Skip to content
Publicly Available Published by De Gruyter May 9, 2015

Organoboranes and tetraorganoborates studied by 11B and 13C NMR spectroscopy and DFT calculations

  • Bernd Wrackmeyer EMAIL logo

Abstract

Care should be taken on recording the sometimes elusive 13C NMR signals for boron-bonded carbon atoms, since it is easy to extract information about coupling constants 1J(13C,11B) by measuring the respective line widths of 13C(B-C) and 11B NMR signals. This information can be confirmed by quantum-chemical calculations [B3LYP (6-311+G(d,p) level of theory] of nJ(13C,11B) in organoboranes and tetraorganoborates. For the latter, the signs for n = 2, 3, 4 were experimentally determined.

1 Introduction

The considerable importance of organoboron compounds in synthesis [1–3] requires careful characterization by reliable and efficient analytical techniques. In this context, 1H NMR spectroscopy is helpful, although many organyl groups do not give simple first order spectra. In spite of the quadrupolar nature of 11B (spin I = 3/2), 11B NMR spectra are extremely useful, since in many cases the chemical shifts δ11B are diagnostic for the surroundings of the boron atom [4]. Typically, 13C NMR spectroscopic measurements [5] often provide conclusive evidence for the solution-state molecular structures. However, even nowadays, with modern powerful NMR spectrometers, one finds frequently the message that the 13C NMR signal of a carbon atom linked to boron was not observed. Indeed, this signal can be rather broad because of partially relaxed scalar 13C-11B spin-spin coupling, and careful measurements are necessary, in particular if 13C nuclear spin relaxation tends to be slow. One can distinguish three cases, depending on 11B nuclear spin relaxation: (i) long relaxation time TQ(11B), typical for highly symmetrical surroundings (or with a small electric field gradient) of the boron atom as in borates, e.g. in [BPh4], for which all 13C–11B spin–spin couplings over one-, two-, three- and even four bonds are perfectly resolved [6]; (ii) moderately fast 11B nuclear spin relaxation, e.g. in trigonal boranes of low molecular weight, which allows sometimes to observe residual splitting of the 13C NMR signal [7–9] (see also Fig. 1); (iii) rather fast 11B nuclear spin relaxation, observed e.g. for most arylboranes, leading to more or less broadened 13C NMR signals for boron-bonded carbon atoms (see examples for 13C NMR spectra in [10, 11]).

In the latter cases (iii), the coupling constants 1J(13C,11B) can be evaluated using eq. (1) [12] with TQ(11B) = [πh1/2(11B)]–1, since scalar relaxation of the second kind (T2SC(13C) [13]) is by far the major source of broadening of these 13C(B-C) NMR signals. In the absence of chemical exchange and when the line width exceeds 10–15 Hz, other contributions to the line width Δνb(13C) can be safely neglected, an advantage of the magnetically diluted 13C nucleus. The data 1J(13C,11B) thus obtained can be compared with calculated data based on quantum-chemical calculations of coupling constants [14, 15], to be checked by comparison with directly measured data 1J(13C,11B) for the cases (i) and (ii). The numerous examples depicted in Table 1 indicate that both evaluation of 1J(13C,11B) from line widths as well as quantum-chemical calculations are fairly reliable.

Fig. 1: Part of the 50.3 MHz 13C{1H} NMR spectrum of methylamino-dimethylborane (24 in Table 1; 5 % in C6D6 at 27 °C) showing the 13C(BMe) NMR signals (hindered rotation about the B–N bond) split by partially relaxed 13C–11B spin–spin coupling (the distance in Hz between the inner lines represents the coupling constant).
Fig. 1:

Part of the 50.3 MHz 13C{1H} NMR spectrum of methylamino-dimethylborane (24 in Table 1; 5 % in C6D6 at 27 °C) showing the 13C(BMe) NMR signals (hindered rotation about the B–N bond) split by partially relaxed 13C–11B spin–spin coupling (the distance in Hz between the inner lines represents the coupling constant).

Table 1

11B and 13C NMR parameters of some organoboranes and -borates.

Compoundaδ11Bδ13CnJ(13C,11B) (n) in Hz calcd.b [exp.]c
1 BMe386.014.849.1 (1) [52.0]
2 [BMe4]–20.713.341.0 (1) [39.6]
3 BEt386.519.8 (BCH2), 8.549.1 (1) [50.0], –1.2 (2)
4 [BEt4]–17.517.6 (BCH2), 11.742.1 (1) [40.7], –1.0 (2) [1.1]
5 BnPr386.031.1 (BCH2), 18.1, 17.649.0 (1) [50.0], –1.0 (2), 2.2 (3)
6 [BnPr4]–17.833.6 (BCH2), 21.6, 18.841.1 (1) [41.0], –0.8 (2) [1.0], 4.8 (3) [4.7]
7 BPh368.0143.3 (br, i), 138.4 (o), 127.3 (m), 131.1 (p)64.9 (1) [66.0], 2.9 (2), 3.4 (3), –0.5 (4)
8 [BPh4]– d–8.0164.8 (i), 128.8 (o), 129.8 (m), 125.8 (p),51.1 (1) [49.4], 1.8 (2) [1.5], 2.9 (3) [2.7], –0.5 (4) [0.5]
9
44.3147.0 (br, 2), 129.9 (3), 112.3 (4), 132.1 (5), 35.2 (NMe)76.6 (1) [75.0], 4.9 (2), 3.5 (3), 2.8 (4)
10
–18.8162.7 (2), 116.9 (3), 106.7 (4), 119.5 (5), 32.5 (NMe)58.7 (1) [57.5], 3.4 (2) [3.5], 2.7 (3) [2.8], 2.0 (4) [2.0]
11
35.0166.7 (br, 2), 132.5 (3), 111.4 (4), 148.9 (5)84.6 (1) [83.0], 7.2 (2), 2.4 (3), 3.8 (4)
12
–19.9185.6 (2), 110.7 (3), 106.8 (4), 137.9 (5)63.2 (1) [64.0], 5.6 (2) [5.8], 2.1 (3) [2.0], 2.4 (4) [2.5]
13
47.3144.2 (br, 2), 141.5 (3), 128.9 (4), 136.4 (5)76.4 (1) [77.0], 4.1 (2), 4.3 (3), 2.5 (4)
14
–14.1179.9 (2), 127.4 (3), 122.3 (4), 127.0 (5)57.5 (1) [58.0], 5.8 (2) [6.0], 2.9 (3) [3.0], 2.5 (4) [2.5]
15
57.0145.0 (br, 2), 140.0 (3), 129.0 (4),136.1 (5), 12.3 (br, BMe)73.9 (1) [71.5], 4.1 (2), 4.1 (3), 22.6 (4)Me: 53.6 (1) [54.0]
16
49.0148.7 (br, 2), 145.1 (3), 128.0 (4), 136.7 (5)

Ph: 146.1 (br, i), 136.8 (o), 129.7 (m), 130.7 (p)
75.1 (1) [73.0], 4.1 (2), 4.2 (3), 2.6 (4)

Ph: 66.9 (1) [65.0], 2.9 (2), 3.5 (3), –0.6 (4)
17
35.2142 (br, 2), 129.8 (3), 127.8 (4), 135.2 (5)

NEt2: 44.5, 16.2
79.9 (1) [78.0], 3.3 (2), 4.5 (3), 2.1 (4)
18
48.5139.1 (br, 2), 143.0 (3), 128.6 (4), 140.2 (5)107.7 (1) [110.0], 5.4 (2), 5.7 (3), 3.8 (4)
19 [B(C≡C–Ph)4]– d–31.0102.8 (B–C≡), 94.1 (≡C–Ph)76.7 (1) [70.0], 15.7 (2) [14.0]
20e
62.1139.5 (br, 9a), 153.4 (4a), 133.0 (2), 129.3 (3), 120.7 (4)76.20 (1) [75.0], 8.2 (2), 3.9 (2), 3.8 (3), 3.1 (3), –0.3 (4)
21 EtBCl263.428.1, 8.371.8 (1) [74.0], –1.0 (2)
22 PhBCl255.0134.0 (br, i), 126.8 (o), 136.8 (m), 134.9 (p)86.6 (1) [88.0], 4.5 (2), 6.0 (3),–0.6 (4)
23 Me2B–NH247.15.655.6 (1) [59.0]
24 Me2B–N(H)Me45.71.5, 6.5, 29.4 (NMe)57.0 (1) [56.0], 56.0 (1) [56.0]
25 Et2B–NH247.311.2 (BCH2), 8.755.7 (1) [57.0], –1.2 (2)
26 [Et3B–NH2]– d–13.521.5 (BCH2), 10.144.0 (1) [43.0], –1.9 (2)

aBoranes measured in CDCl3 solutions, except 2325 (in C6D6), and borates as lithium salts in [D8]THF; bcalculated for optimized geometries at the B3LYP/6-311+G(d,p) level of theory; cdirect measurements from 13C{1H} NMR spectra for 1, 3, 21, 2325 (± 2 Hz); and all borates (± 1 Hz); for all other boranes, the experimental coupling constants 1J(13C,11B) were evaluated (± 3 Hz) from line widths measured from 13C{1H} and 11B{1H} NMR spectra; dsodium salt; eRef. [10].

(1)Δνb(13C)=(4/3) π 3/2(3/2+1) [J(13C,11B)]2 [TQ(11B)] (1)

Expectedly, all calculated coupling constants nJ(13C,11B) for n = 1 possess a positive sign [7, 9, 16]. For n = 2, calculations show negative or positive signs depending on the nature of the intervening carbon atom. In the cases of [BPh4](8) and the heteroaryborates 10, 12, 14, the calculated positive sign of nJ(13C,11B) (n = 2, 3, 4) can be confirmed by appropriate selective heteronuclear 1H{11B} experiments [17], observing 13C satellite signals in the 1H NMR spectra (Fig. 2).

Fig. 2: Two representative examples of the determination of coupling signs nJ(13C,11B) relative to 1J(13C,1H) (known as > 0) by selective low-power irradiation of 11B transitions at low or high frequency, sharpening the respective low- or high-frequency 13C satellite signal in the 1H NMR spectrum. 1H and 11B are the “active” nuclei and 13C (neither irradiated nor observed) is the "passive" nucleus.
Fig. 2:

Two representative examples of the determination of coupling signs nJ(13C,11B) relative to 1J(13C,1H) (known as > 0) by selective low-power irradiation of 11B transitions at low or high frequency, sharpening the respective low- or high-frequency 13C satellite signal in the 1H NMR spectrum. 1H and 11B are the “active” nuclei and 13C (neither irradiated nor observed) is the "passive" nucleus.

2 Experimental

Most compounds (Table 1) were commercially available or obtained following literature procedures [1]. Heteroarylboranes were accessible as reported [18, 19]. Heteroarylborates were prepared as lithium salts in situ, as described previously [20], extracted with ether, dried and dissolved in [D8]THF. NMR measurements: Bruker WP200, ARX 250 and DRX 500 spectrometers in 5 mm (o.d.) tubes at 23 or 27 °C. Chemical shifts are given in ppm relative to BF3-OEt2 (δ11B = 0 ppm for Ξ(11B) = 32.083971 MHz) and SiMe4 [δ13C(CDCl3) = 77.0, δ13C([D8]THF) = 67.4, 25.2 ppm]. Typically, 1000 to 4000 transients were used to record 13C{1H} NMR spectra with a S/N ratio sufficient for line width measurements of broadened 13C(B-C) NMR signals. The decoupler power for selective 1H{11B} experiments (Bruker DRX 500) was carefully calibrated by observing the desired effects for the central signals of the respective multiplets before measuring the 13C satellite signals in the 1H NMR spectra of the tetraorganoborates. In the case of 4J(13C,11B) for [BPh4] (8), the calculated negative sign of the coupling constant could not be confirmed, since there was no differential sharpening of the 13C satellites upon selective 11B irradiation. Presumably the coupling constant 5J(11B,1H) is too small (calcd. 0.25 Hz),

All quantum-chemical calculations were carried out using the Gaussian 09 program package [21].


Corresponding author: Bernd Wrackmeyer, Anorganische Chemie II, Universität Bayreuth, D-95440 Bayreuth, Germany, e-mail:

Acknowledgments

Support by the Deutsche Forschungsgemeinschaft is gratefully acknowledged.

References

[1] R. Köster, Houben-Weyl Methoden der Organischen Chemie, Organobor-Verbindungen I, Band 13, Teil 3a, Thieme, Stuttgart, 1982.Search in Google Scholar

[2] H. C. Brown, Organic Syntheses via Boranes, Wiley, New York 1975.Search in Google Scholar

[3] A. Suzuki, Heterocycles2010, 80, 15–43.10.3987/COM-09-S(S)SummarySearch in Google Scholar

[4] H. Nöth, B. Wrackmeyer, Nuclear Magnetic Resonance Spectroscopy of Boron Compounds, in NMR – Basic Principles and Progress, (Eds.: P. Diehl, E. Fluck, R. Kosfeld), Vol. 14, Springer, Berlin, 1978.10.1007/978-3-642-66757-2Search in Google Scholar

[5] H.-O. Kalinowski, S. Berger, S. Braun, 13C-NMR-Spektroskopie, Thieme, Stuttgart, 1984.Search in Google Scholar

[6] J. D. Odom, L. W. Hall, P. D. Ellis, Org. Magn. Reson.1974, 6, 360–361.Search in Google Scholar

[7] A. J. Zozulin, H. J. Jakobsen, T. F. Moore, A. R. Garber, J. D. Odom, J. Magn. Reson.1980, 41, 458–466.Search in Google Scholar

[8] L. W. Hall, D. W. Lowman, P. D. Ellis, J. D. Odom, Inorg. Chem. 1975, 14, 580–583.Search in Google Scholar

[9] W. McFarlane, H. Nöth, B. Wrackmeyer, Chem. Ber. 1975, 108, 3831–3841.Search in Google Scholar

[10] B. Wrackmeyer, P. Thoma, R. Kempe, G. Glatz, Collect. Czech. Chem. Commun.2010, 75, 743–756.Search in Google Scholar

[11] B. Wrackmeyer, P. Thoma, W. Milius, Z. Naturforsch. 2013, 68b, 493–502.Search in Google Scholar

[12] V. Mlynarek, Prog. Nucl. Magn. Reson. Spectrosc.1986, 18, 277–305.Search in Google Scholar

[13] A. Abragam, The Principles of Nuclear Magnetism, Oxford University Press, Oxford, 1961, pp. 305–315.Search in Google Scholar

[14] R. H. Contreras, V. Barone, J. C. Facelli, J. E. Peralta, Annu. Rep. NMR Spectrosc. 2003, 51, 167–260.Search in Google Scholar

[15] T. Helgaker, M. Jaszunski, M. Pecul, Progr. NMR Spectrosc. 2008, 53, 249–268.Search in Google Scholar

[16] J. D. Kennedy, W. McFarlane, G. S. Pyne, B. Wrackmeyer, J. Chem. Soc. Dalton Trans.1975, 386–390.10.1039/DT9750000386Search in Google Scholar

[17] W. McFarlane, Annu. Rev. NMR Spectrosc. 1968, 1, 135–163.Search in Google Scholar

[18] B. Wrackmeyer, H. Nöth, Chem. Ber. 1976, 109, 1075–1088.Search in Google Scholar

[19] T. Köhler, J. Faderl, H. Pritzkow, W. Siebert, Eur. J. Inorg. Chem.2002, 2942–2946.10.1002/1099-0682(200211)2002:11<2942::AID-EJIC2942>3.0.CO;2-PSearch in Google Scholar

[20] G. E. Pacey, C. E. Moore, Anal. Chim. Acta1979, 105, 353–359.10.1016/S0003-2670(01)83766-1Search in Google Scholar

[21] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox, Gaussian 09 (revision A.02), Gaussian, Inc., Wallingford CT (USA), 2010.Search in Google Scholar

Received: 2015-2-26
Accepted: 2015-4-2
Published Online: 2015-5-9
Published in Print: 2015-6-1

©2015 by De Gruyter

Downloaded on 8.5.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2015-0040/html
Scroll to top button