Next Article in Journal
Morphological and Structural Properties of Amino-Functionalized Fumed Nanosilica and Its Comparison with Nanoparticles Obtained by Modified Stöber Method
Next Article in Special Issue
Magnetic Nanoparticles Fishing for Biomarkers in Artificial Saliva
Previous Article in Journal
The Role of State-of-the-Art Quantum-Chemical Calculations in Astrochemistry: Formation Route and Spectroscopy of Ethanimine as a Paradigmatic Case
Previous Article in Special Issue
m-Carborane as a Novel Core for Periphery-Decorated Macromolecules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reactions of Dihaloboranes with Electron-Rich 1,4-Bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadienes

Department of Chemistry, Southern University of Science and Technology, Shenzhen 518055, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2020, 25(12), 2875; https://doi.org/10.3390/molecules25122875
Submission received: 31 May 2020 / Revised: 14 June 2020 / Accepted: 16 June 2020 / Published: 22 June 2020

Abstract

:
The reactions of electron-rich organosilicon compounds 1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene (1), 2,3,5,6-tetramethyl-1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene (2), and 1,1′-bis(trimethylsilyl)-1,1′-dihydro-4,4′-bipyridine (12) with B-amino and B-aryl dihaloboranes afforded a series of novel B=N-bond-containing compounds 311 and 13. The B=N rotational barriers of 7 (>71.56 kJ/mol), 10 (58.79 kJ/mol), and 13 (58.65 kJ/mol) were determined by variable-temperature 1H-NMR spectroscopy, thus reflecting different degrees of B=N double bond character in the corresponding compounds. In addition, ring external olefin isomers 11 were obtained by a reaction between 2 and DurBBr2. All obtained B=N-containing products were characterized by multinuclear NMR spectroscopy. Compounds 5, 9, 10a, 11, and 13a were also characterized by single-crystal X-ray diffraction analysis.

1. Introduction

Low-valent boron compounds are a class of highly reactive species that have been the focus of intense research because of their unique electronic properties [1,2] as well as their diverse and fascinating reactivity patterns such as inert bond activation [3,4], cycloaddition reaction [5,6,7], and small molecule activation [8]. The progress in this research area is highlighted by the very recent results in terms of borylene-mediated N2 activation [9] and N2 coupling [10]. Nonetheless, the synthetic approach to low-valent boron species is severely limited [11,12,13]. Almost all of the reported synthetic strategies require a strong metallic reducing agent (e.g., Li, K, Na, KC8) [3,4,14,15,16,17,18,19,20,21,22,23], harsh reaction conditions, and a strict moisture- and oxygen-free atmosphere. Therefore, the exploration of metal-free reductants to access low-valent boron species is highly desirable [24,25,26].
Mashima et al. reported a class of electron-rich organosilicon compounds 1, 2, and 12, which can serve as versatile reducing reagents for the group 4–6 metal chloride complexes. The corresponding low-valent metal species were prepared in a salt-free manner [27,28,29,30,31,32,33]. The reducing power mainly derives from the aromatization of the central 1,4-diaza-2,5-cyclohexadiene ring. Deeply inspired by the advantage of the salt-free reduction protocol and easy workup, we decided to examine the ability of the organosilicon compounds 1, 2, and 12 for the reduction of trivalent dihaloboranes. Based on the published results, the disubstituted compounds ArXB(N2C4R4)BXAr are proposed as the reduction products. We hypothesized two possible bonding modes (i.e., A and B in Scheme 1) between the C4N2 ring and the boron atoms. In the first manner, B–N is bound by an electron-precise σ bond (A, Scheme 1) and an additional N―B dative π bond. In the second manner, two nitrogen atoms each provide a π-electron for 6π-aromatization, while the remaining two valence electrons form a lone pair on each N atom, donating to the empty sp2-hybridized orbital of boron, thus leading to the divalent boron radical centers (B, Scheme 1).

2. Results and Discussion

First, we examined compound 1 for its ability to reduce ArBX2. The results are summarized in Scheme 2. Compound 1 [34] and ArBX2 (Ar = 2,3,5,6-tetramethylphenyl (Dur), 2,4,6-trimethylphenyl (Mes)) [35] were prepared according to the literature. The reaction of 1 with an equimolar amount of DurBBr2 and MesBCl2 at ambient temperature afforded the expected monosubstituted products 3 and 4, respectively. Adding the second equiv. of dihaloboranes to the reaction mixture led to the disubstituted products 5 and 6. In stark contrast, the reaction of 1 with an equimolar amount of the less sterically demanding PhBCl2 caused precipitation, which is insoluble in all ordinary solvents. This is most likely due to the polymerization of PhClBC4N2H4SiMe3 by chlorosilane elimination. Hence, the stepwise synthetic protocol is unsuitable for the synthesis of 7. Instead, 1 was directly treated with 2 equiv. of PhBCl2 at room temperature (RT), affording 7 in an acceptable yield (48%). Hence, the reaction of the monosubstituted intermediate (i.e., PhClBC4N2H4SiMe3) with PhBCl2 should be much faster than the self-polymerization process. Compounds 37 were confirmed by NMR spectroscopic (Figures S1–S15) and HRMS studies. Furthermore, the multinuclear NMR spectroscopic study revealed that the isolated 57 all consist of ca. 1:1 cis-trans isomers in the solution phase at ambient temperature due to the nonrotatable B=N double bond (see Electronic Supporting Information (ESI)).
Suitable single crystals of 5 for X-ray diffraction analysis were obtained by slow evaporation of a saturated hexane solution. Two isomers, 5a and 5b, co-crystallized in the unit cell. The result is depicted in Figure 1 and Figure S37. The central C4N2 ring is nearly planar. The endocyclic N1–C2 (1.385(7) Å), N2–C3 (1.415(7) Å), C1–C2 (1.311(9) Å), and C3–C3* (1.326(8) Å) distances lie in the expected range for N–C single bonds and C=C double bonds. The bond lengths of B1–N1 (1.423(8) Å) and B2–N2 (1.400(8) Å) are shorter than that of a B–N single bond, which is indicative of a significant B=N double bond character. All these geometric parameters suggest the bonding mode A in Scheme 1. Therefore, both boron centers adopt a formal oxidation state of +3.
Compound 2, which features a less-negative redox potential (+0.10 V) with respect to 1 (−0.24 V), was further examined to reduce PhBCl2 and DurBBr2 (Scheme 2). Differing from the aforementioned reactions with 1, both monosubstituted products 8 and 9 could be prepared upon a 1:1 ratio reaction of 2 with PhBCl2 and DurBBr2, respectively. Upon the reaction of 2 with two equiv. of PhBCl2 at RT, the disubstituted compounds 10a and 10b were obtained as 1:1 cis-trans isomers. Surprisingly, treatment of 2 with two equiv. of DurBBr2 at ambient temperature led to the formation of 11, which can be regarded as the product from an isomerization of 11′ [36,37]. Compounds 811 were confirmed by NMR spectroscopic (Figures S16–S28) and HRMS studies. There were two sets (intensity ratio of ca. 1:0.3) of 1H signals between 4 and 6 ppm, each consisting of three multiplets with the integration ratio of 1:1:1, which can be assigned to the migrated H and two remaining olefinic protons. These are the most characteristic signs for the formation of the isomerized product. After assigning each peak (with the help of the NOE spectrum, see the ESI Figure S26 for more details), we could determine that the ratio of isomers 11a and 11b was 77:23. Furthermore, the isomerization was also observed upon the treatment of the isolated 9 with an equimolar amount of DurBBr2 at RT.
The structures of 9, 10a, 11a, and 11b were confirmed by single-crystal X-ray diffraction analysis (Figure 2 and Figures S38–S40). All four compounds adopt a boat conformation, which could be explained by the small energy difference between the planar and nonplanar geometry of the C4N2 ring, and the steric congestion between the central exocyclic methyl groups and the bulky boron substituents. Bond lengths (Å) of 9 (B1–N1 1.376(7), N1–C1 1.451(6), C1–C2 1.343(7), N2–C2 1.434(6), N2–Si1 1.759(5)), 10 (B1–N1 1.401(2), N1–C2 1.4494(17), C1–C2 1.328(2), N1–C1* 1.4513(18)) are all as expected. The overall structures of 11a and 11b resemble that of 10a. However, since the C3 position in 11a and 11b accepted one H atom from the methyl group at the C2 position, respectively, and thus became sp3-hybridized, the torsion angles C4–C3–N2–B2 (11a: 94.25°; 11b: 94.69°) are notably greater than those at the other three carbon positions (57–63°) in the central six-membered ring. Due to the disordered nature of the crystal, the bond lengths of 11a and 11b cannot be further discussed.
The reaction of (SiMe3)2NBCl2 with 12 [38] of greater reducing power (redox potential of −0.40 V) [39] was performed at ambient temperature in C6D6. After the removal of the solvent and extraction with hexane, an NMR spectroscopically pure product 13 was obtained with a yield of 75%. Compound 13 was confirmed by NMR spectroscopic (Figures S29–S31) and HRMS studies. Suitable single crystals of 13a for X-ray diffraction analysis were obtained upon storage of the reaction mixture overnight at RT (Figure 3 and Figure S41). The N1–C1/N1–C5 (1.402(8)–1.414(8) Å), C1–C2/C5–C4 (1.332(8)–1.347(9) Å), C2–C3/C3–C4 (1.444(9)–1.452(9) Å), C3–C3* (1.376(12) Å) distances are in line with the Lewis structure depicted in Scheme 3.
Apparently, both the RT-NMR spectroscopic and crystallographic studies failed to prove any successful reduction of the trivalent borane to divalent boron radical. Since the rotational barrier around an N―B dative bond should be lower than that of a B=N double bond, we assumed that any contribution from the bonding mode B (Scheme 1) should slightly lower the rotational barrier around the exocyclic B–N bond. In this context, we conducted a variable-temperature 1H-NMR experiment to provide further insight. Toluene-d8 was selected as the solvent with a temperature ranging from −60 °C to 80 °C. In general, the exocyclic H or Me as marked in Figure 4 (top right) should display two signals if the B=N bond is nonrotatable. The separated signals will coalesce at an elevated temperature when the B–N bond overcomes the rotational barrier and begins to rotate. Determination of the separation (Hz) of two signals and the coalescent temperature allows calculation of the B–N rotational barrier. The results of the VT-NMR experiments and assignment of the signals of interest are depicted in Figure 4 and Figures S32–S36. The obtained ΔG values are summarized in Table 1. Analysis of the VT-NMR spectra revealed 7 with a strong B=N bond, and 10 and 13 with weak B=N bonds, as reflected by their rotational barriers >71.56 kJ/mol (7), 58.79 kJ/mol (10), 58.65 kJ/mol (13) when compared with ordinary B=N double bonds (71–100 kJ/mol) [40]. When taking the aforementioned assumption into account, the remarkably lower B–N rotational barrier of 10 compared to that of 7 is not in line with the fact that the reducing power of 2 is slightly weaker than that of 1, according to the CV data. Therefore, the lower B–N rotational barrier in 10 should be mainly due to its boat conformation, which allows for less steric hindrance. Furthermore, although the central C4N2 rings of 7 and 13 both adopt a planar structure, the B–N rotational barrier in 13 is significantly lower than that of 7. This finding could be explained by the competition in π donation from another B-amino function (–N(SiMe3)2) in 13.

3. Materials and Methods

3.1. General Information

All manipulations were performed under dry argon using standard Schlenk line or glovebox techniques. Solvents were purified by distillation from Na under dry argon. C6D6 was dried over an Na/K alloy and then degassed by freeze–pump–thaw cycles. PhBCl2 was purchased from Beijing MREDA Technologie Co., Ltd., without any special treatment before use. The NMR spectra were acquired on a Bruker AVANCE 400 (1H: 400 MHz, 13C{1H}: 101 MHz, 11B: 128 MHz) NMR spectrometer at 298 K. Variable-temperature NMR experiments were conducted on a Bruker AVANCE 400 NMR spectrometer (1H: 400 MHz, 213–353 K). Chemical shifts are given in ppm. 1H and 13C{1H} NMR spectra were referenced to an external tetramethylsilane (TMS) via the residual protons of the solvent (1H) or the solvent itself (13C{1H}). 11B NMR spectra were referenced to the external BF3·OEt2. High-resolution mass spectrometry (HMRS) was performed with a Thermo Fisher Scientific Q Exactive Mass Spectrometer (MS) system.

3.2. Synthesis of 3 and 4:

In the glove box, DurBBr2 (30.3 mg, 0.1 mmol, 1.0 equiv.) and 1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene (1) (22.6 mg, 0.1 mmol, 1.0 equiv.) were added into C6D6 (0.6 mL) in a J. Young NMR tube. The mixture was rested for 10 min prior to the removal of the volatiles under vacuum to get 3 as a pale yellow solid (23.6 mg, 72 mmol, 63%). Compound 4 was synthesized in a similar manner, with a yield of 64%.
3: 1H-NMR (400 MHz, C6D6): δ = 6.86 (s, 1H, H of Dur), 6.35 (d, J = 6.6 Hz, 1H, H of C4N2), 5.13 (d, J = 6.5 Hz, 1H, H of C4N2), 4.91 (d, J = 6.6 Hz, 1H, H of C4N2), 4.64 (d, J = 6.5 Hz, 1H, H of C4N2), 2.29 (s, 6H, Me of Dur), 2.06 (s, 6H, Me of Dur), −0.23 (s, 9H, Me of TMS). 13C{1H}-NMR (101 MHz, C6D6): δ = 134.5, 133.4, 132.0, 119.8, 118.7, 113.5, 112.8, 19.4, 18.4, −2.3 (9C, C of TMS). The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 34.1. HRMS: calc. for [M]+ C17H26BBrN2Si+ 376.11362; found: 376.11308.
4: 1H-NMR (400 MHz, C6D6): δ = 6.75 (s, 2H, H of Mes), 6.17 (d, J = 6.6 Hz, 1H, H of C4N2), 5.07 (d, J = 6.5 Hz, 1H, H of C4N2), 4.91 (d, J = 6.6 Hz, 1H, H of C4N2), 4.65 (d, J = 6.5 Hz, 1H, H of C4N2), 2.37 (s, 6H, Me of Mes), 2.15 (s, 3H, Me of Mes), 0.21 (s, 9H, Me of TMS). 13C{1H}-NMR (101 MHz, C6D6): δ = 139.2, 137.8, 127.5, 119.4, 118.5, 112.4 (1C, C of C4N2), 112.8 (1C, C of C4N2), 21.3 (2C, o-CH3 of Mes), 20.9 (C, p-CH3 of Mes), −2.3 (9C, C of TMS). The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 34.2. HRMS: calc. for [M]+ C16H24BClN2Si+ 318.14848; found: 318.14877.

3.3. Synthesis of 57

In the glove box, DurBBr2 (60.4 mg, 0.2 mmol, 2.0 equiv.) and 1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene (1) (22.6 mg, 0.1 mmol, 1.0 equiv.) were added into C6D6 (0.6 mL) in a J. Young NMR tube. The mixture was rested overnight prior to the removal of the volatiles under vacuum to get 5 as a yellow oil with a 45% yield. Mixture 5 contains the cis-structure 5a and trans-structure 5b, and the ratio of the cis-trans isomers was about 1:1. Compounds 67 were synthesized in a similar manner, with a cis-trans isomers ratio of about 1:1 (yield: 52% (6) and 48% (7)].
5a + 5b: 1H-NMR (400 MHz, C6D6): δ = 6.86 (s, 2H), 6.82 (s, 2H), 6.52 (s, 2H), 6.26 (d, 3JH-H = 1.6 Hz, 1H), 6.24 (d, 3JH-H =1.6 Hz, 1H), 5.36 (d, 3JH-H = 1.6 Hz, 1H), 5.34 (d, 3JH-H = 1.6 Hz, 1H), 4.99 (s, 2H), 2.09 (s, 12 H), 2.07 (s, 12 H), 2.03 (s, 12 H), 2.01 (s, 12H). 13C{1H}-NMR (101 MHz, C6D6): δ = 134.0, 133.9, 133.6, 133.5, 132.5, 132.3, 118.2, 117.7, 117.2, 116.6, 19.2, 19.1, 18.5, 18.4. The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 38.9. HRMS: calc. for [M]+ C24H30N2B2Br2+ 526.09564; found: 526.09549.
6a + 6b: 1H-NMR (400 MHz, C6D6): δ = 6.69 (s, 4H), 6.67 (s, 4H), 6.34 (s, 2H), 6.09 (d, 3JH-H = 1.6 Hz, 1H), 6.08 (d, 3JH-H = 1.6 Hz, 1H), 5.32 (d, 3JH-H = 1.6 Hz, 1H), 5.30 (d, 3JH-H = 1.6 Hz, 1H), 4.94 (s, 2H), 2.19 (s, 12H), 2.16 (s, 12H), 2.13 (s, 6H), 2.12 (s, 6H).13C{1H}-NMR (101 MHz, C6D6): δ = 138.9, 138.8, 138.6, 138.5, 127.6, 127.6, 117.3, 116.4, 116.3, 115.5, 21.2, 21.1, 20.9, 20.8. The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 38.5. HRMS: calc. for [M]+ C22H26N2B2Cl+ 410.16537; found: 410.16492.
7a + 7b: 1H-NMR (400 MHz, C6D6): δ = 7.86 (s, 2 H), 7.51−7.46 (m, 8 H), 7.19−7.15 (m, 10 H), 6.29 (s, 2H), 6.08 (d, 3JH-H = 1.68 Hz, 1H), 6.06 (d, 3JH-H = 1.64 Hz, 1H), 5.78 (d, 3JH-H = 1.60 Hz, 1H), 5.76 (d, 3JH-H = 1.70 Hz, 1H), 5.54 (s, 2 H). 13C{1H}-NMR (101 MHz, C6D6): δ = 133.3, 133.2, 130.2, 130.1, 127.9, 127.8, 118.0, 117.5, 116.8, 116.5. The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 36.5. HRMS: calc. for [M]+ C16H14N2B2Cl2+ 326.07147; found: 326.07069.

3.4. Synthesis of 8 and 9

Compounds 8 and 9 were synthesized in a similar manner as 3 and 4, with yields of 65% and 75%, respectively.
8: 1H-NMR (400 MHz, C6D6): δ = 7.93–7.90 (m, 2H), 7.85–7.83 (m, 1H), 7.24–7.13 (m, 2H), 2.14 (s, 3H), 1.68 (s, 3H), 1.62 (s, 3H), 1.52 (s, 3H), 0.19 (s, 9H). 13C{1H}-NMR (101 MHz, C6D6): δ = 132.5, 132.2, 131.9, 131.9, 131.6, 128.6, 128.0, 122.8, 121.8, 16.8, 16.7, 16.7, 16.5, 0.2. The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 35.9. HRMS: calc. for [M+H]+ C17H27N2BClSi+ 333.17196; found: 333.17233.
9: 1H-NMR (400MHz, C6D6): δ = 6.88 (s, 1H), 2.43 (s, 3H), 2.29 (s, 3H), 2.24 (s, 3H, Me of Dur), 2.10 (s, 3H), 2.06 (s, 3H), 1.71 (s, 3H), 1.48 (s, 3H), 1.44 (s, 3H), 0.25 (s, 9H). 13C{1H}-NMR (101 MHz, C6D6): δ = 133.8, 133.2, 133.1, 133.0, 132.8, 131.6, 131.4, 122.5, 122.0, 19.6, 19.4, 19.3, 19.3, 19.2, 18.1, 18.0, 16.1, 1.8. The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 37.7. HRMS: calc. for [M+H]+ C21H35N2BBrSi+ 433.18405; found: 433.18483.

3.5. Synthesis of 10

In the glove box, PhBCl2 (31.6 mg, 0.2 mmol, 2.0 equiv.) and 2,3,5,6-tetramethyl-1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene (2) (28.2 mg, 0.1 mmol, 1.0 equiv.) were added into C6D6 (0.6 mL) in a J. Young NMR tube. The mixture was rested overnight prior to the removal of the volatiles under vacuum to get 10 as a pale yellow solid (24.9 mg, 0.53 mmol, 53%).
10a + 10b: 1H-NMR (400 MHz, toluene-d8): δ = 7.72–7.20 (m, 8H), 7.11–7.06 (m, 12H), 1.95 (br, 12H), 1.45 (br, 12H). 13C{1H}-NMR (101 MHz, toluene-d8): δ = 133.2, 130.2, 129.1, 128.6, 128.2, 128.0, 127.9, 127.7, 125.4, 124.9, 20.8, 20.7, 20.3, 20.1. The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, toluene-d8): δ = 36.9. HRMS: calc. for [M+H]+ C20H23N2B2Cl2+ 383.14189; found: 383.14083.

3.6. Synthesis of 11a and 11b

In the glove box, 9 (37.6 mg, 0.1 mmol, 1 equiv.) and DurBBr2 (30.3 mg, 1 mmol, 1.0 equiv.) were added into C6D6 (0.6 mL) in a J. Young NMR tube. The mixture was rested overnight prior to the removal of the volatiles under vacuum to get 11 as a yellow oil with a yield of 76%. Mixture 11 contains two olefin isomers, 11a and 11b, with a ratio of about 1:0.3.
11a: 1H-NMR (400 MHz, C6D6): δ = 6.91 (s, 1H), 6.89 (s, 1H), 5.14–5.13 (m, 1H), 4.72–4.71 (m, 1H), 4.38 (s, 1H), 2.58 (s, 3H), 2.55 (s, 3H), 2.35 (s, 3H), 2.32 (s, 3H), 2.11 (s, 3H), 2.09 (s, 6H), 2.07 (s, 3H), 1.99 (d, 3JH-H = 0.8 Hz, 3H), 1.41 (d, 3JH-H = 0.8 Hz, 3H), 0.91 (d, 3JH-H = 6.6 Hz, 3H).
11b: 1H-NMR (400 MHz, C6D6): δ = 6.89 (s, 1H), 6.88 (s, 1H), 5.75–5.70 (m, 1H), 4.47–4.46 (m, 1H), 4.07–4.06 (m, 1H), 2.54 (s, 3H), 2.52 (s, 3H), 2.36 (s, 3H), 2.33 (s, 3H), 2.10 (s, 3H), 2.09 (s, 9H), 2.07 (s, 3H), 1.98 (d, 3JH-H = 1.1 Hz, 3H), 1.45 (d, 3JH-H = 1.1 Hz, 3H), 1.13 (d, 3JH-H = 8.0 Hz, 3H).
11a + 11b: 13C{1H}-NMR (101 MHz, C6D6): δ = 152.0, 151.4, 133.7, 133.6, 133.6, 133.5, 133.4, 133.3, 133.2, 133.2, 132.6, 132.2, 132.1, 132.0, 131.7, 131.4, 131.4, 131.1, 130.8, 129.9, 105.6, 101.7, 61.2, 60.8, 22.8, 22.6, 20.2, 20.1, 20.0, 19.7, 19.7, 19.6, 19.5, 19.4, 19.3, 19.3, 19.2, 19.2, 19.2, 19.1, 18.9, 18.1, 17.9, 15.4. The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 39.1. HRMS: calc. for [M+H]+ C28H39N2B2Br2+ 585.16402; found: 585.16454.

3.7. Synthesis of 13a and 13b

In the glove box, (TMS)2NBCl2 (86.4 mg, 0.2 mmol, 2 equiv.) and 1,1′-bis(trimethylsilyl)-1H,1′H-4,4′-bipyridinylidene (12) (30.2 mg, 0.1 mmol, 1 equiv.) were added into C6D6 (0.6 mL). The mixture was rested for 10 min prior to the removal of the volatiles under vacuum to get a yellowish green powder. The yellowish green powder was extracted with hexane, filtered, and the solvent was again removed under reduced pressure to yield 13 as a yellow powder (60.0 mg, 0.84 mmol, 84%). The ratio of 13a:13b is ca. 1:1.
13a + 13b: 1H-NMR (400 MHz, C6D6): δ = 6.87 (br, 8H), 5.79 (d,3JH-H = 8.1 Hz, 8H), 0.21 (s, 72H). 13C{1H}-NMR (101 MHz, C6D6): δ = 114.3, 111.3, 2.25 (C of TMS). The carbon atom directly attached to boron was not detected, likely due to quadrupolar broadening. 11B-NMR (128 MHz, C6D6): δ = 34.4. HRMS: calc. for [M]+ C22H44N4B2Cl2Si4+ 568.22007; found: 568.21887.

4. Conclusions

In summary, the reactions of electron-rich organosilicon compounds 1, 2, and 12 with various B-amino and B-aryl dihaloboranes were comprehensively studied. No direct evidence for the presence of divalent boron radical character could be obtained from NMR spectra and single-crystal structures. The rotational barrier around the exocyclic B–N bonds was studied by VT 1H-NMR spectroscopy, which revealed relatively small barriers for 10 and 13. The steric hindrance as well as the competition from additional B-amino functions were the main factors affecting the B–N rotational barrier. In addition, the reaction between 2 and DurBBr2 resulted in 11 via an isomerization process. Although this study does not access the desired biradial species, we believe that the novel B=N-containing products could act as an RXB• source upon the liberation of the aromatic linker (i.e., pyrazine and 4,4′-bipyridine). Studies of the mechanism of the isomerization reaction, as well as the application of 10 and 13 as RXB• transfer reagents to unsaturated organic substrates, are currently underway in our laboratory, and will be reported in due course.

Supplementary Materials

Supplementary materials are available online. Figures S1–S31: NMR spectra for 311, 13. Figures S32–S36: Variable-temperature 1H-NMR spectra for 7, 10, and 13. Figures S37–S41, Single crystal structure for 5, 911, and 13. Table S1: Crystal data for 5, 911, and 13.

Author Contributions

Q.Y. conceived and designed the experiments; L.M., X.Z., S.S., and W.M. performed the experiments; L.M., W.M., and Q.Y. analyzed the data; L.M., X.Z., and X.C. tested and refined single crystals. L.M., W.M., and Q.Y. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the start-up fund of SUSTech.

Acknowledgments

The authors thank Zhanying Ren, Yinhua Yang, and Jianfei Qu (SUSTech) for the VT NMR measurement, and Hua Li (SUSTech) for the HRMS measurement.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Krasowska, M.; Bettinger, H.F. Reactivity of borylenes toward ethyne, ethene, and methane. J. Am. Chem. Soc. 2012, 134, 17094–17103. [Google Scholar] [CrossRef] [PubMed]
  2. Krasowska, M.; Edelmann, M.; Bettinger, H.F. Electronically excited states of borylenes. J. Phys. Chem. A 2016, 120, 6332–6341. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, Y.; Robinson, G.H. Carbene stabilization of highly reactive main-group molecules. Inorg. Chem. 2011, 50, 12326–12337. [Google Scholar] [CrossRef] [PubMed]
  4. Bissinger, P.; Braunschweig, H.; Damme, A.; Dewhurst, R.D.; Kupfer, T.; Radacki, K.; Wagner, K. Generation of a carbene-stabilized bora-borylene and its insertion into a C–H bond. J. Am. Chem. Soc. 2011, 133, 19044–19047. [Google Scholar] [CrossRef] [PubMed]
  5. Bissinger, P.; Braunschweig, H.; Kraft, K.; Kupfer, T. Trapping the elusive parent borylene. Angew. Chem. Int. Ed. 2011, 50, 4704–4707. [Google Scholar] [CrossRef] [PubMed]
  6. Curran, D.P.; Boussonnière, A.; Geib, S.J.; Lacôte, E. The parent borylene: Betwixt and between. Angew. Chem. Int. Ed. 2012, 51, 1602–1605. [Google Scholar] [CrossRef]
  7. Braunschweig, H.; Claes, C.; Damme, A.; Deißenberger, A.; Dewhurst, R.D.; Hörl, C.; Kramer, T. A facile and selective route to remarkably inert monocyclic NHC-stabilized boriranes. Chem. Commun. 2015, 51, 1627–1630. [Google Scholar] [CrossRef]
  8. Frey, G.D.; Lavallo, V.; Donnadieu, B.; Schoeller, W.W.; Bertrand, G. Facile splitting of hydrogen and ammonia by nucleophilic activation at a single carbon center. Science 2007, 316, 439–441. [Google Scholar] [CrossRef] [Green Version]
  9. Légaré, M.-A.; Bélanger-Chabot, G.; Dewhurst, R.D.; Welz, E.; Krummenacher, I.; Engels, B.; Braunschweig, H. Nitrogen fixation and reduction at boron. Science 2018, 359, 896–900. [Google Scholar] [CrossRef] [Green Version]
  10. Légaré, M.-A.; Rang, M.; Bélanger-Chabot, G.; Schweizer, J.I.; Krummenacher, I.; Bertermann, R.; Arrowsmith, M.; Holthausen, M.C.; Braunschweig, H. The reductive coupling of dinitrogen. Science 2019, 363, 1329–1332. [Google Scholar] [CrossRef]
  11. Krasowska, M.; Bettinger, H.F. Ring enlargement of three-membered boron heterocycles upon reaction with organic π systems: Implications for the trapping of borylenes. Chem. Eur. J. 2016, 22, 10661–10670. [Google Scholar] [CrossRef] [PubMed]
  12. Timms, P.L. Boron-fluorine chemistry. I. Boron monofluoride and some derivatives. J. Am. Chem. Soc. 1967, 89, 1629–1632. [Google Scholar] [CrossRef]
  13. Timms, P.L. Chemistry of boron and silicon subhalides. Acc. Chem. Res. 1973, 6, 118–123. [Google Scholar] [CrossRef]
  14. Wang, Y.; Quillian, B.; Wei, P.; Wannere, C.S.; Xie, Y.; King, R.B.; Schaefer, H.F.; Schleyer, P.v.R.; Robinson, G.H. A stable neutral diborene containing a B=B double bond. J. Am. Chem. Soc. 2007, 129, 12412–12413. [Google Scholar] [CrossRef] [PubMed]
  15. Kinjo, R.; Donnadieu, B.; Celik, M.A.; Frenking, G.; Bertrand, G. Synthesis and characterization of a neutral tricoordinate organoboron isoelectronic with amines. Science 2011, 333, 610–613. [Google Scholar] [CrossRef]
  16. Woon, E.C.Y.; Tumber, A.; Kawamura, A.; Hillringhaus, L.; Ge, W.; Rose, N.R.; Ma, J.H.Y.; Chan, M.C.; Walport, L.J.; Che, K.H.; et al. Linking of 2-oxoglutarate and substrate binding sites enables potent and highly selective inhibition of Jmjc histone demethylases. Angew. Chem. Int. Ed. 2012, 51, 1631–1634. [Google Scholar] [CrossRef]
  17. Kong, L.; Li, Y.; Ganguly, R.; Vidovic, D.; Kinjo, R. Isolation of a bis(oxazol-2-ylidene)–phenylborylene adduct and its reactivity as a boron-centered nucleophile. Angew. Chem. Int. Ed. 2014, 53, 9280–9283. [Google Scholar] [CrossRef]
  18. Wang, H.; Zhang, J.; Lin, Z.; Xie, Z. Synthesis and structural characterization of carbene-stabilized carborane-fused azaborolyl radical cation and dicarbollyl-fused azaborole. Organometallics 2016, 35, 2579–2582. [Google Scholar] [CrossRef]
  19. Lu, W.; Li, Y.; Ganguly, R.; Kinjo, R. Alkene-carbene isomerization induced by borane: Access to an asymmetrical diborene. J. Am. Chem. Soc. 2017, 139, 5047–5050. [Google Scholar] [CrossRef]
  20. Wang, H.; Zhang, J.; Lin, Z.; Xie, Z. The synthesis and structure of a carbene-stabilized iminocarboranyl-boron(I) compound. Chem. Commun. 2015, 51, 16817–16820. [Google Scholar] [CrossRef]
  21. Ruiz, D.A.; Melaimi, M.; Bertrand, G. An efficient synthetic route to stable bis(carbene)borylenes [(L1)(L2)BH]. Chem. Commun. 2014, 50, 7837–7839. [Google Scholar] [CrossRef] [PubMed]
  22. Hahn, F.E.; Wittenbecher, L.; Boese, R.; Bläser, D. N,N′-bis(2,2-dimethylpropyl)benzimidazolin-2-ylidene: A stable nucleophilic carbene derived from benzimidazole. Chem. Eur. J. 1999, 5, 1931–1935. [Google Scholar] [CrossRef]
  23. Lavallo, V.; Canac, Y.; Donnadieu, B.; Schoeller, W.W.; Bertrand, G. Cyclopropenylidenes: From interstellar space to an isolated derivative in the laboratory. Science 2006, 312, 722–724. [Google Scholar] [CrossRef] [Green Version]
  24. Connelly, N.G.; Geiger, W.E. Chemical redox agents for organometallic chemistry. Chem. Rev. 1996, 96, 877–910. [Google Scholar] [CrossRef] [PubMed]
  25. Broggi, J.; Terme, T.; Vanelle, P. Organic electron donors as powerful single-electron reducing agents in organic synthesis. Angew. Chem. Int. Ed. 2014, 53, 384–413. [Google Scholar] [CrossRef] [PubMed]
  26. Eberle, B.; Hubner, O.; Ziesak, A.; Kaifer, E.; Himmel, H.J. What makes a strong organic electron donor (or acceptor)? Chem. Eur. J. 2015, 21, 8578–8590. [Google Scholar] [CrossRef]
  27. Tsurugi, H.; Saito, T.; Tanahashi, H.; Arnold, J.; Mashima, K. Carbon radical generation by d0 tantalum complexes with α-diimine ligands through ligand-centered redox processes. J. Am. Chem. Soc. 2011, 133, 18673–18683. [Google Scholar] [CrossRef]
  28. Tsurugi, H.; Tanahashi, H.; Nishiyama, H.; Fegler, W.; Saito, T.; Sauer, A.; Okuda, J.; Mashima, K. Salt-free reducing reagent of bis(trimethylsilyl)cyclohexadiene mediates multielectron reduction of chloride complexes of W(VI) and W(IV). J. Am. Chem. Soc. 2013, 135, 5986–5989. [Google Scholar] [CrossRef]
  29. Saito, T.; Nishiyama, H.; Tanahashi, H.; Kawakita, K.; Tsurugi, H.; Mashima, K. 1,4-bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadienes as strong salt-free reductants for generating low-valent early transition metals with electron-donating ligands. J. Am. Chem. Soc. 2014, 136, 5161–5170. [Google Scholar] [CrossRef]
  30. Tsurugi, H.; Mashima, K. A new protocol to generate catalytically active species of group 4–6 metals by organosilicon-based salt-free reductants. Chem. Eur. J. 2019, 25, 913–919. [Google Scholar] [CrossRef]
  31. Arteaga-Muller, R.; Tsurugi, H.; Saito, T.; Yanagawa, M.; Oda, S.; Mashima, K. New tantalum ligand-free catalyst system for highly selective trimerization of ethylene affording 1-hexene: New evidence of a metallacycle mechanism. J. Am. Chem. Soc. 2009, 131, 5370–5371. [Google Scholar] [CrossRef] [PubMed]
  32. Saito, T.; Nishiyama, H.; Kawakita, K.; Nechayev, M.; Kriegel, B.; Tsurugi, H.; Arnold, J.; Mashima, K. Reduction of (tBuN=)NbCl3(py)2 in a salt-free manner for generating Nb(IV) dinuclear complexes and their reactivity toward benzo[c]cinnoline. Inorg. Chem. 2015, 54, 6004–6009. [Google Scholar] [CrossRef] [PubMed]
  33. Tanahashi, H.; Ikeda, H.; Tsurugi, H.; Mashima, K. Synthesis and characterization of paramagnetic tungsten imido complexes bearing α-diimine ligands. Inorg. Chem. 2016, 55, 1446–1452. [Google Scholar] [CrossRef] [PubMed]
  34. Weyenberg, D.R.; Toporcer, L.H. The synthesis of 3,6-disilyl-1,4-cyclohexadienes by the trapping of benzene anion-radicals. J. Am. Chem. Soc. 1962, 84, 2843–2844. [Google Scholar] [CrossRef]
  35. Braunschweig, H.; Ye, Q.; Radacki, K. High yield synthesis of a neutral and carbonyl-rich terminal arylborylene complex. Chem. Commun. 2012, 48, 2701–2703. [Google Scholar] [CrossRef]
  36. Cowherd, F.G.; Von Rosenberg, J.L. Mechanism of iron pentacarbonyl-catalyzed 1,3-hydrogen shifts. J. Am. Chem. Soc. 1969, 91, 2157–2158. [Google Scholar] [CrossRef]
  37. Hvistendahl, G.; Williams, D.H. Energy barrier to symmetry-forbidden 1,3-hydrogen shifts in simple oxonium ions. Metastable peaks from fast dissociations. J. Am. Chem. Soc. 1975, 97, 3097–3101. [Google Scholar] [CrossRef]
  38. Laguerre, M.; Dunogues, J.; Calas, R.; Duffaut, N. Silylation d′hydrocarbures mono-aromatiques mono- ou disubstitues. J. Organomet. Chem. 1976, 112, 49–59. [Google Scholar] [CrossRef]
  39. Lis, A.V.; Gostevskii, B.A.; Albanov, A.I.; Yarosh, N.O.; Rakhlin, V.I. Synthesis of volatile bis[bis(trimethylsilyl)amide]-substituted boron derivatives. Russ. J. Gen. Chem. 2017, 87, 353–356. [Google Scholar] [CrossRef]
  40. Brown, C.; Cragg, R.H.; Miller, T.J.; Smith, D.O.N. Organoboron compounds: Xxii. A 13C NMR study of some dialkylaminophenylboranes. J. Organomet. Chem. 1983, 244, 209–215. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from the authors.
Scheme 1. Proposed products from the reactions of ArBX2 with 1 and 2, and two possible bonding modes A and B between the central C4N2 ring and the boron centers.
Scheme 1. Proposed products from the reactions of ArBX2 with 1 and 2, and two possible bonding modes A and B between the central C4N2 ring and the boron centers.
Molecules 25 02875 sch001
Scheme 2. Synthesis of 311.
Scheme 2. Synthesis of 311.
Molecules 25 02875 sch002
Scheme 3. Synthesis of 13.
Scheme 3. Synthesis of 13.
Molecules 25 02875 sch003
Figure 1. Molecular structures of 5a (bottom) and 5b (top) in the solid state (ellipsoids set at 50% probability). Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (°) for 5a: B1–N1 1.423(8), N1–C2 1.385(7), C1–C2 1.311(9), C1–C2–N1 123.1(5), C2–N1–C1* 113.9(4); for 5b: B2–N2 1.400(8), N2–C4 1.411(6), N2–C3 1.415(7), C4–C4* 1.328(8), C3–C3* 1.326(8), C4*–C4–N2 123.7(4), C4–N2–C3 112.6(4), N2–C3–C3* 123.7(5).
Figure 1. Molecular structures of 5a (bottom) and 5b (top) in the solid state (ellipsoids set at 50% probability). Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (°) for 5a: B1–N1 1.423(8), N1–C2 1.385(7), C1–C2 1.311(9), C1–C2–N1 123.1(5), C2–N1–C1* 113.9(4); for 5b: B2–N2 1.400(8), N2–C4 1.411(6), N2–C3 1.415(7), C4–C4* 1.328(8), C3–C3* 1.326(8), C4*–C4–N2 123.7(4), C4–N2–C3 112.6(4), N2–C3–C3* 123.7(5).
Molecules 25 02875 g001
Figure 2. Molecular structures of 9, 10a, and 11 in the solid state (ellipsoids set at 50% probability). Hydrogen atoms, except for the C(sp3)–H and the olefinic H in 11a and 11b, are omitted for clarity. Selected bond lengths (Å) and angles (°) for 9: B1–N1 1.376(7), N1–C1 1.451(6), C1–C2 1.343(7), N2–C2 1.434(6), N2–Si1 1.759(5), C2–C1–N1 115.3(4), C1–N1–C1* 111.1(4); for 10a: B1–N1 1.401(2), N1–C2 1.4494(17), C1–C2 1.328(2), N1–C1* 1.4513(18), C1–C2–N1 116.25(12), C2–N1–C1* 110.63(11).
Figure 2. Molecular structures of 9, 10a, and 11 in the solid state (ellipsoids set at 50% probability). Hydrogen atoms, except for the C(sp3)–H and the olefinic H in 11a and 11b, are omitted for clarity. Selected bond lengths (Å) and angles (°) for 9: B1–N1 1.376(7), N1–C1 1.451(6), C1–C2 1.343(7), N2–C2 1.434(6), N2–Si1 1.759(5), C2–C1–N1 115.3(4), C1–N1–C1* 111.1(4); for 10a: B1–N1 1.401(2), N1–C2 1.4494(17), C1–C2 1.328(2), N1–C1* 1.4513(18), C1–C2–N1 116.25(12), C2–N1–C1* 110.63(11).
Molecules 25 02875 g002
Figure 3. Molecular structure of 13a in the solid state (ellipsoids set at 50% probability). Selected bond lengths (Å) and angles (°) for 13a: B1–N1 1.459(8), B1–N2 1.392(10), Si1–N2 1.757(5), Si2–N2 1.770(5), N1–C1 1.402(8), N1–C5 1.414(8), C1–C2 1.332(8), C4–C5 1.347(9), C2–C3 1.452(9), C3–C4 1.444(9), C3–C3* 1.376(12), C1–N1–C5 115.7(5), C2–C1–N1 123.3(6), C1–C2–C3 122.9(6), C4–C3–C2 112.7(5), C5–C4–C3 123.3(6), C4–C5–N1 122.1(6).
Figure 3. Molecular structure of 13a in the solid state (ellipsoids set at 50% probability). Selected bond lengths (Å) and angles (°) for 13a: B1–N1 1.459(8), B1–N2 1.392(10), Si1–N2 1.757(5), Si2–N2 1.770(5), N1–C1 1.402(8), N1–C5 1.414(8), C1–C2 1.332(8), C4–C5 1.347(9), C2–C3 1.452(9), C3–C4 1.444(9), C3–C3* 1.376(12), C1–N1–C5 115.7(5), C2–C1–N1 123.3(6), C1–C2–C3 122.9(6), C4–C3–C2 112.7(5), C5–C4–C3 123.3(6), C4–C5–N1 122.1(6).
Molecules 25 02875 g003
Figure 4. Variable-temperature 1H-NMR (400 MHz, toluene-d8) spectra of 7, 10, and 13.
Figure 4. Variable-temperature 1H-NMR (400 MHz, toluene-d8) spectra of 7, 10, and 13.
Molecules 25 02875 g004
Table 1. Rotational barrier of 7, 10, and 13.
Table 1. Rotational barrier of 7, 10, and 13.
CompoundTc△ν△G
7>80 °C (353 K)85.0 Hz>71.56 kJ/mol
1032 °C (305 K)243.5 Hz58.79 kJ/mol
1320 °C (293 K)96.0 Hz58.65 kJ/mol
Tc = coalescence temperature; Δν = the separation in hertz between the two singlets in the absence of exchange; ∆G = rotational barrier.

Share and Cite

MDPI and ACS Style

Ma, L.; Zhang, X.; Ming, W.; Su, S.; Chang, X.; Ye, Q. Reactions of Dihaloboranes with Electron-Rich 1,4-Bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadienes. Molecules 2020, 25, 2875. https://doi.org/10.3390/molecules25122875

AMA Style

Ma L, Zhang X, Ming W, Su S, Chang X, Ye Q. Reactions of Dihaloboranes with Electron-Rich 1,4-Bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadienes. Molecules. 2020; 25(12):2875. https://doi.org/10.3390/molecules25122875

Chicago/Turabian Style

Ma, Li, Xiaolin Zhang, Wenbo Ming, Shengxin Su, Xiaoyong Chang, and Qing Ye. 2020. "Reactions of Dihaloboranes with Electron-Rich 1,4-Bis(trimethylsilyl)-1,4-diaza-2,5-cyclohexadienes" Molecules 25, no. 12: 2875. https://doi.org/10.3390/molecules25122875

Article Metrics

Back to TopTop