Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Synthesis and stereochemistry of chiral aza-boraspirobifluorenes with tetrahedral boron-stereogenic centers

Yusuke Yoshigoe *, Keiichiro Hashizume and Shinichi Saito *
Department of Chemistry, Faculty of Science, Tokyo University of Science, Shinjuku, Tokyo 162-8601, Japan. E-mail: ssaito@rs.tus.ac.jp; yoshigoe.yusuke@rs.tus.ac.jp

Received 12th October 2022 , Accepted 14th October 2022

First published on 28th October 2022


Abstract

We have synthesized chiral aza-boraspirobifluorenes and evaluated their structural and photophysical properties. Enantiomers were separated by chiral HPLC on a semi-preparative scale, and the absolute stereochemistry was determined by comparison of experimental circular dichroism (CD) spectra and theoretical electronic CD (ECD) spectra. A kinetic analysis combined with theoretical calculations revealed that the rate-determining step of the racemization involves the cleavage of the B–N bond.


Introduction

Chiral tetrahedral boron compounds are promising optical materials1–4 that exhibit circular dichroism (CD) or circular polarized luminescence (CPL), and that have been applied for stereoselective transformations.5–12 A classic tetrahedral borane with four B–C bonds (BC4) was reported by Torssell in 1962 (Fig. 1a).13 Due to the high stability of the B–C bonds, BC4 boranes exhibit high stereochemical stability.14 To date, chiral tetrahedral boranes with B–N,15–27 B–O,28–31 B–H,32,33 and boron–heteroatom34–37 bonds have been synthesized and the stereochemistry of the compounds has been studied extensively. For example, Toyota has reported stable chiral tetrahedral boranes with three B–C bonds and one B–N bond (BC3N), and examined the racemization process (Fig. 1b).38 Here, the cleavage of the B–N bond is the rate-determining step. Recently, He has reported the enantioselective synthesis of BC3N and BC2N2 boranes via asymmetric copper-catalyzed azide–alkyne cycloaddition (CuAAc) reactions and studied their stereochemical stability.39
image file: d2dt03303h-f1.tif
Fig. 1 (a) Chiral tetrahedral boranes with four B–C bonds (BC4) reported by Torssel et al.; (b) chiral tetrahedral boranes with three B–C bonds and one B–N bond (BC3N); (c) an achiral spiro BC3N borane (aza-boraspirobifluorene); (d) this work: chiral spiro BC3N boranes.

Aza-boraspirobifluorenes, synthesized first by Ingleson et al. in 2015,40 are neutral tetrahedral boranes with a BC3N framework, in which the boron atom links an orthogonal π-conjugated pair of a borafluorene and an aza-borafluorene (Fig. 1c). Fukagawa et al. have applied these chemicals to organic fluorescent materials, i.e., OLEDs.41–43 We envisioned that chiral aza-boraspirobifluorenes, whose optical resolution and stereochemical stability have not yet been reported,44 could be synthesized from substituted borafluorenes (Fig. 1d). In the present study, we synthesized chiral aza-boraspirobifluorenes and evaluated their stereochemical stability. The mechanism of the racemization process is discussed based on the obtained kinetic data. The photophysical properties, including circular dichroism (CD) spectra of the homochiral compounds, were also studied.

Results and discussion

We synthesized chiral aza-boraspirobifluorene analogue 1 by two methods (Tables 1 and 2). Compound 1a was obtained in 35% yield by the reaction of 2-(2-(dibromoboraneyl)phenyl)pyridine (2)45 with an excess (4.0 equiv.) of Grignard reagent 3a (Table 1, entry 1). Other substituted aza-boraspirobifluorenes (1b–e) were obtained from the reaction of 2 with the corresponding Grignard reagents (3b–e) in 18–40% yield (entries 2–6).46 The methoxy-substituted compounds 1b and 1e were obtained in higher yield than the other products (entries 2 and 5).
Table 1 Synthesis of aza-boraspirobifluorenes 1a–f

image file: d2dt03303h-u1.tif

Entry Reagent R1 R2 1 Yield (%)
a The Grignard reagent was prepared by the reaction of the dibromobiphenyl derivatives, I2-activated Mg, DIBAL-H, and LiCl.
1 3a H H 1a 35
2 3b H H 1b 40
3 3c[thin space (1/6-em)]a H H 1c 20
4 3d Me H 1d 18
5 3e OMe H 1e 40
6 3f[thin space (1/6-em)]a CF3 H 1f 15


Table 2 Synthesis of aza-boraspirobifluorenes 1g–i

image file: d2dt03303h-u2.tif

Entry Reagent R3 1 Yield (%)a
a Yield over two steps based on compound 4.
1 6 Me 1g 24
2 7 OMe 1h 43
3 8 CF3 1i 33


Aza-boraspirobifluorenes with substituted pyridine rings were obtained via a different route (Table 2). Treatment of compound 4 with nBuLi (2.0 equiv.), followed by the addition of a solution of BCl3 (1.0 equiv.), furnished chloroborafluorene 5.47,48 A solution of 5 was added to aryl lithium 6, and 1g was isolated in 24% yield (entry 1). The 4-methoxy derivative (1h) and 4-trifluoromethyl derivative (1i) were obtained from the corresponding aryl lithium in 43% and 33% yield, respectively (entries 2 and 3).

The solid-state structures of 1a–i were examined by X-ray crystallography. Single crystals of 1a–i were obtained from their solutions by slow evaporation of the solvent. The thermal-ellipsoid plot of rac-1a is shown in Fig. 2. These crystals contained (R)-1a and (S)-1a in a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1 within an achiral crystal system (P21/c). The tetrahedral BC3N framework with an intramolecular B–N bond was confirmed. The torsion between the planes of the 2-phenylpyridine and biphenyl moieties was nearly orthogonal (θ = 86.4°). The B–N bond was slightly shorter (1.615(2) Å) than those of other borane–amine complexes (1.67–1.76 Å),49,50 and close to the length of the B–C bonds (1.60–1.62 Å).51


image file: d2dt03303h-f2.tif
Fig. 2 Molecular structure of 1a with thermal ellipsoids at 50% probability; hydrogen atoms are omitted for clarity.

Selected bond lengths and angles around the boron atom in 1a–f are compiled in Table 3. There is no clear relationship between the substituent introduced to the aza-boraspirobifluorene and the length of the B–N bond: the bond is slightly longer in 1a (1.615(2) Å) and slightly shorter in 1h (1.588(7) Å), while the bond length of other compounds remained constant (1.603(3)–1.61 0(10) Å). The tetrahedral character of the dihedral angle (THCDA)49,52 for the spiro atoms of 1a–i was ∼50%. These values are affected by the substituents on the pyridine ring (R3) and decrease from 62% to 29% in the order 1h (OMe) > 1g (Me) > 1i (CF3). The very small value for 1i can be explained by the reduced Lewis basicity of the nitrogen atom upon introducing an electron-withdrawing group, which leads to an increased three-coordinated character of the boron atom. The bond lengths of B–C1 (1.690(20) Å) and B–C2 (1.700(10) Å) in 1i are slightly longer than those in the other compounds.

Table 3 Selected bond lengths and bond angles of 1a–i[thin space (1/6-em)]a
  Bond length/Å Bond angle (°) THCDA (%)[thin space (1/6-em)]b
B–N B–C1 B–C2 B–C3 C1BC2 C3BN
a See Fig. 2 for assignment of the bond lengths and bond angles. b The tetrahedral character of the dihedral angle (THCDA) was calculated based on six angles around the boron atom using the equation in ref. 22.
1a 1.615(2) 1.615(2) 1.619(2) 1.607(2) 99.8(1) 96.9(1) 50
1b 1.605(3) 1.608(3) 1.624(3) 1.607(3) 99.6(2) 96.7(2) 49
1c 1.603(4) 1.623(4) 1.616(3) 1.600(3) 99.6(2) 97.1(2) 49
1d 1.604(2) 1.611(3) 1.624(3) 1.609(3) 99.8(2) 97.3(2) 51
1e 1.604(2) 1.617(3) 1.622(2) 1.614(3) 99.8(1) 97.3(1) 51
1f 1.603(3) 1.618(2) 1.623(3) 1.605(2) 99.7(1) 97.4(1) 51
1g 1.610(10) 1.610(10) 1.620(10) 1.620(10) 101.1(6) 97.6(6) 54
1h 1.588(7) 1.650(10) 1.593(9) 1.610(10) 100.6(5) 98.4(5) 62
1i 1.610(2) 1.690(20) 1.700(10) 1.608(2) 91.1(8) 97.0(1) 29


UV/vis absorption and photoluminescence spectroscopy in CH2Cl2 were examined to understand the optical properties of 1a–f. Fig. 3 shows the UV/vis absorption and fluorescence spectrum of 1d. We observed an absorption band around 260–340 nm with a peak (λmax, abs.) at 266 nm and a broad fluorescence band around 300–440 nm with two peaks (λmax, em.) at 338 nm and 386 nm. Fluorescence spectra of other aza-boraspirobifluorenes composed of simple substituents emitted violet-blue fluorescence with λmax, em. in the range of 334–463 nm (for details, see the ESI).


image file: d2dt03303h-f3.tif
Fig. 3 Absorption (dashed) and fluorescence (solid) spectrum of 1d (5 μM in CH2Cl2).

Next, we examined the separation of the enantiomers and the stereochemical stability of aza-boraspirobifluorenes 1a–i (Fig. 4a). Using chiral HPLC, rac-1d was separated into two fractions (fr1-1d: ee > 99%; fr2-1d: ee > 99%). The CD spectra of fr1-1d and fr2-1d display mirrored Cotton effects (Fig. 4b). The spectrum for fr1-1d displayed a positive Cotton effect at 275 nm and a negative Cotton effect at 323 nm. The observed spectrum of fr-1d was in good agreement (within 10 nm) with the calculated ECD spectrum of (S)-1d (Fig. 4b).53–55 Based on these results, we assigned the absolute configuration of fr1-1d as (S)-1d and that of fr2-1d as (R)-1d.


image file: d2dt03303h-f4.tif
Fig. 4 (a) HPLC chart of 1d; column: YMC CHIRAL Amylose-SA (L × D: 250 mm × 4.6 mm; particle size: 5 μm); detector: photo diode array (PDA) detector at 254 nm; eluent: n-hexane : CHCl3 : i-PrOH = 93[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]5; elution rate: 0.8 mL min−1; (b) experimental CD spectra (solid) and theoretical ECD spectra (dashed) calculated at the M06-2X/6-31+G(d,p) level, whereby the structures of 1d were optimized at the B3LYP/6-31G(d,p) level: (S)-1d (red) and (R)-1d (blue).

With the separated homochiral compounds in hand, we studied the racemization of 1 under various conditions. Heating a solution of (S)-1d in DMSO to 145 °C induced racemization.56,57 The rate constant of racemization (krac) of (S)-1d at 145 °C (4.00 × 10−5 s−1) was determined by a first-order kinetic analysis.58–60 The kinetic parameters (145 °C: ΔG = 33.7 kcal mol−1, ΔH = 26.8 kcal mol−1, and ΔS = −16.6 cal K−1 mol−1), which were determined by an Eyring-Polanyi plot at three different temperatures (135–145 °C), were higher than those of other reported BC3N systems.39,40

The kinetic parameters for racemization of 1a–i are summarized in Table 4. For 1a–c, which carry a substituent at the 2-position of the borafluorene ring (R1), ΔG increases from 32.9 kcal mol−1 to 34.9 kcal mol−1 (145 °C) in the order 1b (R1 = OMe) < 1a (R1 = Me) < 1c (R1 = CF3). A similar trend was observed for ΔG of 1d–f, which increases in the order 1e (R2 = OMe) ≃ 1d (R2 = Me) < 1f (R2 = CF3). At present, we assume that the electron-withdrawing group increases the Lewis acidity of the boron atom and that the presence of a stronger B–N bond retards the racemization. For 1g–i, which carry a substituent at the 4-positon of the aza-borafluorene ring (R3), a clear substituent effect was observed. ΔG decreases from 33.4 to 31.0 kcal mol−1 in the order 1g (R3 = OMe) > 1h (R3 = Me) > 1i (R3 = CF3). This result can be rationalized in terms of the reduced basicity of the nitrogen atom upon introducing the electron-withdrawing group. In accordance with these results, 1i, which contains a MeO group at the 2-position of the borafluorene ring and a CF3 group at the 4-position of the aza-borafluorene ring, showed the lowest activation energy (ΔG = 31.0 kcal mol−1). Based on the kinetic data obtained, we concluded that the rate-determining step of the racemization involves the cleavage of the B–N bond.

Table 4 Summary of the kinetic parameters for the racemization of 1a–i

image file: d2dt03303h-u3.tif

  R1 R2 R3 k rac (145 °C) 10−5/s−1 ΔG (145 °C)/kcal mol−1 ΔH/kcal mol−1
1a Me H H 5.76 33.4 28.4
1b OMe H H 11.3 32.9 29.6
1c CF3 H H 0.984 34.9 33.9
1d H Me H 4.00 33.7 26.8
1e H OMe H 4.04 33.7 31.6
1f H CF3 H 1.84 34.4 27.1
1g Me H Me 2.70 33.4 28.4
1h Me H OMe 1.77 34.4 30.7
1i Me H CF3 114 31.0 23.5


The mechanism of racemization was further examined by theoretical calculations. The characteristic structures and the energy levels for the inversion process of (S)-1d, which were calculated at the B3LYP/6-311+G(d,p)//B3LYP/6-31G(d) level, are shown in Fig. 5. The cleavage of the B–N bond and the rotation of the B–C3 bond can proceed viaInt. The process from GS-(S) to IntviaTS-1 is favored with an activation energy of ΔG = 36.8 kcal mol−1. The length of the B–N bond increases from 1.62 Å (GS) to 4.11 Å (TS-1). The rotational intermediate Int contains the 5-phenylborafluorene moiety, which carries a three-coordinated boron atom with no B–N interaction. Another transition state (TS-2) was obtained for the formation of the B–N bond.


image file: d2dt03303h-f5.tif
Fig. 5 Energy diagram for the chirality inversion from (S)-1d to (R)-1d; all calculations were carried out at the B3LYP/6-311+G(d,p)//B3LYP/6-31G(d) level.

Conclusions

We synthesized chiral aza-boraspirobifluorenes and examined their molecular structures as well as their photophysical properties. The enantiomers were separated via chiral HPLC. The racemization of aza-boraspirobifluorenes proceeded at elevated temperatures, and the kinetic parameters were determined to understand the mechanism of the isomerization. This study thus offers important insights into the stereochemistry of chiral tetrahedral borane compounds.

Conflicts of interest

There are no conflicts to declare.

References

  1. D. Li, H. Zhang and Y. Wang, Chem. Soc. Rev., 2013, 42, 8416–8433 RSC.
  2. D. Frath, J. Massue, G. Ulrich and R. Ziessel, Angew. Chem., Int. Ed., 2014, 53, 2290–2310 CrossRef CAS PubMed.
  3. S. K. Mellerup and S. Wang, Chem. Soc. Rev., 2019, 48, 3537–3549 RSC.
  4. X. Chen, D. Tan and D.-T. Yang, J. Mater. Chem. C, 2022, 10, 13499–13532 RSC.
  5. D. G. Hall, Boronic Acid, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2011 Search PubMed.
  6. M. G. Davidson, K. Wade, T. B. Marder and A. K. Hughes, Contemporary Boron Chemistry, Royal Society of Chemistry, Cambridge, 2000 Search PubMed.
  7. J. W. B. Fyfe and A. J. B. Watson, Chem, 2017, 3, 31–55 CAS.
  8. C.-Y. Wang, J. Derosa and M. R. Biscoe, Chem. Sci., 2015, 6, 5105–5113 RSC.
  9. D. Leonori and V. K. Aggarwal, Angew. Chem., Int. Ed., 2015, 54, 1082–1096 CrossRef CAS PubMed.
  10. C. Sandford and V. K. Aggarwal, Chem. Commun., 2017, 53, 5481–5794 RSC.
  11. H. Wang, C. Jing, A. Noble and V. K. Aggarwal, Angew. Chem., Int. Ed., 2020, 59, 16859–16872 CrossRef CAS PubMed.
  12. A. J. J. Lennox and G. C. Lloyd-Jones, Chem. Soc. Rev., 2014, 43, 421–443 RSC.
  13. K. Torssell, Acta Chem. Scand., 1962, 16, 87–93 CrossRef CAS.
  14. D. J. Owen, D. Vanderveer and G. B. Schuster, J. Am. Chem. Soc., 1998, 120, 1705–1717 CrossRef CAS.
  15. L. W.-Y. Wong, A. S.-F. Au Yeung, G. S.-S. Tam, J. W.-H. Kan, H. H.-Y. Sung, F. K. Sheong, Z. Lin and I. D. Williams, RSC Adv., 2018, 8, 1451–1460 RSC.
  16. L. W.-Y. Wong, J. W.-H. Kan, T.-H. Nguyen, H. H.-Y. Sung, D. Li, A. S.-F. Au-Yeung, R. Sharma, Z. Lin and I. D. Williams, Chem. Commun., 2015, 51, 15760–15763 RSC.
  17. D. Uraguchi, F. Ueoka, N. Tanaka, T. Kizu, W. Takahashi and T. Ooi, Angew. Chem., Int. Ed., 2020, 59, 11456–11461 CrossRef CAS PubMed.
  18. J. Huskens and M. Reetz, Eur. J. Org. Chem., 1999, 1775–1786 CrossRef CAS.
  19. D. Liu, Z. Shan, Y. Zhou, X. Wu and J. Qin, Helv. Chim. Acta, 2004, 87, 2310–2317 CrossRef CAS.
  20. Z. He, L. Liu, Z. Zhao, S. K. Mellerup, Y. Ge, X. Wang, N. Wang and S. Wang, Chem. – Eur. J., 2020, 26, 12403–12410 CrossRef CAS PubMed.
  21. B. Deng, X. Wang and S. Wang, Chem. – Eur. J., 2019, 25, 14694–14700 CrossRef CAS PubMed.
  22. P. Novoseltseva, H. Li, X. Wang, F. Sauriol and S. Wang, Chem. – Eur. J., 2020, 26, 2276–2284 CrossRef CAS PubMed.
  23. M. Grandl, B. Rudolf, Y. Sun, D. F. Bechtel, A. J. Pierik and F. Pammer, Organometallics, 2017, 36, 2527–2535 CrossRef CAS.
  24. T. A. Bartholome, J. J. Martinez, A. Kaur, D. J. D. Wilson, J. L. Dutton and C. D. Martin, Organometallics, 2021, 40, 1966–1973 CrossRef CAS.
  25. M. Grandl, Y. Sun and F. Pammer, Org. Chem. Front., 2018, 5, 336–352 RSC.
  26. F. Pammer, J. Schepper, J. Glöckler, Y. Sun and A. Orthaber, Dalton Trans., 2019, 48, 10298–10312 RSC.
  27. M. Grandl, Y. Sun and F. Pammer, Chem. – Eur. J., 2016, 22, 3976–3980 CrossRef CAS PubMed.
  28. M. Braun, S. Schlecht, M. Engelmann, W. Frank and S. Grimme, Eur. J. Org. Chem., 2008, 5221–5225 CrossRef CAS.
  29. P. F. Kaiser, J. M. White and C. A. Hutton, J. Am. Chem. Soc., 2008, 130, 16450–16451 CrossRef CAS PubMed.
  30. V. G. Jiménez, F. M. F. Santos, S. Castro-Fernández, J. M. Cuerva, P. M. P. Gois, U. Pischel and A. G. Campaña, J. Org. Chem., 2018, 83, 14057–14062 CrossRef PubMed.
  31. S. Schlecht, W. Frank and M. Braun, Beilstein J. Org. Chem., 2011, 7, 615–621 CrossRef CAS PubMed.
  32. L. Charoy, A. Valleix, T. Le Gall, P. Pham Van Chuong, C. Mioskowski and L. Toupet, Chem. Commun., 2000, 2275–2276 RSC.
  33. C. Aupic, A. Abdou Mohamed, C. Figliola, P. Nava, B. Tuccio, G. Chouraqui, J.-L. Parrain and O. Chuzel, Chem. Sci., 2019, 10, 6524–6530 RSC.
  34. L.-Y. Wang, Z.-F. Liu, K.-X. Teng, L.-Y. Niu and Q.-Z. Yang, Chem. Commun., 2022, 58, 3807–3810 RSC.
  35. E. Vedejs, S. C. Fields, R. Hayashi, S. R. Hitchcock, D. R. Powell and M. R. Schrimpf, J. Am. Chem. Soc., 1999, 121, 2460–2470 CrossRef CAS.
  36. E. Vedejs, S. C. Fields, S. Lin and M. R. Schrimpf, J. Org. Chem., 1995, 60, 3028–3034 CrossRef CAS.
  37. T. Imamoto and H. Morishita, J. Am. Chem. Soc., 2000, 122, 6329–6330 CrossRef CAS.
  38. S. Toyota, F. Ito, N. Nitta and T. Hakamata, Bull. Chem. Soc. Jpn., 2004, 77, 2081–2088 CrossRef CAS.
  39. B. Zu, Y. Guo and C. He, J. Am. Chem. Soc., 2021, 143, 16302–16310 CrossRef CAS PubMed.
  40. D. L. Crossley, J. Cid, L. D. Curless, M. L. Turner and M. J. Ingleson, Organometallics, 2015, 34(24), 5767–5774 CrossRef CAS PubMed.
  41. H. Fukagawa, T. Sasaki, T. Tsuzuki, Y. Nakajima, T. Takei, G. Motomura, M. Hasegawa, K. Morii and T. Shimizu, Adv. Mater., 2018, 30, 1706768 CrossRef PubMed.
  42. H. Fukagawa, M. Hasegawa, K. Morii, K. Suzuki, T. Sasaki and T. Shimizu, Adv. Mater., 2019, 31, 1904201 CrossRef CAS PubMed.
  43. T. Sasaki, M. Hasegawa, K. Inagaki, H. Ito, K. Suzuki, T. Oono, K. Morii, T. Shimizu and H. Fukagawa, Nat. Commun., 2021, 12, 2706 CrossRef CAS PubMed.
  44. For the synthesis and chiral resolution of structurally related 9,9′-spirobifluorenes, see: C. Stobe, R. Seto, A. Schneider and A. Lützen, Eur. J. Org. Chem., 2014, 6513–6518 CrossRef CAS.
  45. R. Wada, S. Kaga, Y. Kawai, K. Futamura, T. Murai and F. Shibahara, Tetrahedron, 2021, 83, 131978 CrossRef CAS.
  46. Our attempts for improving the yield of 1d were unsuccessful; for details, see the ESI..
  47. X. Mu, H. Zhang, P. Chen and G. Liu, Chem. Sci., 2014, 5, 275–280 RSC.
  48. The synthesis of 5–7 was performed with reference to the synthesis of 5-chloro-5H-dibenzo[b,d]borole; for details, see: Y. Shoji, N. Tanaka, S. Muranaka, N. Shigeno, H. Sugiyama, K. Takenouchi, F. Hajjaj and T. Fukushima, Nat. Commun., 2016, 7, 12704 CrossRef PubMed.
  49. H. Höpfl, J. Organomet. Chem., 1999, 581, 129–149 CrossRef.
  50. S. Toyota and M. Oki, Bull. Chem. Soc. Jpn., 1991, 64, 1554–1562 CrossRef CAS.
  51. F. H. Allen, O. Kennard, D. G. Watson, L. Brammer, A. G. Orpen and R. Taylor, J. Chem. Soc., Perkin Trans. 2, 1987, S1 RSC.
  52. The tetrahedral character of the dihedral angle (THCDA) can be calculated using the Höpfl equation: image file: d2dt03303h-t1.tif.
  53. H. Takekawa, K. Tanaka, E. Fukushi, K. Matsuo, T. Nehira and M. Hashimoto, J. Nat. Prod., 2013, 76, 1047–1051 CrossRef CAS PubMed.
  54. T. Ito and T. Nehira, Tetrahedron Lett., 2014, 55, 314–318 CrossRef CAS.
  55. D. J. Tantillo, Nat. Prod. Rep., 2013, 30, 1079–1086 RSC.
  56. When a solution of the isolated (S)-1d in DMSO was kept at 25 °C without irradiation, racemization was not observed.
  57. Racemization was not observed when a solution of (S)-1d in DMSO was irradiated with UV light (λ = 365 nm), blue LED light (λ = 460 nm), or a common domestic lamp at room temperature.
  58. C. H. Bamford and C. F. H. Tipper, Comprehensive Chemical Kinetics, vol. 1. The practice of kinetics, Elsevier B.V., Amsterdam, 1969 Search PubMed.
  59. M. F. Boehm and J. L. Bada, Proc. Natl. Acad. Sci. U. S. A., 1984, 81, 5263–5266 CrossRef CAS PubMed.
  60. L. Canoira, M.-J. García-Martínez, J. F. Llamas, J. E. Ortíz and T. D. Torres, Int. J. Chem. Kinet., 2003, 35, 576–591 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. CCDC 2203027–2203033, 2176259 and 2215757. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d2dt03303h

This journal is © The Royal Society of Chemistry 2022
Click here to see how this site uses Cookies. View our privacy policy here.