Ru [18], and chiral ligands (very often amines), normally generated in situ. In this regard, excellent reviews discussing these methods have been reported by Brunel [19], Wang [20], Bryliakov et al. [21], and Liao et al. [22] (Scheme 1 and Scheme 2).
Scheme 1. Chiral Cu-amine catalytic systems employed in the construction of BINOL derivatives with axial chirality [1][2][3][4][5][6][7][8][9].
Scheme 2. Chiral V, Fe, and Ru catalytic systems employed in the construction of optically pure C2-symmetric BINOL derivatives [12][13][14][15][16][17][18][19][20][21].
Oxidative coupling may occur through three different mechanisms: (1) radical–radical coupling, (2) heterolytic coupling of cationic species with 2-naphthol, or (3) radical–anion coupling, the latter generally being the most accepted to support this type of transformation [17][23][24][25][26]. An important step is the one where the catalytic species is complexed with directing groups or coordination assistants at the C3 position of 2-naphthols—notably ester groups [3][7]—which, in many cases, is a “sine qua no” condition for the success of the synthetic protocols (Scheme 3).
Scheme 3. General aspects of the mechanism for aerobic radical–anion coupling of 2-napthols in the presence of metallic chiral complex catalysts (Mn+). DG: coordination aids. Adapted from Brunel et al [19].
With due recognition of the particularities of each case, the radial–anion mechanism [17][23][24][25][26][27] (Scheme 3) usually proceeds via generation of the radical species B resulting from an oxidation of 2-naphthol A by a metal catalyst (Mn+). B is then added to another neutral 2-naphthol molecule to form a new C-C bond and generate the C-radical, which is further oxidized by O2 to restore aromaticity.
The most recent methods for obtaining enantiomerically pure BINOLs are still based on the catalytic dyad metal–chiral ligands. In this sense, Chen and colleagues [28] (Scheme 4) have developed a new chiral 1,5-N,N-bidentate ligand based on a spirocyclic skeleton of pyrrolidine oxazoline and CuBr to couple 2-naphtols 3b. The efficient catalytic species formed in situ allows for (S)-BINOL derivatives (1) with high enantioselectivity (up to 99% ee) and good yields (up to 87%) to be obtained. Based on experimental results and the literature, the authors proposed that this coupling proceeds via radical–anion coupling, where the complex generated in situ coordinates to form species D in the presence of air, which couples with radical E (generated through an electron transfer from the outer sphere with another Cu(II) complex) to form the intermediate F (Scheme 5). The coupling product is obtained after tautomerization of H. The authors found experimental evidence that, during the coupling process, the attack from E to F by the Si face was favored, probably because of the greater steric impediment to attack by the Re face.
Scheme 4. Recent synthetic protocols for the construction of axially chiral BINOL derivatives.
Scheme 5. Mechanistic proposal for the enantioselective aerobic coupling of 2-naphthols based on a ligand catalytic system containing a spirocyclic skeleton of pyrrolidine oxazoline/CuBr.
Che and co-workers introduced a chiral aminopyridine-like ligand—bisquinolyldiamine [(1R,2R)-N1,N2-di(quinolin-8-yl)cyclohexane-1,2-diamine (BQCN)]— and applied it to the iron-catalyzed asymmetric cis-dihydroxylation of alkenes [29]. Inspired by this work, Liu’s group [30] (Scheme 6) established a methodology for the asymmetric oxidative homo-coupling of 2-naphthols (3c), leading to the synthesis of (S)-BINOL derivatives (1) mediated by a Fe complex and generated in situ from Fe(ClO4)2 and the BQCN ligand. Excellent yields (up to 99%) and enantiomeric excesses (up to 81%) have been reported.
Scheme 6. Enantioselective coupling between 2-naphthols (3c) mediated by an iron/bisquinolyldiamine ligand complex.
From the same perspective, Uchida’s group [31] (Scheme 7) developed remarkable enantioselective aerobic coupling between 2-naphthols 3d in the presence of the (aqua)ruthenium complex (salen). The protocol provided (R)-BINOLs (1) with yields between 55 and 85% and enantiomeric excesses up to 94%. Through mechanistic studies, these researchers concluded that, in this case, cross-coupling selectivity is dominated by steric rather than electronic effects, which can be controlled by chemoselective oxidation via single electron transfer (SET) and oxidative carbon–carbon bond formation, a process for which ruthenium(salen) catalyst proved to be suitable [24]. Therefore, the authors have proposed that this transformation proceeds via oxidation of one of the coupling partners to the electrophilic intermediate radical I, which is converted to the desired BINOL after chemoselective coupling [24].
Scheme 7. (Aqua)ruthenium (salen) catalyzed enantioselective aerobic coupling between 2-naphthols for access to C1-symmetric BINOL derivatives.
Recently, Subramanian et al. [32] (Scheme 8) developed a Cu(II)-2+4-μ4-oxo tetranuclear open frame macrocyclic/BINAN complex and employed it in the asymmetric oxidative coupling of 2-naphthols 3e, obtaining (R)-BINOL derivatives (1) with good to excellent yields (70–96%) and enantiomeric excesses between 68 and 74%.
Scheme 8. Asymmetric oxidative coupling of 2-naphthols mediated by a macrocyclic Cu(II) complex.
Ishihara and co-workers [33] (Scheme 9) developed a method for enantioselective oxidative coupling of 2-naphthol derivatives 3d in the presence of a chiral Fe(II)-diphosphine oxide complex. The products of interest were obtained with yields up to 98 % and enantiomeric excesses between 60 and 85%.
Scheme 9. Chiral diphosphine oxide-iron(II) complex catalyzed enantioselective aerobic coupling between 2-naphthols to access C1-symmetric BINOL derivatives.
A copper catalyst prepared in situ from a ligand synthesized by the fusion of chelating picolinic acid/substituted BINOLs and CuI was employed in the asymmetric oxidative coupling of 2-naphthols (3e). In this work, published by Zhang et al. [34] (Scheme 10), 6,6’-disubstituted (R)-BINOLs (1) were obtained with yields of up to 89% and excellent enantioselectivities (up to 96% ee). The reaction was accompanied by Mass Spectroscopy, and identification of a peak corresponding to the complex J allowed the authors to propose a mechanism pathway through the transition state K.
Scheme 10. Copper-catalyzed asymmetric oxidative coupling of 2-naphthols for the synthesis of 6,6′- disubstituted BINOLs.
Continuing work involving multifunctional chiral catalysis via double activation, Takizawa’s group [35] developed complexes A-C (Scheme 11)—from VOSO4 and Schiff base ligands generated via condensation of (S)-tert-leucine and 3,3 ‘-formyl-(R)-BINOL—which have been successfully applied in the synthesis of (R)- and (S)-BINOL (1) with yields between 46 and 76%, in addition to enantiomeric excesses of up to 91%.
Scheme 11. Atroposelective synthesis of (R)- and (S)-BINOLs (1) via mono- and binuclear vanadium catalysts.

2. Electrochemical Synthesis

Despite the inherent advantages of electrochemical synthesis, notably in terms of sustainability [36], few examples of enantioselective coupling for the construction of chiral BINOLs have been reported so far. In 1994, which appears to be the first record of this type of synthesis, Bobbitt et al. [37] (Scheme 12) established a method for enantioselective coupling of 2-naphthols (3f) on a TEMPO-modified graphite electrode in the presence of (-)-sparteine in acetonitrile to afford (S)-BINOL (1) with excellent yields and enantiomeric excesses.
Scheme 12. Electrochemical synthesis of (S)-BINOL (1) using a TEMPO-modified graphite electrode.
Recently, Mei’s group [36] (Scheme 13) demonstrated the first example of a Ni-catalyzed enantioselective electrochemical reductive coupling of 2-naphtols (4) in an undivided cell for the construction of axially chiral BINOL derivatives (1) with good yields (up 91%) and enantiomeric excess of up to 98%.
Scheme 13. Enantioselective Ni-promoted electrochemical synthesis of (R)-BINOL derivatives (1).

References

  1. Feringa, B.; Wynberg, H. Biomimetic Asymmetric Oxidative Coupling of Phenols. Bioorg. Chem. 1978, 7, 397–408.
  2. Brussee, J.; Groenendijk, J.L.G.; te Koppele, J.M.; Jansen, A.C.A. On the Mechanism of the Formation of s(−)-(1, 1’-Binaphthalene)-2,2’-Diol via Copper(II)Amine Complexes. Tetrahedron 1985, 41, 3313–3319.
  3. Nakajima, M.; Miyoshi, I.; Kanayama, K.; Hashimoto, S.; Noji, M.; Koga, K. Enantioselective Synthesis of Binaphthol Derivatives by Oxidative Coupling of Naphthol Derivatives Catalyzed by Chiral Diamine Copper Complexes. J. Org. Chem. 1999, 64, 2264–2271.
  4. Caselli, A.; Giovenzana, G.B.; Palmisano, G.; Sisti, M.; Pilati, T. Synthesis of C2-Symmetrical Diamine Based on (1R)-(+)-Camphor and Application to Oxidative Aryl Coupling of Naphthols. Tetrahedron Asymmetry 2003, 14, 1451–1454.
  5. Gao, J.; Reibenspies, J.H.; Martell, A.E. Structurally Defined Catalysts for Enantioselective Oxidative Coupling Reactions. Angew. Chem. Int. Ed. 2003, 42, 6008–6012.
  6. Li, X.; Hewgley, J.B.; Mulrooney, C.A.; Yang, J.; Kozlowski, M.C. Enantioselective Oxidative Biaryl Coupling Reactions Catalyzed by 1,5-Diazadecalin Metal Complexes: Efficient Formation of Chiral Functionalized BINOL Derivatives. J. Org. Chem. 2003, 68, 5500–5511.
  7. Kim, K.H.; Lee, D.-W.; Lee, Y.-S.; Ko, D.-H.; Ha, D.-C. Enantioselective Oxidative Coupling of Methyl 3-Hydroxy-2-Naphthoate Using Mono-N-Alkylated Octahydrobinaphthyl-2,2′-Diamine Ligand. Tetrahedron 2004, 60, 9037–9042.
  8. Temma, T.; Habaue, S. Highly Selective Oxidative Cross-Coupling of 2-Naphthol Derivatives with Chiral Copper(I)–Bisoxazoline Catalysts. Tetrahedron Lett. 2005, 46, 5655–5657.
  9. Zhang, Q.; Cui, X.; Chen, L.; Liu, H.; Wu, Y. Syntheses of Chiral Ferrocenophanes and Their Application to Asymmetric Catalysis. Eur. J. Org. Chem. 2014, 2014, 7823–7829.
  10. Egami, H.; Katsuki, T. Iron-Catalyzed Asymmetric Aerobic Oxidation: Oxidative Coupling of 2-Naphthols. J. Am. Chem. Soc. 2009, 131, 6082–6083.
  11. Tkachenko, N.V.; Lyakin, O.Y.; Samsonenko, D.G.; Talsi, E.P.; Bryliakov, K.P. Highly Efficient Asymmetric Aerobic Oxidative Coupling of 2-Naphthols in the Presence of Bioinspired Iron Aminopyridine Complexes. Catal. Commun. 2018, 104, 112–117.
  12. Hon, S.-W.; Li, C.-H.; Kuo, J.-H.; Barhate, N.B.; Liu, Y.-H.; Wang, Y.; Chen, C.-T. Catalytic Asymmetric Coupling of 2-Naphthols by Chiral Tridentate Oxovanadium(IV) Complexes. Org. Lett. 2001, 3, 869–872.
  13. Chu, C.-Y.; Hwang, D.-R.; Wang, S.-K.; Uang, B.-J. Chiral Oxovanadium Complex Catalyzed Enantioselective Oxidative Coupling of 2-Naphthols. Chem. Commun. 2001, 11, 980–981.
  14. Barhate, N.B.; Chen, C.-T. Catalytic Asymmetric Oxidative Couplings of 2-Naphthols by Tridentate N-Ketopinidene-Based Vanadyl Dicarboxylates. Org. Lett. 2002, 4, 2529–2532.
  15. Luo, Z.; Liu, Q.; Gong, L.; Cui, X.; Mi, A.; Jiang, Y. The Rational Design of Novel Chiral Oxovanadium(IV) Complexes for Highly Enantioselective Oxidative Coupling of 2-Naphthols. Chem. Commun. 2002, 8, 914–915.
  16. Somei, H.; Asano, Y.; Yoshida, T.; Takizawa, S.; Yamataka, H.; Sasai, H. Dual Activation in a Homolytic Coupling Reaction Promoted by an Enantioselective Dinuclear Vanadium(IV) Catalyst. Tetrahedron Lett. 2004, 45, 1841–1844.
  17. Guo, Q.-X.; Wu, Z.-J.; Luo, Z.-B.; Liu, Q.-Z.; Ye, J.-L.; Luo, S.-W.; Cun, L.-F.; Gong, L.-Z. Highly Enantioselective Oxidative Couplings of 2-Naphthols Catalyzed by Chiral Bimetallic Oxovanadium Complexes with Either Oxygen or Air as Oxidant. J. Am. Chem. Soc. 2007, 129, 13927–13938.
  18. Tanaka, H.; Nishikawa, H.; Uchida, T.; Katsuki, T. Photopromoted Ru-Catalyzed Asymmetric Aerobic Sulfide Oxidation and Epoxidation Using Water as a Proton Transfer Mediator. J. Am. Chem. Soc. 2010, 132, 12034–12041.
  19. Brunel, J.M. BINOL: A Versatile Chiral Reagent. Chem. Rev. 2005, 105, 857–898.
  20. Wang, H. Recent Advances in Asymmetric Oxidative Coupling of 2-Naphthol and Its Derivatives. Chirality 2010, 22, 827–837.
  21. Tkachenko, N.V.; Bryliakov, K.P. Transition Metal Catalyzed Aerobic Asymmetric Coupling of 2-Naphthols. Mini. Rev. Org. Chem. 2019, 16, 392–398.
  22. Liao, G.; Zhou, T.; Yao, Q.-J.; Shi, B.-F. Recent Advances in the Synthesis of Axially Chiral Biaryls via Transition Metal-Catalysed Asymmetric C–H Functionalization. Chem. Commun. 2019, 55, 8514–8523.
  23. Matsushita, M.; Kamata, K.; Yamaguchi, K.; Mizuno, N. Heterogeneously Catalyzed Aerobic Oxidative Biaryl Coupling of 2-Naphthols and Substituted Phenols in Water. J. Am. Chem. Soc. 2005, 127, 6632–6640.
  24. Egami, H.; Matsumoto, K.; Oguma, T.; Kunisu, T.; Katsuki, T. Enantioenriched Synthesis of C1-Symmetric BINOLs: Iron-Catalyzed Cross-Coupling of 2-Naphthols and Some Mechanistic Insight. J. Am. Chem. Soc. 2010, 132, 13633–13635.
  25. Narute, S.; Parnes, R.; Toste, F.D.; Pappo, D. Enantioselective Oxidative Homocoupling and Cross-Coupling of 2-Naphthols Catalyzed by Chiral Iron Phosphate Complexes. J. Am. Chem. Soc. 2016, 138, 16553–16560.
  26. Libman, A.; Shalit, H.; Vainer, Y.; Narute, S.; Kozuch, S.; Pappo, D. Synthetic and Predictive Approach to Unsymmetrical Biphenols by Iron-Catalyzed Chelated Radical–Anion Oxidative Coupling. J. Am. Chem. Soc. 2015, 137, 11453–11460.
  27. Shibasaki, M.; Matsunaga, S. Design and Application of Linked-BINOL Chiral Ligands in Bifunctional Asymmetric Catalysis. Chem. Soc. Rev. 2006, 35, 269.
  28. Tian, J.; Wang, A.; Yang, J.; Zhao, X.; Tu, Y.; Zhang, S.; Chen, Z. Copper-Complex-Catalyzed Asymmetric Aerobic Oxidative Cross-Coupling of 2-Naphthols: Enantioselective Synthesis of 3,3′-Substituted C1 -Symmetric BINOLs. Angew. Chem. 2019, 131, 11139–11143.
  29. Zang, C.; Liu, Y.; Xu, Z.-J.; Tse, C.-W.; Guan, X.; Wei, J.; Huang, J.-S.; Che, C.-M. Highly Enantioselective Iron-Catalyzed Cis -Dihydroxylation of Alkenes with Hydrogen Peroxide Oxidant via an Fe III -OOH Reactive Intermediate. Angew. Chem. 2016, 128, 10409–10413.
  30. Wu, L.Y.; Usman, M.; Liu, W.B. Enantioselective Iron/Bisquinolyldiamine Ligand-catalyzed Oxidative Coupling Reaction of 2-naphthols. Molecules 2020, 25, 852.
  31. Hayashi, H.; Ueno, T.; Kim, C.; Uchida, T. Ruthenium-Catalyzed Cross-Selective Asymmetric Oxidative Coupling of Arenols. Org. Lett. 2020, 22, 1469–1474.
  32. Chinnaraja, E.; Arunachalam, R.; Pillai, R.S.; Peuronen, A.; Rissanen, K.; Subramanian, P.S. One-Pot Synthesis of -Helicate-like Macrocycle and 2+4-Μ4-Oxo Tetranuclear Open Frame Complexes: Chiroptical Properties and Asymmetric Oxidative Coupling of 2-Naphthols. Appl. Organomet. Chem. 2020, 34, e5666.
  33. Horibe, T.; Nakagawa, K.; Hazeyama, T.; Takeda, K.; Ishihara, K. An Enantioselective Oxidative Coupling Reaction of 2-Naphthol Derivatives Catalyzed by Chiral Diphosphine Oxide-Iron(II) Complexes. Chem. Commun. 2019, 55, 13677–13680.
  34. Wang, P.; Cen, S.; Gao, J.; Shen, A.; Zhang, Z. Novel Axially Chiral Ligand-Enabled Copper-Catalyzed Asymmetric Oxidative Coupling of 2-Naphthols for the Synthesis of 6,6′-Disubstituted BINOLs. Org. Lett. 2022, 24, 2321–2326.
  35. Kumar, A.; Sasai, H.; Takizawa, S. Atroposelective Synthesis of C-C Axially Chiral Compounds via Mono- and Dinuclear Vanadium Catalysis. Acc. Chem. Res. 2022, 55, 2949–2965.
  36. Qiu, H.; Shuai, B.; Wang, Y.Z.; Liu, D.; Chen, Y.G.; Gao, P.-S.; Ma, H.X.; Ma, H.X.; Chen, S.; Mei, T.S. Enantioselective Ni-Catalyzed Electrochemical Synthesis of Biaryl Atropisomers. J. Am. Chem. Soc. 2020, 142, 9872–9878.
  37. Osa, T.; Kashiwagi, Y.; Yanagisawa, Y.; Bobbitt, J.M. Enantioselective, Electrocatalytic Oxidative Coupling of Naphthol, Naphthyl Ether and Phenanthrol on a TEMPO-Modified Graphite Felt Electrode in the Presence of (-)-Sparteine (TEMPO = 2,2,6,6-Tetramethylpiperidin-1-Yloxyl). J. Chem. Soc. Chem. Commun. 1994, 1, 2535–2537.
More
Information
Subjects: Chemistry, Organic
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 937
Revisions: 2 times (View History)
Update Date: 10 Jan 2023
1000/1000
Video Production Service