Author Manuscript

Title: L3C3P3: Tricarbontriphosphide Cage Radicals and Cations Stabilized by Cy- clic (alkyl)(amino)carbenes

Authors: Hansjorg¨ Grutzmacher,¨ Prof.; Zhongshu Li; Yuanfeng Hou; Yaqi Li; Alexan- der Hinz; Jeffrey Harmer; Chen-Yong Su; Guy Bertrand

This is the author manuscript accepted for publication and has undergone full peer review but has not been through the copyediting, typesetting, pagination and proofrea- ding process, which may lead to differences between this version and the Version of Record.

To be cited as: 10.1002/ange.201710099

Link to VoR: https://doi.org/10.1002/ange.201710099 L3C3P3: Tricarbontriphosphide Cage Radicals and Cations Stabilized by Cyclic (alkyl)(amino)carbenes

Zhongshu Li,[a] Yuanfeng Hou,[a] Yaqi Li,[a] Alexander Hinz,[b] Jeffrey R. Harmer,*[c] Cheng-Yong Su,[a] Guy Bertrand,[e] and Hansjörg Grützmacher*[a,d]

[a] Dr. Z. Li, Y. Hou, Y. Li, Prof. Dr. C.-Y. Su, Lehn Institute of Functional Materials (LIFM) School of Sun Yat-Sen University 510275 Guangzhou, China E-mail: [email protected] [b] Dr. A. Hinz University of Oxford, Chemistry Research Laboratory 12 Mansfield Road, OX1 3TA, Oxford, UK E-mail: [email protected] [c] Assoc. Prof. Dr. J. R. Harmer. Centre for Advanced Imaging, University of Queensland, Brisbane, QLD, 4072, Australia E-mail: [email protected] [d] Prof. Dr. H. Grützmacher Department of Chemistry and Applied Biosciences ETH Zurich 8093 Zurich, Switzerland E-mail: [email protected] [e] Prof. Dr. G. Bertrand UCSD/CNRS Joint Research Chemistry Laboratory Department of Chemistry University of California San Diego La Jolla, CA 92521–0403, USA E-mail: [email protected]

1

This article is protected by copyright. All rights reserved Dedicated to Prof. Dr. Dieter Fenske on the oaccasion of his 75 th birthday.

Abstract: usually oligomerize to give rings with a conjugated -electron system. In contrast, phosphaalkynes, R-CP, frequently give compounds with polycyclic structures which are thermodynamically more stable than the corresponding

-conjugated isomers. Here we report the syntheses of the first C3P3 cages with either radical or cation ground states stabilized by cyclic (alkyl)(amino)carbenes (CAACs). These compounds may be considered as examples of tricarbontriphosphide coordinated by carbenes and are likely formed via trimerization of the corresponding mono-radicals CAAC-CP• . The mechanism for the formation of these tricarbontriphosphide cage radicals has been rationalized by a combination of experiments and DFT calculations.

Organo- compounds with CnPn cage skeletons that are exclusively composed of and phosphorus atoms are still relatively rare and their properties remain to be explored. To the best of our knowledge, all reported CnPn cages are prepared by oligomerization of phosphaalkynes, mainly tBu-CP.[1] For instance, phosphaalkyne tetramers A and B (Scheme 1) could be prepared via thermally induced or metal mediated oligomerization.[2,3] Other cage types of tetramers, pentamers, and even hexamers are also obtained from the corresponding

[1] phosphaalkynes. Remarkably, stable cage skeletons of the composition C3P3 seemingly have never been isolated while 1,3,5-triphosphabenzenes C and

1,3,5-triphospha-Dewar-benzenes D are well established and their conversion into cage compounds with additional reagents is documented.[1,3b] This agrees well with calculations on (H,C,P)3 isomers, which predict that the cyclic triphosphabenzenes are more than 30 kcal mol–1 lower in energy than prismane cages.[4]

2

This article is protected by copyright. All rights reserved

Scheme 1. Possible structures for tetra- or trimers of RCP (R = H, organyl groups).

Singlet N-heterocyclic carbenes (NHCs)[5] and cyclic diamidocarbenes (DACs)[6] may stabilize otherwise reactive molecules and recently cyclic and acyclic C2P2 skeletons as possible forms of “dicarbondiphosphide” could be isolated with these as substituents.[7] Here we report the synthesis of neutral tricarbontriphosphide radicals

 + (L3C3P3) and their oxidation to the corresponding cations (L3C3P3) (L = CAAC). Phosphaketenes like 1 are easily prepared from chlorodiazaphospholenes and Na(OCP)[7] and are quantitatively rearranged to phosphallenes

L=C=P[PO(NDipp)2(CH)2] upon reaction with diisopropylphenyl (Dipp) substituted

[7a,8] NHC or DAC as carbene L. These phosphaallenes cleanly react with KC8 to give the anionic oxydiazaphospholene 3 and cyclic or acyclic dicarbondiphosphide

[8]  compounds L2C2P2 (L = NHC, DAC). Carbene-bound CP radicals L=C=P 4 may be formed as first intermediates. We reasoned that with sterically less demanding carbenes higher oligomers of these radicals may be obtained. Cyclic (alkyl)(amino)carbenes, CAACs, can be easily prepared on a multi-gram scale and their steric demand can be facilely tuned.[9] Phosphaketene 1[8] was reacted with

CAACs L1 – L5 (Scheme 2). All reactions yield phosphaallenes 2 as bright pink powders in good to excellent yields which show two doublets in the 31P NMR spectra

1 at δ = 99.1 ppm (averaged) and δ = 17.5 ppm (averaged) with a coupling constant JP– 13 P = 398.3 Hz (averaged). In the C NMR spectra, doublet resonances in the range

1 from δ = 273.1 ppm (2a) to δ = 282.8 ppm (2e) with JP–C coupling constants of about 20 Hz are observed for the carbon nucleus bonded directly to CAACs which is in between that of structurally similar phosphaallenes stabilized with NHC (289.5 ppm)[7a] or DAC (249.8 ppm).[8] This ordering of the frequencies from high

3

This article is protected by copyright. All rights reserved to low values reflects the π-electron acceptor strength of carbenes which increases in the order NHC < CAAC < DAC.[9] The structures of 2b and 2e were determined by single crystal X-ray diffraction analysis. As an example, the structure of 2b is shown in Figure 1a (see the Supporting Information for further details on all the reported structures herein). These molecules contain a slightly bent C2-C1-P1 unit (average angle at C1 is 172.0) with contracted C=C (C1–C2 1.326 Å, averaged) and C=P (C1– P1 1.627 Å, averaged) double bonds as compared to typical double bonds (C=C 1.34

Å; C=P 1.69 Å).[10]

Scheme 2. Synthesis of 2, 3, 5, and 6.

Compounds 2a – c react cleanly with one equivalent of KC8 in tetrahydrofuran (THF)

4

This article is protected by copyright. All rights reserved to give compound 3. Further products were not detected by 31P NMR spectroscopy but the reaction solutions showed very strong EPR signals (vide infra). After work-up

 the neutral tricarbontriphosphide radicals (L3C3P3) 5a – c were isolated in excellent yields (> 82%) as extremely air-sensitive green powders. The isolation of these cage radicals is a very strong indication for the formation of intermediate L=C=P• radicals 4. The molecular structures of 5a – c were determined unambiguously by single-crystal X-ray diffraction methods. As an example the structure of 5b is shown in Figure 1b (all substituents on the CAAC ring were omitted for clarity). a b

c d

Fig 1. Plots of the molecular structure of 2b, 5b, 6e, and 7b. Ellipsoids are set to 50% probability; H atoms, solvent molecules, anion, and substituents on CAAC for 5b and 7b are omitted for clarity. Selected distances [Å] and angles [°]: a) 2b: P1-P2 2.2360(6), P1-C1 1.623(2), C1-C2 1.324(3), P2-P1-C1 99.79(8), P1-C1-C2 173.18(16); b) 5b: P1-P2 2.6846(6), P1-P3 2.2082(6), P1-C1 1.8494(18), C1-P2

5

This article is protected by copyright. All rights reserved 1.8577(18), P2-C3 1.8578(18), C3-P1 1.8412(18), P2-C5 1.8828(18), P3-C5 1.8040(18), C1-C2 1.357(2), C3-C4 1.352(2), C5-C6 1.386(2), C5-P3-P1 94.40(6), P3-P1-C3 94.15(6), P3-P1-C1 91.61(6); c) 6e: P1-P1' 2.2508(8), P1-C1 1.6336(17), C1-C2 1.324(2), P1'-P1-C1 99.30(7), P1-C1-C2 175.67(14); d) 7b: P1-P2 2.755, P1-P3 2.2252(7), P1-C1 1.939(2), C1-P2 1.941(2), P2-C3 1.837(2), C3-P1 1.910(2), P2-C5 1.856(2), P3-C1 1.808(2), P3-C5 1.821(2), C1-C2 1.443(3), C3-C4 1.365(3), C5-C6 1.371(3), C5-P3-P1 84.05(9), C5-P3-C1 84.05(9), P3-P1-C3 99.22(7), P3-P1-C1 54.09(6).

The structural parameters of all three compounds 5a – c vary only marginally (see the

Supp. Info for details). The center of the molecules comprises a P3C3 cage which is bonded to three CAAC substituents. Within the cage, all C–P bonds are in the range of single C–P bonds (P–C 1.86 Å) with the exception of the C5-P3 bond which is slightly shorter (1.83 Å, averaged).[10] Likewise, all C=C bonds between the CAAC

[10] groups and the C3P2 cage are in the range of double bonds (C=C 1.34 Å) with the exception of the C5=C6 bonds which are slightly longer (1.37 Å, averaged). These data indicate significant -electron conjugation across the P3–C5–C6 unit (vide infra). The P3–P1 bonds (2.22 Å, averaged) correspond to standard P-P single bonds bond lengths. The structures of 5a – c can be viewed as [4+2] cycloadducts between a C2P2

• ring in L2C2P2 and a L=C=P radical where the addition took place across the P1, P2 vector. Note that also in reactions between L2C2P2 and H2, a cis-specific addition to give L2C2(PH)2 took place at the phosphorus centers which are therefore considered to be the reactive centers of carbene stabilized dicarbondiphosphides.[8] These additions cause the basal C2P2 heterocycles to deviate from planarity as seen in L2C2P2 and dihedral angles of 38.9° (averaged) are observed in 5a – 5c.

6

This article is protected by copyright. All rights reserved

Fig 2. a) Continuous wave (CW) EPR spectrum of 5c at 100 K; b) Spin density (density = 0.004) of 5cM; c) CW EPR spectrum of 8c at 100 K; d) Spin density (density = 0.004) of 8cM.

The room-temperature EPR spectra of a hexane solution of 5a – c all display a

31 doublet due to a large isotropic P hyperfine interaction (aiso  187 MHz) which indicates a large percentage of the spin density is localized on one of the phosphorus nuclei. To determine the anisotropic coupling constants, the corresponding frozen-solution CW EPR spectra in hexane were recorded, as shown in Fig. 2a for the example of 5c. The main features of this spectrum result from the largest 31P hyperfine interaction with one large (features at 326, 348 mT) and two small (features around 335 mT) principal values, and a small g-value anisotropy. Simulations yielded the following principal g values and 31P hyperfine couplings for 5a: g = [2.0027, 2.0071, 2.0151], A(31P3) = [–19, –40, 618] MHz; 5b: g = [2.0022, 2.0073, 2.0154], A(31P3) = [–28, –35, 623] MHz; and for 5c: g = [2.0025, 2.0068, 2.0140], A(31P3) = [– 28, –37, 626] MHz (Table S11). The large anisotropic 31P hyperfine coupling confirms that a majority of the unpaired spin density is localized in a phosphorus p-type orbital at P3. The fine splittings resolved on the central signal (around 335 mT) indicate that the radical’s SOMO is delocalized onto the nitrogen nuclei in the CAAC groups. 7

This article is protected by copyright. All rights reserved Indeed, our spin Hamiltonian model contains small 14N hyperfine couplings for N3 of 11 , 12 and 9 MHz for 5a, 5b and 5c, respectively. Such small 14N hyperfine couplings could also be observed for a CAAC supported P2 radical, but not for the NHC

[11] supported P2 or P3 radicals. Thus, in agreement with the structure data, the single occupied molecular orbital (SOMO) is a conjugated π–system composed of the P3– C5 bond connected to the co-planar CAAC bonded to C5. This is consistent with the large anisotropy of the hyperfine coupling (P3) in particular with a large p-type orbital contribution to the SOMO. The couplings from EPR provide an estimation of the spin population ρ = 0.73–0.74 e on P3. These values agree well with DFT results computed from a simplified model 5M which give ρ = 0.76 e on P3 (see Fig 2b for a plot of the spin density).

The compounds 2d, e bearing bulkier carbenes react with one equivalent of KC8 in THF to afford compound 3 and the corresponding CAAC stabilized acyclic

31 diamagnetic linear C2P2 compounds 6d, e as green powders. In the P NMR spectra, two singlets in the range between  = 150 – 160 ppm were observed which we assign to cis- and trans isomers with respect of the nitrogen centers in the CAAC moieties in the two P=C=CAAC units. Indeed, for 6d (Fig S26) the cis-isomer and for 6e the trans-isomer (Fig. 1c) were crystallized from the mixture of products and the structures determined by X-ray diffraction methods (see the Supp. Info for details).

The structural parameters in 6e (C1=C2 1.32 Å); C1=P1 1.63 Å); P1–P1' (2.25 Å) are

[8] comparable to the DAC stabilized acyclic C2P2 core.

Scheme 3. Synthesis of 7 and 8.

8

This article is protected by copyright. All rights reserved The cyclic neutral and diamagnetic dicarbondiphosphides L2C2P2 (L = NHC) are

•+ [8] easily oxidized (E1/2=‒0.451 V) to give stable cyclic radical cations [L2C2P2] . In contrast, cyclic voltammograms of the paramagnetic neutral radical cages 5a – c show

+ – no reversible redox waves but are chemically oxidized upon reaction with Fc PF6 or

+ – ‒ Fc BARF as oxidants (BARF = [B(3,5-(CF3)2C6H3)4] ). Crystalline yellow to orange cations 7a – c were isolated as major products (Scheme 3) after rapid work up. All structures were determined by X-ray diffraction methods and as an example, the structure of 7b is shown in Figure 1d (all substituents on the CAAC ring were omitted for clarity). The structural parameters of all three compounds 7a – c vary only marginally (see the SI for more details). Upon oxidation a new bond between P3 and C1 is formed which converts the bicyclic structure of 5a – c into tricyclic structures

7a – c with acute P3–P1–C1 angles (54.1°, averaged). Notably, this newly formed P3– C1 bond is significantly shorter (1.81 Å) than the other two, P1–C1 and P2–C1, which are 1.94 Å long. All other P–C bonds are in the range of 1.80 to 1.86 Å as observed in the neutral radical cages 5a – c. As expected, the C–C bond from C1 in a tetrahedral coordination sphere is longer (C1–C2 1.43 Å) than the ones at C3 and C5 which are in a trigonal planar coordination sphere (1.37 Å). Therefore, the C1–C2 can be assumed to be a single bond and the CAAC substituent bound to C1 formally carries the positive charge as shown in Scheme 3. This assumption is supported by the fact that the N1–C2 bond (1.32 Å, average) is longer than a typical C=N double bond (1.27 Å[10]) but significantly shorter than the corresponding C–N bonds in the neutral radicals 5a – c (1.40 Å, average). Despite the fact that 7a – c are expected to be diamagnetic, the NMR spectra are broad and ill defined. Experimentally, a weak CW EPR signal was observed from isolated 7c. Both room temperature and frozen solution CW EPR spectra indicate the formation of a CAAC stabilized cyclic C2P2 cation radical, as compared to the reported NHC analog.[8] When 5a – c were reacted with 1.5 equivalents of Fc+BARF– in diethyl ether for more than 10 hours at room temperature, red oily precipitates formed after evaporation of the solvent and washing with hexane. These solidified under vacuum to yield red powders which could not be further purified. The frozen 9

This article is protected by copyright. All rights reserved solution (100 K) CW EPR spectra for 8a, c in toluene display a strong signal and a characteristic spectrum is shown in Fig. 2c for the example of 8c. This spectrum shows a splitting pattern described by two equivalent 31P hyperfine couplings with two small and one large principal value and a small g-value anisotropy. Because of the weak CW EPR signal intensity of 8b, we are not sure about its structure and it is not discussed further. Simulations yielded for 8a: g = [2.0020, 2.0062, 2.0098], 2 × A(31P3) = [505, –5, –11] MHz; and for 8c: g = [2.0022, 2.0068, 2.0097],

2 × A(31P3) = [507, –4, –5] MHz (Table S11). The 31P hyperfine coupling provides an estimate of the spin population ρ = 0.53 e on each phosphorus atom, which compares well to the DFT value where ρ = 0.59 e (Fig 2d). These spin Hamiltonian parameters are very similar to our previously reported NHC stabilized cyclic C2P2 radical cation.[8] The Supporting Information contains further details on all the reported EPR analysis herein including room temperature data. We cannot devise with certainty a mechanism for the formation of the C2P2 radical cations 8, but we assume that the neutral C3P3 radical cages 5 are first oxidized to their corresponding cations which are further oxidatively decomposed as indicated in Scheme 3. To gain further insight into a possible reaction mechanism for the formation of the cage radical 5, a minimum-energy reaction pathway (MERP) with simplified model compounds was calculated with DFT (BP86/def2-TZVPP) (Scheme 4).[12] The dimerization of radical 4M to give 6M proceeds via a barrierless exothermic reaction (ΔG = –12.9 kcal mol–1). This is in accord with the experimental findings and isolation of 6. An alternative reaction path was searched in order to explain the formation of cyclic C2P2 compounds and an zwitterionic intermediate I was found in a local minimum as the product of a slightly endothermic dimerization (ΔG = 4.0 kcal mol–1).

10

This article is protected by copyright. All rights reserved

Scheme 4. Reaction MERP of the reaction coordinates for the formation of trimer radical simplified model.

The conversion of I to II requires passing an activation energy of Ea = 15.2 kcal mol–1

–1 at TS 1 to give the cyclic C2P2 isomer II, which is 16.0 kcal mol more stable than

M –1 M linear C2P2 6 , in an exothermic reaction (ΔG = ‒28.9 kcal mol ). A third radical 4 may then attack one of the P atoms of the cyclic C2P2 to give intermediate III (ΔG = –

–1 M 36.4 kcal mol ). In the final step, the C3P3 cage 5 is formed in a ring closing reaction with a very small activation barrier of only 5.5 kcal mol–1. Thus, the trimerisation of L=C=P• radicals is an overall exothermic reaction (ΔG = –51.0 kcal mol–1) with smooth activation barriers along the MERP.

The results reported here support the assumption that the reduction of phosphaallenes

• L=C=P[PO(NR)2CR’)2] lead to carbene bound CP radicals, L=C=P , which may either dimerize to rhombic or linear L2C2P2 dimers or trimerize to cage radicals

• (L3C3P3) depending on the electronic and steric properties of the carbenes L. This

11

This article is protected by copyright. All rights reserved implies that eventually also the monomer, L=C=P• is assessable with highly bulky carbenes.[7c, 13]

Acknowledgements: This work was supported by the National Natural Science Foundation of China (21603280, 21720102007), the Fundametnal Research Funds for the Central Universities (171gpy78), the ETH Zürich, the STP Project of Guangzhou (201504010031), and the NSF of Guangdong Province (S2013030013474).

References

[1] See the selected recent reports and lit. cited therein: a) R. L. Falconer, C. A. Russell, Coord. Chem. Rev. 2015, 297–298, 146–167; b) Chirila, R. Wolf, J. C. Slootweg, K. Lammertsma, Coord. Chem. Rev. 2014, 270–271, 57–74. c) M. Trincado,

A. J. Rosenthal, M. Vogt, H. Grützmacher, Eur. J. Inorg. Chem. 2014, 10, 1599-1604; d) L. E. Longobardi, C. A. Russell, M. Green, N. S. Townsend, K. Wang, A. J. Holmes, S. B. Duckett, J. E. McGrady, D. W. Stephan, J. Am. Chem. Soc. 2014, 136, 13453–13457; e) F. Mathey in Phosphorus-Carbon Heterocyclic Chemistry: The Rise of a New Domain. Elsevier, London, 2001. [2] a) T. Wettling. B. Geissler, R. Schneider, S. Barth, P. Binger, M. Regitz, Angew. Chem. Int. Ed. Engl., 1992, 31, 758–759; Angew. Chem., 1992, 104, 761–762; b) T. Wettling, J. Schneider, O. Wagner, C. G. Kreiter, M. Regitz, Angew. Chem. Int. Ed.

1989, 28, 1013–1014; Angew. Chem., 1989, 101, 1035–1037. [3] a) P. Binger, G. Glaser, B. Gabor, R. Mynott, Angew. Chem. Int. Ed. 1995, 34, 81–83; Angew. Chem. 1995, 107, 114–116; b) P. Binger, S. Leininger, J. Stannek, B. Gabor, R. Mynott, J. Bruckmann, C. Krüger, Angew. Chem. Int. Ed. 1995, 34, 2227–

2230; Angew. Chem. 1995, 107, 2411–2414. [4] a) T. R. Galeev, A. I. Boldyrev, Phys. Chem. Chem. Phys. 2011, 13, 20549–20556; b) M. Hofmann, P. v. R. Schleyer, M. Regitz, Eur. J. Org. Chem. 1999, 3291–3303. [5] Selected reviews: a) M. N. Hopkinson, C. Richter, M. Schedler, F. Glorius,

Nature 2014, 510, 485–496; b) Y. Wang, G. H. Robinson, Inorg. Chem. 2014, 53,

12

This article is protected by copyright. All rights reserved 11815–11832; c) C. D. Martin, M. Soleilhavoup, G. Bertrand, Chem. Sci. 2013, 4, 3020–3030; d) D. J. D. Wilson, J. L. Dutton, Chem. Eur. J. 2013, 19, 13626–13637; e) D. Martin, M. Soleilhavoup, G. Bertrand, Chem. Sci. 2011, 2, 389–399.

[6] J. P. Moerdyk, D. Schilter, C. W. Bielawski, Acc. Chem. Res. 2016, 49, 1458– 1468. [7] a) Z. Li, X. Chen, Z. Benkö, L. Liu, D. A. Ruiz, J. L. Peltier, G. Bertrand, C.-Y. Su, H. Grützmacher, Angew. Chem. Int. Ed. 2016, 55, 6018–6022; Angew. Chem.

2016, 128, 6122–6126; b) Z. Li, X. Chen, C.-Y. Su, H. Grützmacher, Chem. Commun. 2016, 52, 11343–11346; c) L. Liu, D. A. Ruiz, D. Munz, G. Bertrand, Chem 2016, 1, 147–153; d) M. M. Hansmann, G. Bertrand, J. Am. Chem. Soc. 2016, 138, 15885– 15888; e) M. M. Hansmann, R. Jazzar, G. Bertrand, J. Am. Chem. Soc. 2016, 138,

8356–8359. f) Z. Li. X. Chen, M. Bergeler, M. Reiher, C.-Y. Su, H. Grützmacher, Dalton Trans. 2015, 44, 6431–6438. [8] Z. Li, X. Chen, D. Andrada, G. Frenking, Z. Benkö, Y. Li, J. R. Harmer, C.-Y. Su, H. Grützmacher, Angew. Chem. Int. Ed. 2017, 56, 5744–5749; Angew. Chem. 2017,

129, 5838–5843. [9] a) See the review: M. Melaimi, R. Jazzar, M. Soleihavoup, G. Bertrand, Angew. Chem. Int. Ed. 2017, 56, 10046–10068; Angew. Chem. 2017, 129, 10180–10203; and literature cited therein. b) R. Jazzar, R. D. Dewhurst, J.-B. Bourg, B. Donnadieu, Y.

Canac, G. Bertrand, Angew. Chem. Int. Ed. 2007, 46, 2899–2902; Angew. Chem. 2007, 119, 2957–2960; c) W. Lavallo, G. D. Frey, S. Kousar, B. Donnadieu, G. Bertrand, Proc. Natl. Acad. Sci. USA. 2007, 104, 13569–13573; d) V. Lavallo, Y. Canac, C. Präsang, B. Donnadieu, G. Bertrand, Angew. Chem. Int. Ed. 2005, 44, 5705–5709;

Angew. Chem. 2005, 117, 5851–5855. [10]a) P. Pyykko, M. Atsumi, Chem. Eur. J. 2009, 15, 12770–12779; b) F. H. Allen, O. Kennard, D. G. Watson, L. Brammer, A. G. Orpen, R. Taylor, J. Chem. Soc. Perkin Trans. II 1987, S1–S19. [11] a) A. M. Tondreau, Z. Benkö, J. R. Harmer, H. Grützmacher, Chem. Sci. 2014, 5, 1545–1554; b) O. Back, B. Donnadieu, P. Parameswaran, G. Frenking, G. Bertrand, Nat. Chem. 2010, 2, 369–373. 13

This article is protected by copyright. All rights reserved [12] M. J. Frisch, et al., Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2009 (full reference see ESI†). [13] E. A. Romero, P. M. Olsen, R. Jazzar, M. Soleilhavoup, M. Gembicky, G.

Bertrand, Angew. Chem. Int. Ed. 2017, 56, 4024–4027.

14

This article is protected by copyright. All rights reserved