Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Sep 26;88(19):14222–14226. doi: 10.1021/acs.joc.3c01534

Atroposelective Ir-Catalyzed C–H Borylation of Phthalazine Heterobiaryls

Paul Stehrer , Anke Spannenberg , Marko Hapke †,‡,*
PMCID: PMC10563123  PMID: 37751525

Abstract

graphic file with name jo3c01534_0005.jpg

The atroposelective iridium-catalyzed borylation of menthyloxy-substituted phthalazine heterobiaryls with diborons is reported. Utilizing [Ir(OMe)(COD)]2/2-aminopyridine as a rarely used efficient catalyst system, the heterobiaryls were selectively borylated in the 2-position of the carbocycle, exclusively yielding only one of the atropisomers, depending on the substitution of the phthalazine with (+)-menthyl or (−)-menthyl moieties. Exemplary further functionalization of a borylated atropisomer demonstrated that nickel-catalyzed Suzuki-Miyaura cross-coupling with an aryl halide was able to provide stereoretention to a certain degree (up to 75% de).


The position-selective and stereoselective transition metal-catalyzed C–H borylation reaction has evolved into a highly useful methodology for the introduction of functional groups, especially for arene substrates but also for alkyl C–H bonds.1 The approach circumvents the use of functionalized substrates like aryl halides, where the halide can be transformed into the boronate by using organometallic bases (R-Li, Grignard reagents) and quenching with borates, transition metal-catalyzed processes with diborons,2 or other functional group transformations.3 Also, a number of transition metal-free processes have been reported lately, relying on electrophilic borylation.4 Borylation reactions have been used for the synthesis of active pharmaceutical ingredients.5 Within the evolution of the methodology, recently a lot of attention is focusing on developing position-selective borylation reactions for the meta or para position of arenes6,7 or for asymmetric borylation reactions.8

Asymmetric assembling of atropisomers by catalytic methodologies saw a huge advance over the last two decades.9 The synthesis of selectively borylated (hetero)biaryls using guidance by nitrogen atoms has been less developed10 but could deliver highly interesting building blocks for the assembly of functionalized molecules for photochemical applications.11 Substitution in the 2-position of a sufficiently large arene connected to a nitrogen-containing ring as directing group can increase the configurational stability of the biaryl axis. Under ideal conditions, it can be utilized for the preparation of chiral compounds, e.g., as demonstrated for dynamic kinetic resolution via asymmetric Heck reactions12 or the synthesis of enantiopure QUINAP derivatives by asymmetric catalysis.13

The synthesis of heterobiaryls containing other stereocenters beside the chiral axis potentially leads to resolvable diastereomers, which could be separated by, e.g., chromatographic methods.14 Since we are experienced in the de novo synthesis of heterobiaryls by [2+2+2] cycloaddition, our aim was to venture into the selective postfunctionalization of the 2-position of the carbocycles by C–H borylation using chiral diborons like DB1 (see Table 1 for structure). Preliminary experiments suggested phthalazine heterocycles, utilized before as a heterobiaryl backbone in the assembly of PINAP ligands,15 as suitable test substrates for the borylation of the heterobiaryl carbocycle in the 2-position. The test substrates containing a chiral (+)- or (−)-menthol (1) group can be easily assembled from commercially available 1,4-dichloro-phthalazine in three steps via 2a with a final Suzuki-Miyaura coupling (Scheme 1).16 The heterobiaryls (+)-3ak and (−)-3a, b, and e were conveniently obtained by the same protocol with generally very good yields.

Table 1. Screening of Ligands for Directed Borylation Reaction with Substrate 3a.

graphic file with name jo3c01534_0003.jpg

No. Ligand Yield [%]a
1 1,10-phenanthroline (L1) (94)b
2 3,4,7,8-tetramethyl-1,10-phenanthroline (L2) (67)b
3 2,2′-bipyridine (L3) (68)b
4 4,4′-di-tert-butyl-2,2′-bipyridine (L4) (79)b
5 2-aminopyridine (L5) 88
6 2-aminomethylpyridine (L6) n.r.c
7 2-(dimethyamino)-pyridine (L7) 74
8 2-(benzylamino)-pyridine (L8) 87
9 2-(benzylimino)-pyridine (L9) 68
a

Isolated yields.

b

Reaction yielded mixture of indeterminable regioisomerically borylated heterobiaryls.

c

No reaction and quantitative recovery of starting material.

Scheme 1. Exemplary Synthesis of (+)-Menthyloxy 1-naphthyl-phthalazines.

Scheme 1

THF was used as solvent.

In the next step, we investigated the optimal conditions for the Ir-catalyzed borylation of the prepared heterobiaryls using commercially available chiral diboron DB1. Table 1 shows the results of the ligand screening for the Ir-catalyzed process at the optimized temperature of 80 °C. Utilizing 2-aminopyridine (L5) as very rare ligand for C–H borylation gave excellent 88% yield exclusively of the desired regioisomer (contrary to entries 1–4).17 It came as a genuine surprise that only one stereoisomer was found as product, suggesting a sufficient energetical difference between diastereomers on an intermediary catalytic stage. Except for ligand L6, leading to no reaction, all other ligands L7 to L9 gave good to very good yields (entries 7–9, Table 1).

After having identified that the pyridine L5 afforded the most suitable catalyst, we investigated the borylation of different substrates 3ak to evaluate the role of the chiral menthyloxy substitutent as well as the chiral diboron DB1, its enantiomer DB2, and bis(pinacolato)diboron (DB3) in more detail. The results are displayed in Scheme 2. The most intriguing observation is the dependence of the configuration of the biaryl axis from the configuration of the menthyloxy moiety attached to the phthalazine ring while overriding any effect of the chiral diborons DB1/DB2, thus also allowing selective borylations using achiral B2pin2 (DB3).

Scheme 2. Synthesis of Borylated Phthalazine Derivatives with Chiral Diborons DB1 and DB2 and Achiral B2pin2 (DB3).

Scheme 2

Repeated on a 1 mmol scale, providing 4g in 77% yield.

Repeated on a 1 mmol scale, providing 5a in 85% yield.

2.0 equiv of B2pin2 used.

While (+)-menthyloxy substitution led to the (Ra)-configuration (compounds 4a, c, e, f, g, h), the introduction of a (−)-menthyloxy substituent led to the (Sa)-configuration (compounds 4b, d), which was derived from structure determination in the series of heterobiaryls 5 (vide infra). The isolated yields for the heterobiaryls 4 are in general very good and range from 70 to 88% (Scheme 2, left column).18 Replacing the chiral diborons DB1 and DB2 by B2pin2 (DB3) to evaluate the role of the chiral boryl group demonstrated that atroposelective borylations were also occurring in this case (5ak, Scheme 2, right column). The dibenzothiophene 3j and phenanthrene 3k were either not reactive in the borylation (e.g., to 5l) or underwent unselective borylations.19,20 These experiments therefore established that the chiral boron moiety does not exert a direct influence on the selectivity of the C–H borylation process; however, the sterically crowded iridium-boryl complex species might pronounce the effect of the chiral menthyl auxiliar for the borylation process.21 Assignation of the absolute configuration of the biaryl axis was possible by SC-XRD of compound 5c, unambiguously confirming the (Ra)-configuration of the biaryl axis (see Figure S1, Supporting Information).22,23

Finally, we exemplarily investigated the further functionalization of the 2-boronate function from either the 4g or 4k-SI atropisomer by a Suzuki-Miyaura cross-coupling reaction to evaluate the configurational stability of the biaryl axis in the products 6a (Table 2). Such C–C coupling reactions with chiral boronic esters are very rare.24 We turned our attention to nickel-catalyzed cross-coupling reactions at lower reaction temperatures with 4-haloacetophenones to 6a (entries 1–4, Table 2).25 Application of Ni[P(n-Bu)3]2(COD) as nickel precursor together with 1,4-bis(diphenylphosphino)butane (dppb) as ligand increased the yield as well as even more significantly the de value for 6a to 75% as the best case; beneficially, the aryl bromide as well as iodide reacted identically (entries 2–4, Table 2). We investigated lower reaction temperatures to improve the selectivity and found the reaction with these catalysts to be possible even at room temperature, unfortunately without improvement of the de value (entries 5 and 6, Table 2). Shorter reaction times do not improve the stereoretention at lower yields (entries 7 and 8, Table 2). However, the developed catalytic system represents one of the very few nickel-catalyzed Suzuki-Miyaura couplings successfully working at mild reaction temperatures with Bpinane esters.

Table 2. Cross-coupling Reactions with Borylated Atropisomers.

graphic file with name jo3c01534_0004.jpg

No. Atropisomer/Aryl halide/Catalyst system/Conditions Yield [%] Stereoret. [% de]
1 4g, 4-Iodoacetophenone, Ni(COD)2, dppb, 2-MeTHF/H2O, i-Pr2EtN, 50 °C, 12 h 75 25
2 4g, 4-Bromoacetophenone, Ni[P(n-Bu)3]2(COD), dppb, 2-MeTHF/H2O, i-Pr2EtN, 50 °C, 12 h 81 74
3 4g, 4-Iodoacetophenone, Ni[P(n-Bu)3]2(COD), dppb, 2-MeTHF/H2O, i-Pr2EtN, 50 °C, 12 h 82 75
4 4k-SI, 4-Iodoacetophenone, Ni[P(n-Bu)3]2(COD), dppb, 2-MeTHF/H2O, i-Pr2EtN, 50 °C, 12 h 82 67
5 4g, 4-Iodoacetophenone, Ni[P(n-Bu)3]2(COD), 2-MeTHF/H2O, i-Pr2EtN, rt, 12 h 67 72
6 4g, 4-Iodoacetophenone, Ni(dppb)(COD), 2-MeTHF/H2O, i-Pr2EtN, rt, 12 h 85 71
7 4g, 4-Iodoacetophenone, Ni(dppb)(COD), 2-MeTHF/H2O, i-Pr2EtN, 50 °C, 0.5 h 10 70
8 4g, 4-Iodoacetophenone, Ni(dppb)(COD), 2-MeTHF/H2O, i-Pr2EtN, 50 °C, 2 h 33 69

In summary, we could demonstrate the ability of menthyl groups as convenient chiral auxiliaries in menthyloxy-substituted phthalazine heterobiaryls to direct the stereo- and position-selective C–H borylation with achiral and chiral diborons. The chirality of the used diborons did not play a role for the stereoselectivity of the borylation process. A rarely used in situ-generated Ir(I)-2-aminopyridine catalytic system proved to be most efficient for the process. Further functionalization of the atropisomeric 2-aryl-borylated heterobiaryls by nickel-catalyzed Suzuki-Miyaura coupling reaction with aryl halides at room temperature led to up to 75% de and 82% yield for the triaryl coupling product.

Acknowledgments

M.H. thanks the JKU for financial support. We thank Dr. Matthias Bechmann (JKU, Institute of Organic Chemistry) for additional NMR experiments.

Data Availability Statement

The data underlying this study are available in the published article and its Supporting Information.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.3c01534.

  • Experimental procedures, synthesis of starting materials, data of screening experiments, compound characterization data, SC-XRD data for 5c, NMR spectra (PDF)

Author Contributions

P.S. performed the experiments for the manuscript and A.S. performed the X-ray single crystal studies. The manuscript was written through the contributions of all authors. All authors have given approval to the final version of the manuscript.

The authors declare no competing financial interest.

Dedication

Dedicated to the 60th birthday of Prof. Dr. Matthias Beller.

Supplementary Material

References

  1. a Veth L.; Grab H. A.; Dydio P. Recent Trends in Group 9 Catalyzed C–H Borylation Reactions: Different Strategies To Control Site-, Regio-, and Stereoselectivity. Synthesis 2022, 54, 3482–3498. 10.1055/a-1711-5889. [DOI] [Google Scholar]; b Hartwig J. F. Borylation and Silylation of C–H Bonds: A Platform for Diverse C–H Bond Functionalizations. Acc. Chem. Res. 2012, 45, 864–873. 10.1021/ar200206a. [DOI] [PubMed] [Google Scholar]; c Mkhalid I. A. I.; Barnard J. H.; Marder T. B.; Murphy J. M.; Hartwig J. F. C-H Activation for the Construction of C-B Bonds. Chem. Rev. 2010, 110, 890–931. 10.1021/cr900206p. [DOI] [PubMed] [Google Scholar]
  2. Selected recent examples:; a Dong J.; Guo H.; Hu Q.-S. Room Temperature Ni0/PCy3-Catalyzed Coupling Reactions of Aryl Arenesulfonates with Bis(pinacolato)diboron. Eur. J. Org. Chem. 2017, 2017, 7087–7090. 10.1002/ejoc.201701300. [DOI] [Google Scholar]; b Yamamoto Y.; Matsubara H.; Yorimitsu H.; Osuka A. Base-Free Palladium-Catalyzed Borylation of Aryl Chlorides with Diborons. ChemCatChem. 2016, 8, 2317–2320. 10.1002/cctc.201600456. [DOI] [Google Scholar]; c Guerrand H. D. S.; Marciasini L. D.; Jousseaume M.; Vaultier M.; Pucheault M. Borylation of Unactivated Aryl Chlorides under Mild Conditions by Using Diisopropylaminoborane as a Borylating Reagent. Chem. Eur. J. 2014, 20, 5573–5579. 10.1002/chem.201304861. [DOI] [PubMed] [Google Scholar]
  3. Wang M.; Shi Z. Methodologies and Strategies for Selective Borylation of C–Het and C–C Bonds. Chem. Rev. 2020, 120, 7348–7398. 10.1021/acs.chemrev.9b00384. [DOI] [PubMed] [Google Scholar]
  4. a Wang Z. J.; Chen X.; Wu L.; Wong J. J.; Liang Y.; Zhao Y.; Houk K. N.; Shi Z. Metal-Free Directed C–H Borylation of Pyrroles. Angew. Chem., Int. Ed. 2021, 60, 8500–8504. 10.1002/anie.202016573. [DOI] [PubMed] [Google Scholar]; b Sadek O.; Le Gac A.; Hidalgo N.; Mallet-Ladeira S.; Miqueu K.; Bouhadir G.; Bourissou D. Metal-Free Phosphorus-Directed Borylation of C(sp2)–H Bonds Angew. Chem. Int. Ed. 2022, 61, e202110102 10.1002/anie.202110102. [DOI] [PubMed] [Google Scholar]; c Lv J.; Chen X.; Xue X.-S.; Zhao B.; Liang Y.; Wang M.; Jin L.; Yuan Y.; Han Y.; Zhao Y.; Lu Y.; Zhao J.; Sun W.-Y.; Houk K. N.; Shi Z. Metal-Free Directed sp2-C–H Borylation. Nature 2019, 575, 336–340. 10.1038/s41586-019-1640-2. [DOI] [PubMed] [Google Scholar]; d Légaré M.-A.; Courtemanche M.-A.; Rochette É.; Fontaine F.-G. Metal-Free Catalytic C-H Bond Activation and Borylation of Heteroarenes. Science 2015, 349, 513–516. 10.1126/science.aab3591. [DOI] [PubMed] [Google Scholar]
  5. Examples:; a Hickey A.; Merz J.; Al Mamari H. H.; Friedrich A.; Marder T. B.; McGlacken G. P. Iridium-Catalyzed Borylation of 6-Fluoroquinolines: Access to 6-Fluoroquinolones. J. Org. Chem. 2022, 87, 9977–9987. 10.1021/acs.joc.2c00973. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Karlsson S.; Benson H.; Cook C.; Currie G.; Dubiez J.; Emtenäs H.; Hawkins J.; Meadows R.; Smith P. D.; Varnes J. From Milligram to Kilogram Manufacture of AZD4573: Making It Possible by Application of Enzyme-, Iridium-, and Palladium-Catalyzed Key Transformations. Org. Process Res. Dev. 2022, 26, 601–615. 10.1021/acs.oprd.1c00058. [DOI] [Google Scholar]; c Sieser J. E.; Maloney M. T.; Chisowa E.; Brenek S. J.; Monfette S.; Salisbury J. J.; Do N. M.; Singer R. A. Ir-Catalyzed Borylations as an Efficient Route to a Nicotine Hapten. Org. Process Res. Dev. 2018, 22, 527–534. 10.1021/acs.oprd.8b00053. [DOI] [Google Scholar]; d Homer J. A.; Sperry J. A. Short Synthesis of the Endogenous Plant Metabolite 7-Hydroxyoxindole-3-Acetic Acid (7-Oh-OXIAA) Using Simultaneous C–H Borylations. Tetrahedron Lett. 2014, 55, 5798–5800. 10.1016/j.tetlet.2014.08.104. [DOI] [Google Scholar]
  6. Actual overview:Sinha S. K.; Guin S.; Maiti S.; Biswas J. P.; Porey S.; Maiti D. Toolbox for Distal C–H Bond Functionalizations in Organic Molecules. Chem. Rev. 2022, 122, 5682–5841. 10.1021/acs.chemrev.1c00220. [DOI] [PubMed] [Google Scholar]
  7. Selected examples for para-borylation:; a Saito Y.; Yamanoue K.; Segawa Y.; Itami K. Selective Transformation of Strychnine and 1,2-Disubstituted Benzenes by C–H Borylation. Chem. 2020, 6, 985–993. 10.1016/j.chempr.2020.02.004. [DOI] [Google Scholar]; b Saito Y.; Segawa Y.; Itami K. Para-C–H Borylation of Benzene Derivatives by a Bulky Iridium Catalyst. Angew. Chem., Int. Ed. 2015, 137, 5193–5198. 10.1021/jacs.5b02052. [DOI] [PubMed] [Google Scholar]; c Yang L.; Semba K.; Nakao Y. Para-Selective C–H Borylation of (Hetero)Arenes by Cooperative Iridium/Aluminum Catalysis. Angew. Chem., Int. Ed. 2017, 56, 4853–4857. 10.1002/anie.201701238. [DOI] [PubMed] [Google Scholar]; Selected examples for meta-borylation:; d Trouvé J.; Zardi P.; Al-Shehimy S.; Roisnel T.; Gramage-Doria R. Enzyme-like Supramolecular Iridium Catalysis Enabling C–H Bond Borylation of Pyridines with Meta-Selectivity. Angew. Chem., Int. Ed. 2021, 60, 18006–18013. 10.1002/anie.202101997. [DOI] [PubMed] [Google Scholar]; e Yang L.; Uemura N.; Nakao Y. Meta-Selective C–H Borylation of Benzamides and Pyridines by an Iridium–Lewis Acid Bifunctional Catalyst. J. Am. Chem. Soc. 2019, 141, 7972–7979. 10.1021/jacs.9b03138. [DOI] [PubMed] [Google Scholar]
  8. Reviews:; a Su B.; Hartwig J. F. Development of Chiral Ligands for the Transition-Metal-Catalyzed Enantioselective Silylation and Borylation of C–H Bonds. Angew. Chem., Int. Ed. 2022, 61, e202113343 10.1002/anie.202113343. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Ghosh S.; Ghosh A.; Pyne P.; Hajra A. Asymmetric C(sp3)-H borylation. An update. Org. Biomol. Chem. 2022, 20, 4496–4511. 10.1039/D2OB00688J. [DOI] [PubMed] [Google Scholar]; c Hu J.; Ferger M.; Shi Z.; Marder T. B. Recent advances in asymmetric borylation by transition metal catalysis. Chem. Soc. Rev. 2021, 50, 13129–13188. 10.1039/D0CS00843E. [DOI] [PubMed] [Google Scholar]; d Zhan M.; Song P.; Jiao J.; Li P. Novel Chiral Ligands-Enabled Transition-Metal-Catalyzed Asymmetric C—H Borylation. Chin. J. Chem. 2020, 38, 665–667. 10.1002/cjoc.202000023. [DOI] [Google Scholar]
  9. Recent overviews:; a Cheng J. K.; Xiang S.-H.; Li S.; Ye L.; Tan B. Recent Advances in Catalytic Asymmetric Construction of Atropisomers. Chem. Rev. 2021, 121, 4805–4902. 10.1021/acs.chemrev.0c01306. [DOI] [PubMed] [Google Scholar]; b Carmona J. A.; Rodríguez-Franco C.; Fernández R.; Hornillos V.; Lassaletta J. M. Atroposelective transformation of axially chiral (hetero)biaryls. From desymmetrization to modern resolution strategies. Chem. Soc. Rev. 2021, 50, 2968–2983. 10.1039/D0CS00870B. [DOI] [PubMed] [Google Scholar]; c 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. 10.1039/C9CC03967H. [DOI] [PubMed] [Google Scholar]
  10. Examples for N-directed borylations:; a Hoque M. E.; Hassan M. M. M.; Chattopadhyay B. emarkably Efficient Iridium Catalysts for Directed C(sp2)–H and C(sp3)–H Borylation of Diverse Classes of Substrates. J. Am. Chem. Soc. 2021, 143, 5022–5037. 10.1021/jacs.0c13415. [DOI] [PubMed] [Google Scholar]; b Kawamorita S.; Miyazaki T.; Ohmiya H.; Iwai T.; Sawamura M. Rh-Catalyzed Ortho-Selective C-H Borylation of N-Functionalized Arenes with Silica-Supported Bridgehead Monophosphine Ligands. J. Am. Chem. Soc. 2011, 133, 19310–19313. 10.1021/ja208364a. [DOI] [PubMed] [Google Scholar]; c Ros A.; Estepa B.; López-Rodríguez R.; Álvarez E.; Fernández R.; Lassaletta J. M. Use of Hemilabile N,N Ligands in Nitrogen-Directed Iridium-Catalyzed Borylations of Arenes. Angew. Chem., Int. Ed. 2011, 50, 11724–11728. 10.1002/anie.201104544. [DOI] [PubMed] [Google Scholar]
  11. a Pais V. F.; Alcaide M. M.; López-Rodríguez R.; Collado D.; Nájera F.; Pérez-Inestrosa E.; Álvarez E.; Lassaletta J. M.; Fernández R.; Ros A.; Pischel U. Strongly Emissive and Photostable Four-Coordinate Organoboron N,C Chelates and Their Use in Fluorescence Microscopy. Chem. Eur. J. 2015, 21, 15369–15376. 10.1002/chem.201501626. [DOI] [PubMed] [Google Scholar]; b Pais V. F.; Lassaletta J. M.; Fernández R.; El-Sheshtawy H. S.; Ros A.; Pischel U. Organic Fluorescent Thermometers Based on Borylated Arylisoquinoline Dyes. Chem. Eur. J. 2014, 20, 7638–7645. 10.1002/chem.201402027. [DOI] [PubMed] [Google Scholar]; c Pais V. F.; El-Sheshtawy H. S.; Fernández R.; Lassaletta J. M.; Ros A.; Pischel U. Borylated Arylisoquinolines: Photophysical Properties and Switching Behavior of Promising Tunable Fluorophores. Chem. Eur. J. 2013, 19, 6650–6661. 10.1002/chem.201203887. [DOI] [PubMed] [Google Scholar]
  12. Carmona J. A.; Hornillos V.; Ramírez-López P.; Ros A.; Iglesias-Sigüenza J.; Gómez-Bengoa E.; Fernández R.; Lassaletta J. M. Dynamic Kinetic Asymmetric Heck Reaction for the Simultaneous Generation of Central and Axial Chirality. J. Am. Chem. Soc. 2018, 140, 11067–11075. 10.1021/jacs.8b05819. [DOI] [PubMed] [Google Scholar]
  13. a Alcock N. W.; Brown J. M.; Pearson M.; Woodward S. Synthesis and Easy Racemisation of an Atropisomerically Chiral Phosphinamine. Tetrahedron: Asymmetry 1992, 3, 17–20. 10.1016/S0957-4166(00)82305-8. [DOI] [Google Scholar]; b Bhat V.; Wang S.; Stoltz B. M.; Virgil S. C. Asymmetric Synthesis of QUINAP via Dynamic Kinetic Resolution. J. Am. Chem. Soc. 2013, 135, 16829–16832. 10.1021/ja409383f. [DOI] [PubMed] [Google Scholar]; c Han S.-J.; Bhat V.; Stoltz B. M.; Virgil S. C. Atroposelective Synthesis of PINAP via Dynamic Kinetic Asymmetric Transformation. Adv. Synth. Catal. 2019, 361, 441–444. 10.1002/adsc.201801248. [DOI] [Google Scholar]
  14. Fischer F.; Jungk P.; Weding N.; Spannenberg A.; Ott H.; Hapke M. Diastereomeric Atropisomers from a Chiral Diyne by Cobalt(I)-Catalyzed Cyclotrimerization. Eur. J. Org. Chem. 2012, 2012, 5828–5838. 10.1002/ejoc.201200402. [DOI] [Google Scholar]
  15. The approach has been used for the synthesis and separation of QUINAP ligands: see ref (13c).
  16. Suzuki coupling conditions in accordance to:Ros A.; Estepa B.; López-Rodríguez R.; Álvarez E.; Fernández R.; Lassaletta J. M. Use of Hemilabile N,N Ligands in Nitrogen-Directed Iridium-Catalyzed Borylations of Arenes. Angew. Chem., Int. Ed. 2011, 50, 11724–11728. 10.1002/anie.201104544. [DOI] [PubMed] [Google Scholar]
  17. Borylations with 2-aminopyridine are described in:Hale L. V. A.; McGarry K. A.; Ringgold M. A.; Clark T. B. Role of Hemilabile Diamine Ligands in the Amine-Directed C–H Borylation of Arenes. Organometallics 2015, 34, 51–55. 10.1021/om5007837. [DOI] [Google Scholar]
  18. Additional structurally related products 4i-SI to 4l-SI as well as 5m-SI can be found in the Supporting Information. Products of unselective borylations are mentioned as well for reasons of completeness (5n-SI to 5q-SI).
  19. See the SI for details.
  20. The application of 3,6-dichloro-4,5-dimethylpyridazine (SI-1) instead of 1,4-dichlorophthalazine as diazine component was also successful, as were subsequent borylation reactions to compound SI-4 (see SI, paragraph 4.3). However, in this particular case, we were not able to separate unreacted menthol from the products, even after different attempts of purification. From these experiments, we draw the conclusion to use just a minimal excess of the menthyl component as described in Scheme 1.
  21. The presence of a menthyl group is described to be without effect in the Cu-catalyzed coupling of phthalazine N-oxides and 1-diazonaphthoquinones:Biswas A.; Pan S.; Samanta R. Cu(II)-Catalyzed Construction of Heterobiaryls using 1-Diazonaphthoquinones: A General Strategy for the Synthesis of QUINOX and Related P. N Ligands. Org. Lett. 2022, 24, 1631–1636. 10.1021/acs.orglett.2c00127. [DOI] [PubMed] [Google Scholar]
  22. CCDC 2224469 contains the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service www.ccdc.cam.ac.uk/structures.
  23. The 13C NMR spectra of the products 4 and 5 in cases appeared to show a doubling of the signals of the menthyl or boryl groups, which after additional NMR experiments can be interpreted as conformers (or rotamers) of the menthyloxy group by rotation around the C–O bond and different orientations in the heterobiaryl products.
  24. Initial cross-coupling attempts with iodobenzene and 4-bromoanisole using palladium-based catalyst systems gave the expected coupling products; however, epimerization of the atropisomers took place to a significant extent (Table S3, SI).
  25. Guo X.; Dang H.; Wisniewski S. R.; Simmons E. M. Nickel-Catalyzed Suzuki–Miyaura Cross-Coupling Facilitated by a Weak Amine Base with Water as a Cosolvent. Organometallics 2022, 41, 1269–1274. 10.1021/acs.organomet.2c00197. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

The data underlying this study are available in the published article and its Supporting Information.


Articles from The Journal of Organic Chemistry are provided here courtesy of American Chemical Society

RESOURCES