Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Mar 5.
Published in final edited form as: Chem Commun (Camb). 2019 Mar 5;55(20):2892–2895. doi: 10.1039/c8cc10087j

Synthesis of 9-Borafluorene Analogues Featuring a Three-Dimensional 1,1’-Bis(o-carborane) Backbone

Sam Yruegas a, Jonathan C Axtell b, Kent O Kirlikovali b, Alexander M Spokoyny b,c, Caleb D Martin a
PMCID: PMC6624192  NIHMSID: NIHMS1035562  PMID: 30706915

Abstract

The synthesis of [1,1’-bis(o-carboranyl)]boranes was achieved through the deprotonation of 1,1’-bis(o-carborane) reagents followed by salt metathesis with (iPr)2NBCl2. X-ray crystallography confirms planar central BC4 rings and Gutmann-Beckett studies reveal an increase in Lewis acidity at the boron center in comparison to their biphenyl congener, 9-borafluorene.


Polyhedral carborane clusters are viewed as three-dimensional aromatic analogues to the ubiquitous two-dimensional aromatic arenes (e.g. benzene).1 These species share high delocalization within the cage and ring resulting in high kinetic stability.2 The significant difference is that carboranes exhibit a aromaticity while benzene is a classical n aromatic molecule. Due to their unique steric profile and electronic structure, o-carboranes have been explored as a substitute for phenyl groups in molecules. The lability of the C-H vertices (pKa = 22 c.f. benzene = 43) of o-carborane facilitates selective derivatization to incorporate carboranes into molecular architectures.3 1,1’-Bis(o-carborane, B) can be viewed as a three-dimensional analogue to a biphenyl unit, a common ligand scaffold in organometallic chemistry (A, Figure 1).4 The facile manipulation and high stability has resulted in complexes featuring B being investigated in medicine and electronic materials.2a, 5

Fig 1.

Fig 1.

Relationship of biphenyl (A) to 1,1’-bis(o-carborane) (B) and the corresponding chelated boranes investigated in this work.

9-Borafluorenes (1A) contain a biphenyl backbone linked by a tricoordinate boron center and have been recognized as attractive targets for molecular sensors6, reagents for the synthesis of polycyclic aromatic hydrocarbons7 as well as components in organic light emitting diodes (OLEDs)8 and organic photovoltaics (OPVs).9 The vacant pz orbital on the boron center extends conjugation throughout the three fused rings. We envisioned that 1,1’-bis(o-carborane) could replace the biphenyl framework in 9-borafluorenes to generate a species with a three-dimensional backbone.

The initial strategies to access the target [1,1’-bis(o-carboranyl)]boranes were inspired by effective methods for the synthesis of 9-borafluorenes, specifically transmetallation of a stannole or dilithiated species with RBX2.10 The corresponding [1,1’-bis(o-carboranyl)]stannole4j was recently reported and the [1,1’-bis(o-carboranyl)]dilithium species11 has been generated and utilized in situ. Unfortunately, all attempts to access the [1,1’-bis(o-carboranyl)]borane via these reagents were unsuccessful (Tables S-1 and S-2). In addition, the transmetallation reaction with the [1,1’-bis(o-carboranyl)]magnesium species did not generate the desired boracycle (Table S-3). Potassium bis(trimethylsilyl)amide [K(HMDS)] is also an effective base for the deprotonation of the C-H vertices and the resultant salt, Kz[B], is easier to generate and offers enhanced solubility in comparison to the dilithiated reagent.4k,12 After several attempts using a variety of conditions (Table S-4), the room temperature generation of K2[B] in THF followed by addition of (iPr)2NBCl2 proved to be an effective method to furnish the desired [1,1’-bis(o-carboranyl)]borane 1B. Acquiring a 11B{1H} NMR spectrum of the crude reaction mixture showed a three-coordinate peak at 32.9 ppm, slightly shifted from (iPr)2NBCl2 (31.3 ppm), coupled with the disappearance of one of the diagnostic signals corresponding to B (−2.2 ppm) and emergence of a doublet at 1.8 ppm, suggesting restricted rotation about the C-C bond in B.13 After isolation, the product was dissolved in CDCl3 and the subsequent 1H NMR spectrum contained no C-H carborane signal at 3.51 ppm, indicating successful deprotonation of the carboranyl moieties and the product was isolated in 89% yield (Scheme 1). The identity of 1B was further confirmed based on single crystal X-ray diffraction studies (Figure 2). The synthetic route was compatible with the octa-methylated variant 1BMe814 featuring a 11B{1H} NMR resonance at 33.7 ppm corresponding to the (iPr)2NB-center, and a singlet at 6.3 ppm resulting from k2-C,C’-chelation of the bis(o-carborane). X-ray diffraction studies confirmed the structural identity of 1BMe8, which was isolated in 67% yield (Figure 2).

Scheme 1.

Scheme 1.

Synthesis of 1B and 1BMe8.

Fig 2.

Fig 2.

Solid-state structures of 1B and 1BMe8. Thermal ellipsoids are depicted at 50% probability and hydrogen atoms are removed for clarity. The diisopropyl group in 1BMe8 is positionally disordered and only the major component is shown.

A notable structural feature of 1B and 1BMe8 are highly planar central BC4 rings (maximum deviation from planarity = 029 Å and 0.011 Å, respectively), which is comparable to their borafluorene counterpart 1A (0.020 Å). The boron atom of the central ring and adjacent nitrogen atom of 1B are trigonal planar [Σangles: B(1) = 360.0(18)° and N(1) = 360.0(17)°, Table 1]. Positional disorder of the isopropyl groups on the nitrogen atom of 1BMe8 prevents an in-depth analysis of the metrical parameters of the substituents. The endocyclic carbon-carbon bonds of 1B and 1BMe8 are longer than 1A10b [1B: C(1)-C(2) 1.649(3) Å, C(2)-C(3) 1.528(3) Å, and C(3)-C(4) 1.649(3) Å, 1BMe8:, C(1)-C(2) 1.652(3) Å, C(2)-C(3) 1.524(3) Å, and C(3)-C(4) 1.646(3) Å, 1A: C(1)-C(2) 1.418(3) Å, C(2)-C(3) 1.474(3) Å, and C(3)-C(4) 1.413(3) Å] but contracted from the parent B3j [C(1)-C(2) 1.630(3) Å, C(2)-C(3) 1.528(3) Å, and C(3)-C(4) 1.649(3) Å]. The B-N bond lengths of 1B and 1BMe8 are slightly shorter compared to previously reported B-N length of 1A [1.371(3) Å and 1.384(4) Å c.f. 1.396(3) Å]10b, 15, indicating strong π-donation from the nitrogen lone pair to boron.16

Table 1.

Salient bond lengths (Å) and angles [°] in compounds 1B, 1BMe8, and 1A.

1B 1BMe8 1A

B(1)-C(1) 1.631(3) 1.622(4) 1.593(3)
C(1)-C(2) 1.649(3) 1.652(3) 1.418(3)
C(2)-C(3) 1.528(3) 1.524(3) 1.474(3)
C(3)-C(4) 1.649(3) 1.646(3) 1.413(3)
C(4)-B(1) 1.630(3) 1.626(4) 1.601(3)
B(1)-N(1) 1.371(3) 1.384(4) 1.396(3)
N(1)-B(1)-C(4) 126.06(19) 125.50(2) 128.97(13)
C(1)-B(1)-N(1) 125.61(18) 125.40(2) 127.51(19)
C(1)-B(1)-C(4) 108.33(16) 109.00(2) 103.44(17)
B(1)-N(1)-C(5) 119.94(17) 120.90(2)
B(1)-N(1)-C(8) 120.09(18) 119.76(18)
C(5)-N(1)-C(8) 119.96(16) 119.35(19)

The UV-Vis spectra of 1B and 1BMe8 in CH2Cl2 (Figure 3A) exhibit absorption maxima at 232 and 233 nm, respectively, blueshifted from 1A (248 nm).10b Cyclic voltammetry (CV) measurements conducted on 1B show a quasi-reversible one-electron reduction wave at −0.93 V and a second irreversible one-electron reduction at −1.86 V versus the ferrocenium/ferrocene couple (Fc+/Fc). In comparison, 1BMe8 exhibits only an irreversible reduction at −2.09 V whereas 1A showed only a reversible reduction at −2.95 V, indicating that the bis(o-carboranyl) backbone imparts an electron-withdrawing effect facilitating reduction (Figure 3B).10b

Fig 3.

Fig 3.

A) UV-Vis absorption emission spectra for 1B and 1BMe8 obtained from solutions of CH2CI2 (λ= 232 and 233 nm respectively). B) Cyclic voltammograms of 1B and 1BMe8 recorded in anhydrous tetrahydrofuran with 0.1 M [NnBu4l [PF6] and referenced to the ferrocenium/ferrocene redox couple (Fc+/Fc; scan rate = 0.1 V/s).

In order to understand the electronic effects of the bis(o-carboranyl) ligand scaffold, density functional theory (DFT) calculations were carried out. The geometries of 1A, 1B, and 1BMe8 were optimized based on the X-ray structure of 1B at the PBE-D3(BJ):TZP level, and single-point calculations were carried out at the B3LYP-D3(BJ):TZ2P level of theory (Figure S-13). The frontier orbital diagrams for 1B and 1BMe8 are similar, where the highest occupied molecular orbital (HOMO) is predominantly of π-character with respect to the B-N fragment, and the lowest occupied molecular orbital (LUMO) primarily resides on the bis(o-carboranyl) borane fragment. In contrast to the HOMO for 1A is entirely on the biphenyl fragment with no contribution from the amine, and the LUMO for 1A is localized on the biphenyl borane fragment. The HOMO-LUMO gaps for 1B and 1BMe8 are comparable (5.99 eV and 6.03 eV, respectively), and significantly larger than 1A (4.17 eV). These data corroborate similar absorption maxima for 1B and 1BMe8 as well as a bathochromic shift relative to the absorption maximum of 1A (Figure 3A). The calculated higher-lying LUMO for 1BMe8 (−1.74 eV) relative to that of 1B (−2.05 eV) is consistent with the observed more negative reduction potential for 1BMe8 (−2.09 V and −1.86 V, respectively; Figure 3B).

To experimentally gauge Lewis acidity, the Gutmann-Beckett method was utilized.17 This method involves the addition of an excess of Et3PO to a solution of the borane and monitoring the change in chemical shift of the 31P{1H} NMR signal (δ31Psample - 41.0). Multiplying this value by 2.21 gives the acceptor number (AN), where a greater AN signifies stronger Lewis acidity. The AN of 1A is 13.5 in C6D610b and performing the analogous study with 1B gave an AN value of 15.3. Methyl substitution at the peripheral boron vertices have an inductive effect, in this case acting as electron-withdrawing groups.3k, 18 Subsequent Gutmann-Beckett studies of 1BMe8 corroborated this hypothesis with an AN of 20.3, aligning with an increase of Lewis acidity at the boron center.

In summary, we have taken advantage of the lability of the C-H bonds of 1,1’-bis(o-carborane) to access 9-borafluorene analogues with a three-dimensional backbone. These species represent the first examples of 1,1’-bis(carboranyl)boranes and feature a highly planar central ring with enhanced Lewis acidity in comparison to 9-borafluorenes. Methyl substitution at the 8,9,10,12-B-vertices results in an increase of the overall Lewis acidity of the molecule. The results demonstrate the potential of utilizing bis(o-carboranes) as biphenyl analogues to create unique boracyclic architectures.

Supplementary Material

1

Acknowledgments

‡ C. D. M. and S. Y. are grateful to the Welch Foundation (Grant No. AA-1846) and the National Science Foundation for a CAREER Award (Award No. 1753025) for their generous support of this work. A. M. S. is thankful to 3M for a Non-Tenured Faculty Award, Alfred P. Sloan Foundation for a Fellowship in Chemistry, NIGMS(R35GM124746), and Research Corporation for Science Advancement (RCSA) for a Cottrell Fellowship.

Footnotes

Electronic Supplementary Information (ESI) available: Experimental procedures, multinuclear NMR Spectra, X-ray crystallographic data (CCDC 1884761–1884762), FT-IR spectra, and computational details (PDF). See DOI: 10.1039/x0xx00000x

Conflicts of interest

There are no conflicts to declare.

Notes and references

  • 1.(a)Schleyer P. v. R. and Najafian K, Inorg. Chem, 1998, 37, 3454–3470; [DOI] [PubMed] [Google Scholar]; (b)King RB, Chem. Rev, 2001, 101, 1119–1152; [DOI] [PubMed] [Google Scholar]; (c)Chen Z and King RB, Chem. Rev, 2005, 105, 3613–3642. [DOI] [PubMed] [Google Scholar]
  • 2.(a)Plesek J, Chem. Rev, 1992, 92, 269–278; [Google Scholar]; (b)Fox MA and Hughes AK, Coord. Chem. Rev, 2004, 248, 457–476; [Google Scholar]; (c)Grimes RN, Dalton Trans, 2015, 44, 5939–5956; [DOI] [PubMed] [Google Scholar]; (d)Grimes RN, Carboranes, Academic Press, New York, 2016; [Google Scholar]; (e)Estrada J, Lugo CA, McArthur SG and Lavallo V, Chem. Commun, 2016, 52, 1824–1826. [DOI] [PubMed] [Google Scholar]
  • 3.(a)Zakharkin LI and Kovredov AI, Bull. Acad. Sci. USSR, Div. Chem. Sci, 1973, 22, 1396–1396; [Google Scholar]; (b)Zakharkin LI and Shemyakin NF, Bull. Acad. Sci. USSR, Div. Chem. Sci, 1977, 26, 2184–2186; [Google Scholar]; (c)Zakharkin LI and Shemyakin NF, Bull. Acad. Sci. USSR, Div. Chem. Sci, 1978, 27, 1267–1268; [Google Scholar]; (d)Bokii NG, Yanovskii AI, Struchkov YT, Shemyakin NF and Zakharkin LI, Bull. Acad. Sci. USSR, Div. Chem. Sci, 1978, 27, 328–334; [Google Scholar]; (e)Antipin MY, Furmanova NG, Yanovskii AI and Struchkov YT, Bull. Acad. Sci. USSR, Div. Chem. Sci, 1978, 27, 1264–1267; [Google Scholar]; (f)Zakharkin LI and Shemyakin NF, Bull. Acad. Sci. USSR, Div. Chem. Sci, 1981, 30, 1525–1527; [Google Scholar]; (g)Zakharkin LI and Shemyakin NF, Bull. Acad. Sci. USSR, Div. Chem. Sci, 1984, 33, 2572–2573; [Google Scholar]; (h)Love RA and Bau R, J. Am. Chem. Soc, 1972, 94, 8274–8276; [Google Scholar]; (i)Harwell DE, McMillan J, Knobler CB and Hawthorne MF, Inorg. Chem, 1997, 36, 5951–5955; [DOI] [PubMed] [Google Scholar]; (j)Ren S and Xie Z, Organometallics, 2008, 27, 5167–5168; [Google Scholar]; (k)Puga AV, Teixidor F, Sillanpää R, Kivekäs R, Arca M, Barberà G and Viñas C, Chem. - Eur. J, 2009, 15, 9755–9763; [DOI] [PubMed] [Google Scholar]; (l)Qiu Z, Ren S and Xie Z, Acc. Chem. Res, 2011, 44, 299–309; [DOI] [PubMed] [Google Scholar]; (m)Spokoyny AM, Machan CW, Clingerman DJ, Rosen MS, Wiester MJ, Kenn e dy RD, Stern CL, Sarjeant AA and Mirkin CA, Nat. Chem, 2011, 3, 590–596; [DOI] [PubMed] [Google Scholar]; (n)Ren S, Qiu Z and Xie Z, J. Am. Chem. Soc, 2012, 134, 3242–3254; [DOI] [PubMed] [Google Scholar]; (o)Asay M, Kefalidis CE, Estrada J, Weinberger DS, Wright J, Moore CE, Rheingold AL, Maron L and Lavallo V, Angew. Chem. Int. Ed, 2013, 125, 11774–11777; [DOI] [PubMed] [Google Scholar]; (p)Chan AL, Fajardo J, Wright JH, Asay M and Lavallo V, Inorg. Chem, 2013, 52, 12308–12310; [DOI] [PubMed] [Google Scholar]; (q)Wright JH, Kefalidis CE, Tham FS, Maron L and Lavallo V, Inorg. Chem, 2013, 52, 6223–6229; [DOI] [PubMed] [Google Scholar]; (r)Yao Z-J, Zhang Y-Y and Jin G-X, J. Organomet. Chem, 2015, 798, 274–277; [Google Scholar]; (s)Lugo CA, Moore CE, Rheingold AL and Lavallo V, Inorg. Chem, 2015, 54, 2094–2096; [DOI] [PubMed] [Google Scholar]; (t)Wong YO, Smith MD and Peryshkov DV, Chem. - Eur. J, 2016, 22, 6764–6767; [DOI] [PubMed] [Google Scholar]; (u)Zhao D, Zhang J, Lin Z and Xie Z, Chem. Commun, 2016, 52, 9992–9995; [DOI] [PubMed] [Google Scholar]; (v)Wong YO, Smith MD and Peryshkov DV, Chem. Commun, 2016, 52, 12710–12713; [DOI] [PubMed] [Google Scholar]; (w)Axtell JC, Kirlikovali KO, Djurovich PI, Jung D, Nguyen VT, Munekiyo B, Royappa AT, Rheingold AL and Spokoyny AM, J. Am. Chem. Soc, 2016, 138, 15758–15765; [DOI] [PubMed] [Google Scholar]; (x)Dziedzic RM, Martin JL, Axtell JC, Saleh LMA, Ong T-C, Yang Y-F, Messina MS, Rheingold AL, Houk KN and Spokoyny AM, J. Am. Chem. Soc, 2017, 139, 7729–7732; [DOI] [PubMed] [Google Scholar]; (y)Chan TL and Xie Z, Chem. Sci, 2018, 9, 2284–2289; [DOI] [PMC free article] [PubMed] [Google Scholar]; (z)Cheng R, Li B, Wu J, Zhang J, Qiu Z, Tang W, You S-L, Tang Y and Xie Z, J. Am. Chem. Soc, 2018, 140, 4508–4511; [DOI] [PubMed] [Google Scholar]; (aa)Quan Y, Qiu Z and Xie Z, Chem. - Eur. J, 2018, 24, 2795–2805; [DOI] [PubMed] [Google Scholar]; (ab)Duttwyler S, Pure Appl. Chem, 2018, 90, 733. [Google Scholar]
  • 4.(a)Wiesboeck RA and Hawthorne MF, J. Am. Chem. Soc, 1964, 86, 1642–1643; [Google Scholar]; (b)Spokoyny AM, Pure Appl. Chem, 2013, 85, 903–919; [DOI] [PMC free article] [PubMed] [Google Scholar]; (c)Thiripuranathar G, Man WY, Palmero C, Chan APY, Leube BT, Ellis D, McKay D, Macgregor SA, Jourdan L, Rosair GM and Welch AJ, Dalton Trans, 2015, 44, 5628–5637; [DOI] [PubMed] [Google Scholar]; (d)Martin MJ, Man WY, Rosair GM and Welch AJ, J. Organomet. Chem, 2015, 798, 36–40; [Google Scholar]; (e)Riley LE, Chan APY, Taylor J, Man WY, Ellis D, Rosair GM, Welch AJ and Sivaev IB, Dalton Trans, 2016, 45, 1127–1137; [DOI] [PubMed] [Google Scholar]; (f)Kirlikovali KO, Axtell JC, Gonzalez A, Phung AC, Khan SI and Spokoyny AM, Chem. Sci, 2016, 7, 5132–5138; [DOI] [PMC free article] [PubMed] [Google Scholar]; (g)Sivaev I, Commun. Inorg. Synth, 2016, 21–28; [Google Scholar]; (h)Riley LE, Krämer T, McMullin CL, Ellis D, Rosair GM, Sivaev IB and Welch AJ, Dalton Trans, 2017, 46, 5218–5228; [DOI] [PubMed] [Google Scholar]; (i)Thiripuranathar G, Chan APY, Mandal D, Man WY, Argentari M, Rosair GM and Welch AJ, Dalton Trans, 2017, 46, 1811–1821; [DOI] [PubMed] [Google Scholar]; (j)Axtell JC, Kirlikovali KO, Dziedzic RM, Gembicky M, Rheingold AL and Spokoyny AM, Eur. J. Inorg. Chem, 2017, 4411–4416; [Google Scholar]; (k)Kirlikovali KO, Axtell JC, Anderson K, Djurovich PI, Rheingold AL and Spokoyny AM, Organometallics, 2018, 37, 3122–3131; [Google Scholar]; (l)Chan APY, Rosair GM and Welch AJ, Inorg. Chem, 2018, 57, 8002–8011. [DOI] [PubMed] [Google Scholar]
  • 5.(a)Scholz M and Hey-Hawkins E, Chem. Rev, 2011, 111, 7035–7062; [DOI] [PubMed] [Google Scholar]; (b)Issa F, Kassiou M and Rendina LM, Chem. Rev, 2011, 111, 5701–5722; [DOI] [PubMed] [Google Scholar]; (c)Zhu L, Lv W, Liu S, Yan H, Zhao Q and Huang W, Chem. Commun, 2013, 49, 10638–10640; [DOI] [PubMed] [Google Scholar]; (d)Lesnikowski ZJ, J. Med. Chem, 2016, 59, 7738–7758; [DOI] [PubMed] [Google Scholar]; (e)Dash BP, Satapathy R, Gaillard ER, Norton KM, Maguire JA, Chug N and Hosmane NS, Inorg. Chem, 2011, 50, 5485–5493; [DOI] [PubMed] [Google Scholar]; (f)McArthur SG, Geng L, Guo J and Lavallo V, Inorg. Chem. Front, 2015, 2, 1101–1104; [Google Scholar]; (g)Mukherjee S and Thilagar P, Chem. Commun, 2016, 52, 1070–1093; [DOI] [PubMed] [Google Scholar]; (h)Núñez R, Tarrés M, Ferrer-Ugalde A, de Biani FF and Teixidor F, Chem. Rev, 2016, 116, 14307–14378; [DOI] [PubMed] [Google Scholar]; (i)Li X, Yan H and Zhao Q, Chem. -Eur. J, 2016, 22, 1888–1898; [DOI] [PubMed] [Google Scholar]; (j)Tu D, Shao D, Yan H and Lu C, Chem. Commun, 2016, 52, 14326–14329. [DOI] [PubMed] [Google Scholar]
  • 6.(a)Muhammad S, Janjua MR and Su Z, J. Phys. Chem. C, 2009, 113, 12551–12557; [Google Scholar]; (b)Adams IA and Rupar PA, Macromol. Rapid Commun, 2015, 36, 1336–1340; [DOI] [PubMed] [Google Scholar]; (c)Matsumoto T, Takamine H, Tanaka K and Chujo Y, Mater. Chem. Front, 2017, 1, 2368–2375. [Google Scholar]
  • 7.(a)Biswas S, Maichle-Mössmer C and Bettinger HF, Chem. Commun, 2012, 48, 4564–4566; [DOI] [PubMed] [Google Scholar]; (b)Müller M, Maichle-Mössmer C and Bettinger HF, Angew. Chem. Int. Ed, 2014, 53, 9380–9383; [DOI] [PubMed] [Google Scholar]; (c)Yruegas S, Martinez JJ and Martin CD, Chem. Commun, 2018, 54, 6808–6811; [DOI] [PubMed] [Google Scholar]; (d)Yruegas S, Barnard JH, Al-Furaiji K, Dutton JL, Wilson DJD and Martin CD, Organometallics, 2018, 37, 1515–1518; [Google Scholar]; (e)Bluer KR, Laperriere LE, Pujol A, Yruegas S, Adiraju VAK and Martin CD, Organometallics, 2018, 37, 2917–2927; [Google Scholar]; (f)von Grotthuss E, John A, Kaese T and Wagner M, Asian J. Org. Chem, 2018, 7, 37–53; [Google Scholar]; (g)Zhang W, Li G, Xu L, Zhuo Y, Wan W, Yan N and He G, Chem. Sci, 2018, 9, 4444–4450; [DOI] [PMC free article] [PubMed] [Google Scholar]; (h)Radtke J, Mellerup SK, Bolte M, Lerner H-W, Wang S and Wagner M, Org. Lett, 2018, 20, 3966–3970. [DOI] [PubMed] [Google Scholar]
  • 8.(a)Entwistle CD and Marder TB, Chem. Mater, 2004, 16, 4574–4585; [Google Scholar]; (b)Thanthiriwatte KS and Gwaltney SR, J. Phys. Chem. A, 2006, 110, 2434–2439; [DOI] [PMC free article] [PubMed] [Google Scholar]; (c)Yamaguchi S and Wakamiya A, Pure Appl. Chem, 2006, 78, 1413. [Google Scholar]
  • 9.Lorenz-Rothe M, Schellhammer KS, Jägeler-Hoheisel T, Meerheim R, Kraner S, Hein MP, Schünemann C, Tietze ML, Hummert M, Ortmann F, Cuniberti G, Körner C and Leo K, Adv. Electron. Mater, 2016, 2, 1600152–1600163. [Google Scholar]
  • 10.(a)Romero PE, Piers WE, Decker SA, Chau D, Woo TK and Parvez M, Organometallics, 2003, 22, 1266–1274; [Google Scholar]; (b)Smith MF, Cassidy SJ, Adams IA, Vasiliu M, Gerlach DL, Dixon DA and Rupar PA, Organometallics, 2016, 35, 3182–3191; [Google Scholar]; (c)Zhang W, Yu D, Wang Z, Zhang B, Xu L, Li G, Yan N, Rivard E and He G, Org. Lett, 2018, DOI: 10.1021/acs.orglett.8b03538. [DOI] [PubMed] [Google Scholar]
  • 11.Reiner JR, Alexander RP and Schroeder H, Inorg. Chem, 1966, 5, 1460–1462. [Google Scholar]
  • 12.Popescu A-R, Musteti AD, Ferrer-Ugalde A, Viñas C, Núñez R and Teixidor F, Chem. - Eur. J, 2012, 18, 3174–3184. [DOI] [PubMed] [Google Scholar]
  • 13.Teixidor F, Vinas C and Rudolph RW, Inorg. Chem, 1986, 25, 3339–3345. [Google Scholar]
  • 14.Herzog A, Maderna A, Harakas GN, Knobler CB and Hawthorne MF, Chem. - Eur. J, 1999, 5, 1212–1217. [Google Scholar]
  • 15.(a)Araki T, Wakamiya A, Mori K and Yamaguchi S, Chem. - Asian J, 2012, 7, 1594–1603; [DOI] [PubMed] [Google Scholar]; (b)Ge F, Kehr G, Daniliuc CG, Mück-Lichtenfeld C and Erker G, Organometallics, 2015, 34, 4205–4208. [Google Scholar]
  • 16.Paetzold P, Pure Appl. Chem, 1991, 63, 345–350. [Google Scholar]
  • 17.(a)Beckett MA, Strickland GC, Holland JR and Sukumar Varma K, Polymer, 1996, 37, 4629–4631; [Google Scholar]; (b)Sivaev IB and Bregadze VI, Coord. Chem. Rev, 2014, 270–271, 75–88. [Google Scholar]
  • 18.(a)Heřmánek S, Chem. Rev, 1992, 92, 325–362; [Google Scholar]; (b)Heřmánek S, Inorganica Chim. Acta, 1999, 289, 20–44; [Google Scholar]; (c)Teixidor F, Barberà G, Vaca A, Kivekäs R, Sillanpää R, Oliva J and Viñas C, J. Am. Chem. Soc, 2005, 127, 10158–10159. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

1

RESOURCES