Direct Functionalization of White Phosphorus to

2 hours ago - Search; Citation; Subject. Search in: Anywhere, Title .... Download Citation · Email a Colleague ... Explore by: Author of this Article ...
0 downloads 0 Views 753KB Size
Subscriber access provided by UNIV OF LOUISIANA

Communication

Direct Functionalization of White Phosphorus to Cyclotetraphosphanes: Selective Formation of Four P-C Bonds Shanshan Du, Jimin Yang, Jingyuan Hu, Zhengqi Chai, Gen Luo, Yi Luo, Wen-Xiong Zhang, and Zhenfeng Xi J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b02628 • Publication Date (Web): 16 Apr 2019 Downloaded from http://pubs.acs.org on April 16, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Shanshan Du,†,§ Jimin Yang,‡,§ Jingyuan Hu,† Zhengqi Chai,† Gen Luo,‡ Yi Luo,*,‡ Wen-Xiong Zhang,*,† and Zhenfeng Xi† †

Beijing National Laboratory for Molecular Sciences (BNLMS), Key Laboratory of Bioorganic Chemistry and Molecular Engineering of Ministry of Education, College of Chemistry, Peking University, Beijing 100871, China ‡

State Key Laboratory of Fine Chemicals, School of Chemical Engineering, Dalian University of Technology, Dalian 116024, China Supporting Information Placeholder ABSTRACT: Converting elemental white phosphorus directly into organophosphorus or polyphosphorus is meaningful, challenging and attractive. The ate-complexes of aluminacyclopentadienes 1a,b react with P4 to afford selectively the cyclotetraphosphanes 2a,b featuring four newly formed P−C bonds and a planar square cyclo-P4 ring. DFT calculations show that the conversion of tetrahedral P 4 to planar cyclo-P4 moiety undergoes through an unexpected 1,1-P-insertion/D-A reaction/isomerization cascade process. The reaction of 2a with iodomethane or p-benzoquinone afforded the P-methylation product 3 and the metal-free cyclotetraphosphane 4, respectively. Interestingly, reduction of 4 generated the phospholyl anions 5 and 6 while treatment of 4 with iodomethane afforded the phospholyl cation 7.

The functionalization of white phosphorus (P4) through direct P−C bond formation has been a continuous interest because of the chlorine-free synthesis of organophosphorus compounds.1,2 Since the first report of P−C bond formation by reaction of organolithium with P4 in 1963,3 continuous efforts have been made in order to convert P4 into organophosphorus compounds directly.4-5 Although the conversions of P4 to various [MxPy]n complexes have gained great achievements,6-9 the direct formation of organophosphorus compounds from P4 are rare.10-13 In fact, making organophosphorus compounds through direct P−C bond formation from P4 generally suffers from: i) high electrophilic reactivity of the P4 tetrahedron, ii) the low selectivity for the P−P bond rupture after the first P−P bond cleavage, and iii) the low conversion efficiency of the phosphorus atoms in P4. Thus, the controllable and atom-efficient functionalization of P4 to construct directly organophosphorus or polyphosphorus compounds is highly desirable. Herein, we report an unprecedented metal-mediated reaction pattern of P4 (Scheme 1), in which four P−C bonds were selectively formed between aluminacyclopentadienes and P 4, affording a series of organometallic and organic cyclotetraphosphanes in excellent yields. These cyclotetraphosphanes are a class of important organic polyphosphanes which are not easy to access by other methods.14-16 DFT calculations show this P4 functionalization undergoes through an unexpected 1,1-P-insertion/D-A reaction/isomerization cascade process.

Scheme 1. Direct Functionalization of P4 to Cyclotetraphosphanes: Selective Formation of Four P−C Bonds

The ate-complexes of aluminacyclopentadienes 1a,b, which are readily accessed by the metathesis reaction of dilithio reagents and Et2AlCl,17 reacted with 0.5 equivalents of P4 in THF at 50 °C leading to the formation of compounds 2a,b as the sole product monitored by 31P NMR (Scheme 2). After recrystallization, 2a,b were obtained as air- and moisture-sensitive yellow solids in good yields. This reaction has a 100% atom economy, and all the atoms from both 1a,b and P4 were converted to the products.

Scheme 2. Synthesis of 2

The solid-state structures of 2a and 2b were determined by single-crystal X-ray diffraction analysis. 2a and 2b are isostructural and isomorphous, and only the structure of 2a is discussed here (Figure 1 for 2a, see SI for X-ray structure of 2b). 2a adopts a D2h symmetry with an inversion center and reveals an anti-arrangement of two 5,6-diphospha-7-aluminanorbornene moieties. The neutral cyclo-P4 ring is a perfect planar square, and almost perpendicular (105.2°and 105.7°) to the other two planes defined by P1−P2−C4 and P1′−P2′−C4′. The P1−P2−P1′−P2′ torsional angle is zero and the four P−P bond lengths are almost same (2.2597(11) Å and 2.2537(12) Å, respectively), which are longer than that (2.194 Å) in P4 but within the normal range of P−P single bonds.1 The bond

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Scheme 3. DFT Calculated Mechanism for the Formation of 2a2-

Figure 1. Solid-state structure of 2a. Ellipsoids are set at 30% probability level; hydrogen atoms, lithium atoms and coordinating THF molecules are omitted for clarity. Selected bond lengths [Å] and angles [°]: C1−C2 1.517(5), C2−C3 1.345(5), C3−C4 1.516(5), P1−P2 2.2597(11), P1−P2′ 2.2537 (12), P1−C1 1.858(3), P2−C4 1.866(4); P2−P1−P2′ 89.72(4), P1−P2−P1′ 90.28(4), P2′−P1−C1 105.19(11), P1′−P2−C4 105.71(12). angles are close to 90°(89.7°and 90.3°, respectively). Furthermore, DFT calculated MOs and WBI values indicate that the aluminum atoms don’t have interactions with the cyclo-P4 unit (see SI for the detailed bonding analysis). It is the first time that white phosphorus is converted to the all-alkyl-substituted cyclo-P4 rings, which are significantly different from the metal-cyclo-P4 complexes reported before.18 The DFT calculated mechanism for the formation of 2a2- is shown in Scheme 3. Both the decomposition of P 4 to P2 and the isomerization of P4 to tetraphosphacyclobutadiene in the absence of Al complex are energetically unfavorable (Scheme S1). Considering the energetically feasible formation of 1a- (anion species of 1a, Scheme S2), the reaction of 1a- with P4 was used as a model reaction in the calculations. It is noteworthy that P 4 could not coordinate to Al center of 1a-, due to the lack of coordination site. The addition of P4 to an Al−C(sp2) bond of 1a- could occur via

TSI to give INT1. Subsequently, INT1 isomerizes to INT1′ which contains a P=P double bond. Then, [4+2]-cycloaddition via D-A reaction forming two P−C bonds takes place between INT1′ and 1a- via TSIIa with an energy barrier of 21.5 kcal/mol, giving intermediate INT2a as a syn-chair conformation (See SI for other anti-boat, anti-chair, and syn-boat conformations, Scheme S3). Such [4+2]-cycloaddition between cyclo-P3 and diene was also documented before.5b,10a,19 Finally, the P2−C bond formation of INT2a together with cleavage of P2−P4 bond and formation of Al−C and P1−P4 bonds occurs to yield 2a2-. This step has an energy barrier of 24.4 kcal/mol. The whole process of white phosphorus activation is exergonic by 25.4 kcal/mol. This mechanism shows that both the Al−C bonds and butadiene moiety of 1a- facilitate the relatively controllable transformation of P 4 tetrahedron to the planar cyclo-P4 moiety. To convert 2a into metal-free organopolyphosphorus compounds, we treated 2a with excess iodomethane. A cyclodiphosphino-1,3-diphosphonium species 3 was produced exclusively instead of the preconceived product corresponding to cleavages of Al−C bonds (Scheme 4a). The solid-state structure of 3 was determined by single-crystal X-ray diffractions (Figure 2). The cyclo-P4 ring in 3 has different structural features from that in 2a.20 It is close to be planar with the P1−P2−P3−P4 torsional angle as 1.8°. Compared with that in 2a, the P1−P4 and P2−P3 bonds (average 2.2625 Å) are elongated, while the P1−P2 and P3−P4 bonds (average 2.2172 Å) are shortened. The P−P−P bond angles are divided into two acute angles (the tricoordinate phosphine center, average 83.7°) and two obtuse angles (the tetracoordinate phosphonium center, average 96.3°). The phosphonium-carbon bonds (P2−C4, P2−C17, P4−C18, P4−C34, average 1.7815 Å) are slightly shorter than the phosphine-carbon bonds (P1−C1, P3−C19, average 1.8170 Å). Interestingly, oxidation of 2a by p-benzoquinone in THF at room temperature could provide a metal-free cyclotetraphosphane 4 (Scheme 4b). The structure of 4 can be identified by NMR techniques and HRMS. Only one singlet is found in 31P NMR at 9.87 ppm indicating all the phosphorus atoms to be identical, which is also close to the value of 9 ppm for the tetraphenyl-substituted cyclotetraphosphane (C6H5P)4.14 The protons on the methyl groups are a triplet at 2.21 ppm with 4JPH at 3.1 Hz in 1H NMR. In 13C

ACS Paragon Plus Environment

Page 2 of 6

Page 3 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society NMR, each carbon of 4 splits into a triplet locating at 1.3, 18.8, 143.7, 156.8 ppm with 3JPC, 3JPC, 1JPC, 2JPC values at 2.3, 2.5, 19.0, 6.6 Hz, respectively (See Figure S33 in SI). For comparison, the methyl groups on 2a are a singlet in 1H NMR, and the carbon atoms connected with phosphorus atoms locate at 37.1-37.6 ppm as a multiple peak. The different NMR characteristics between 4 and 2a have proved the butadiene skeletons in 4. Besides, HRMS analysis (calcd. for C24H49P4Si4 [M+H]+: 573.1856, found 573.1851) performs an accurate molecular weight matching well with the molecular formula. Furthermore, by DFT calculations, the cyclo-P4 ring in 4 is clearly puckered (See Figure S42 in SI) upon releasing the metal.21-23

crystal X-ray diffractions (5 is in accord with previous report,13c see SI for X-ray structures of 6 and 7). It is worth mentioning that traditional synthetic methods of these phosphole derivatives usually take PCl3 or R2PCl as necessary reagents which are toxic and environmentally unfriendly.24 This work indicates that the polyphosphorus compounds derived from P 4 can be applied to synthesize small molecules like 5-7.

Scheme 5. Reactions of 4

Scheme 4. Synthesis of 3 and 4

In summary, we have presented the aluminacyclopentadienemediated selective and controllable functionalization of P 4 to a series of cyclotetraphosphane derivatives in high yields. This work shows a promising synthetic route to the cyclotetraphosphanes starting from P4 and throws a light on the functionalization of P 4 to organophosphorus compounds.

Figure 2. Solid-state structure of 3. Ellipsoids are set at 30% probability level; hydrogen atoms are omitted for clarity. Selected bond lengths [Å] and angles [°]: P1−P2 2.2206(17), P2−P3 2.2706(17), P3−P4 2.2138(17), P1−P4 2.2544(17), P2−C17 1.786(5), P4−C34 1.795(5), P1−C1 1.809(5), P2−C4 1.768(5), P3−C21 1.825(5), P4−C18 1.777(5); P2−P1−P4 83.82(6), P1−P2−P3 95.93(6), P2−P3−P4 83.60(6), P1−P4−P3 96.59(6). All these new compounds, despite of different structural features, have constructed four P−C bonds selectively and efficiently, which is essential to the synthesis of organophosphorus from P4.1,2 Hence, the formation of these new compounds presents an efficient and promising method for cyclotetraphosphanes, which are not easy to access by other methods.14-16 Considering the deficient studies on the chemical properties of cyclotetraphosphanes, the reactivity of 4 was studied (Scheme 5). Treatment of 4 with excess lithium or potassium graphite resulted in the formation of the corresponding phospholyl lithium 5 and potassium 6 in high yields, respectively. In addition, treatment of 4 with excess iodomethane generated phospholium 7 in 72% yield. The structures of 5-7 were determined unambiguously by single-

The Supporting Information is available free of charge on the ACS Publication website at http://pubs.acs.org. Experimental details, X-ray data of 2a, 2b, 3, 6 and 7, Computational details, and NMR spectra of new products.

*[email protected] *[email protected] Shanshan Du: 0000-0001-7655-9765 Jimin Yang: 0000-0002-6493-9813 Jingyuan Hu: 0000-0002-7250-1344 Zhengqi Chai: 0000-0003-2151-1507 Gen Luo: 0000-0002-5297-6756 Yi Luo: 0000-0001-6390-8639 Wen-Xiong Zhang: 0000-0003-0744-2832 Zhenfeng Xi: 0000-0003-1124-5380

§S.D.

and J.Y. contributed equally to this work.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The authors declare no competing financial interests.

This work was supported by the National Natural Science Foundation of China (nos. 21725201, 21890721, 21572005). G.L. and Y.L. thank the Fundamental Research Funds for the Central Universities (DUT18RC(3)002, DUT18GJ201).

(1) (a) Cossairt, B. M.; Piro, N. A.; Cummins, C. C. Early-transitionmetal-mediated activation and transformation of white phosphorus. Chem. Rev. 2010, 110, 4164–4177. (b) Caporali, M.; Gonsalvi, L.; Rossin, A.; Peruzzini, M. P4 activation by late-transition metal complexes. Chem. Rev. 2010, 110, 4178–4235. (c) Scheer, M.; Balázs, G.; Seitz, A. P4 activation by main group elements and compounds. Chem. Rev. 2010, 110, 4236–4256. (2) (a) Giffin, N. A.; Masuda, J. D. Reactivity of white phosphorus with compounds of the p-block. Coord. Chem. Rev. 2011, 255, 1342–1359. (b) Khan, S.; Sen, S. S.; Roesky, H. W. Activation of phosphorus by group 14 elements in low oxidation states. Chem. Commun. 2012, 48, 2169–2178. (c) Holthausen, M. H.; Weigand, J. J. The chemistry of cationic polyphosphorus cages-syntheses, structure and reactivity. Chem. Soc. Rev. 2014, 43, 6639–6657. (d) Borger, J. E.; Ehlers, A. W.; Slootweg, J. C.; Lammertsma, K. Functionalization of P4 through direct P–C bond formation. Chem. Eur. J. 2017, 23, 11738–11746. (3) (a) Rauhut, M. M.; Semsel, A. M. Reactions of elemental phosphorus with organometallic compounds. J. Org. Chem. 1963, 28, 471–473. (b) Rauhut, M. M.; Semsel, A. M. Reactions of elemental phosphorus with organometallic compounds and alkyl halides. The direct synthesis of tertiary phosphines and cyclotetraphosphines. J. Org. Chem. 1963, 28, 473– 477. (4) (a) Barbaro, P.; Peruzzini, M.; Ramirez, J. A.; Vizza. F. Rhodiummediated functionalization of white phosphorus: a novel formation of C–P bonds. Organometallics 1999, 18, 4237–4240. (b) Hübner, A.; Bernert, T.; Sänger, I.; Alig, E.; Bolte, M.; Fink, L.; Wagner, M.; Lerner, H.-W. Solvent-free mesityllithium: solid-state structure and its reactivity towards white phosphorus. Dalton Trans. 2010, 39, 7528–7533. (c) Arrowsmith, M.; Hill, M. S.; Johnson, A. L.; Kociok-Köhn, G.; Mahon, M. F. Attenuated organomagnesium activation of white phosphorus. Angew. Chem., Int. Ed. 2015, 54, 7882−7885. (5) (a) Borger, J. E.; Ehlers, A. W.; Lutz, M.; Slootweg, J. C.; Lammertsma, K. Functionalization of P4 using a Lewis acid stabilized bicyclo-[1.1.0]tetraphosphabutane anion. Angew. Chem., Int. Ed. 2014, 53, 12836–12839. (b) Borger, J. E.; Ehlers, A. W.; Lutz, M.; Slootweg, J. C.; Lammertsma, K. Selective [3+1] fragmentations of P4 by “P” transfer from a Lewis acid stabilized [RP4]– butterfly anion. Angew. Chem., Int. Ed. 2017, 56, 285–290. (c) Bhattacharyya, K. X.; Dreyfuss, S.; Saffon-Merceronb, N.; Mézailles, N. Chem. Commun. 2016, 52, 5179–5182. (6) (a) Khan, S.; Michel, R.; Dieterich, J. M.; Mata, R. A.; Roesky, H. W.; Demers, J.-P.; Lange, A.; Stalke, D. Preparation of RSn(I)–Sn(I)R with two unsymmetrically coordinated Sn(I) atoms and subsequent gentle activation of P4. J. Am. Chem. Soc. 2011, 133, 17889–17894. (b) Dube, J. W.; Graham, C. M. E.; Macdonald, C. L. B.; Brown, Z. D.; Power, P. P.; Ragogna, P. J. Reversible, photoinduced activation of P4 by low-coordinate main group compounds. Chem. Eur. J. 2014, 20, 6739–6744. (7) (a) Figueroa, J. S.; Cummins, C. C. Diorganophosphanylphosphinidenes as complexed ligands: synthesis via an anionic terminal phosphide of niobium. Angew. Chem., Int. Ed. 2004, 43, 984–988. (b) Tran, B. L.; Singhal, M.; Park, H.; Lam, O. P.; Pink, M.; Krzystek, J.; Ozarowski, A.; Telser, J.; Meyer, K.; Mindiola, D. J. Reactivity studies of a masked three-coordinate vanadium(II) complex. Angew. Chem., Int. Ed. 2010, 49, 9871–9875. (c) Pinter, B.; Smith, K. T.; Kamitani, M.; Zolnhofer, E. M.; Tran, B. L.; Fortier, S.; Pink, M.; Wu, G.; Manor, B. C.; Meyer, K.; Baik, M.-H.; Mindiola, D. J. Cyclo-P3 complexes of vanadium: redox properties and origin of the 31P NMR chemical shift. J. Am. Chem. Soc. 2015, 137, 15247–15261. (8) (a) Demange, M.; Le Goff, X.-F.; Le Floch, P.; Mézailles, N. P4 activation with Pt0 metal centers: selective formation of a dinuclear {Pt2(μ,η2:2-P2)} complex. Chem. Eur. J. 2010, 16, 12064–12068. (b) Yao, S.; Lindenmaier, N.; Xiong, Y.; Inoue, S.; Szilvási, T.; Adelhardt, M.; Sutter, J.; Meyer, K.; Driess, M. A neutral tetraphosphacyclobutadiene ligand in cobalt(I) complexes. Angew. Chem., Int. Ed. 2015, 54, 1250–1254. (c) Spitzer, F.; Graßl, C.; Balázs, G.; Zolnhofer, E. M.; Meyer, K.; Scheer,

Page 4 of 6

M. Influence of the nacnac ligand in iron(I)-mediated P4 transformations. Angew. Chem., Int. Ed. 2016, 55, 4340–4344. (9) (a) Konchenko, S. N.; Pushkarevsky, N. A.; Gamer, M. T.; Köppe, R.; Schnöckel, H.; Roesky, P. W. [{(η5-C5Me5)2Sm}4P8]: a molecular polyphosphide of the rare-earth elements. J. Am. Chem. Soc. 2009, 131, 5740–5741. (b) Huang, W.; Diaconescu, P. L. P4 activation by group 3 metal arene complexes. Chem. Commun. 2012, 48, 2216–2218. (c) Huang, W.; Diaconescu, P. L. P4 activation by lanthanum and lutetium naphthalene complexes supported by a ferrocene diamide ligand. Eur. J. Inorg. Chem. 2013, 4090–4096. (10) (a) Masuda, J. D.; Schoeller, W. W.; Donnadieu, B.; Bertrand, G. Carbene activation of P4 and subsequent derivatization. Angew. Chem., Int. Ed. 2007, 46, 7052–7055. (b) Martin, C. D.; Weinstein, C. M.; Moore, C. E.; Rheingold, A. L.; Bertrand, G. Exploring the reactivity of white phosphorus with electrophilic carbenes: synthesis of a P4 cage and P8 clusters. Chem. Commun. 2013, 49, 4486–4488. (11) Tofan, D.; Cummins, C. C. Photochemical incorporation of diphosphorus units into organic molecules. Angew. Chem., Int. Ed. 2010, 49, 7516–7518. (12) (a) Cossairt, B. M.; Cummins, C. C. Radical synthesis of trialkyl, triaryl, trisilyl and tristannyl phosphines from P4. New J. Chem. 2010, 34, 1533–1536. (b) Heinl, S.; Reisinger, S.; Schwarzmaier, C.; Bodensteiner, M.; Scheer, M. Selective functionalization of P4 by metal-mediated C–P bond formation. Angew. Chem., Int. Ed. 2014, 53, 7639–7642. (c) Heinl, S.; Scheer, M. Activation of group 15 based cage compounds by [CpBIGFe(CO)2] radicals. Chem. Sci. 2014, 5, 3221–3225. (13) (a) Xu, L.; Chi, Y.; Du, S.; Zhang, W.-X.; Xi, Z. Direct synthesis of phospholyl lithium from white phosphorus. Angew. Chem., Int. Ed. 2016, 55, 9187–9190. (b) Du, S.; Yin, J.; Chi, Y.; Xu, L.; Zhang, W.-X. Dual functionalization of white phosphorus: formation, characterization, and reactivity of rare-earth-metal cyclo-P3 complexes. Angew. Chem., Int. Ed. 2017, 56, 15886–15890. (c) Du, S.; Zhang, W.-X.; Xi, Z. Diversified aggregation states of phospholyl lithiums. Organometallics 2018, 37, 2018–2022. (d) Du, S.; Hu, J.; Chai, Z.; Zhang, W.-X.; Xi, Z. Isolation and characterization of four phosphorus cluster anions P73-, P144-, P162- and P264from the nucleophilic functionalization of white phosphorus with 1,4dilithio-1,3-butadienes. Chin. J. Chem. 2019, 37, 71–75. (14) Henderson, W. A.; Epstein, M.; Seichter, F. S. Some aspects of the chemistry of cyclopolyphosphines. J. Am. Chem. Soc. 1963, 85, 2462–2466. (15) Huy, N. H. T.; Inubushi, Y.; Ricard, L.; Mathey, F. Thermal [2+2] cycloadditions between diphosphene complexes and alkynes. Organometallics 1997, 16, 2506–2508. (16) (a) Kato, T.; Gornitzka, H.; Schoeller, W. W.; Baceiredo, A.; Bertrand, G. Dimerization of a cyclo-1σ4,3σ2,4σ2-triphosphapentadienyl radical: evidence for phosphorus–phosphorus odd-electron bonds. Angew. Chem., Int. Ed. 2005, 44, 5497–5500. (b) Weigand, J. J.; Burford, N.; Davidson, R. J.; Cameron, T. S.; Seelheim, P. New synthetic procedures to catena-phosphorus cations: preparation and dissociation of the first cyclophosphino-halophosphonium salts. J. Am. Chem. Soc. 2009, 131, 17943– 17953. (17) Zhang, Y.; Wei, J.; Zhang, W.-X.; Xi, Z. Lithium aluminate complexes and alumoles from 1,4-dilithio-1,3-butadienes and AlEt2Cl. Inorg. Chem. 2015, 54, 10695–10700. (18) (a) Peng, Y.; Fan, H.; Zhu, H.; Roesky, H. W.; Magull, J.; Hughes, C. E. [{HC(CMeNAr)2}2Al2P4](Ar = 2,6-iPr2C6H3): a reduction to a formal {P4}4- charged species. Angew. Chem., Int. Ed. 2004, 43, 3443–3445. (b) Seidel, W. W.; Summerscales, O. T.; Patrick, B. O.; Fryzuk, M. D. Activation of white phosphorus by reduction in the presence of a zirconium diamidodiphosphine macrocycle: formation of a bridging square-planar cyclo-P4 unit. Angew. Chem., Int. Ed. 2009, 48, 115–117. (c) Camp, C.; Maron, L.; Bergman, R. G.; Arnold, J. Activation of white phosphorus by low-valent group 5 complexes: formation and reactivity of cyclo-P4 inverted sandwich compounds. J. Am. Chem. Soc. 2014, 136, 17652–17661. (d) Dielmann, F.; Timoshkin, A.; Piesch, M.; Balázs, G.; Scheer, M. Selective P4 activation by a highly reduced cobaltate: synthesis of dicobalt tetraphosphido complexes. Angew. Chem., Int. Ed. 2017, 56, 1671–1675. (e) Pelties, S.; Maier, T.; Herrmann, D.; de Bruin, B.; Rebreyend, C.; Gärtner, S.; Shenderovich, I. G.; Wolf, R. The cobalt cyclo-P4 sandwich complex and its role in the formation of polyphosphorus compounds. Chem. Eur. J. 2017, 23, 6094–6102. (f) Cavaillé, A.; Saffon-Merceron, N.; Nebra, N.; Fustier-Boutignon, M.; Mézailles, N. Synthesis and reactivity of an enddeck cyclo-P4 iron complex. Angew. Chem., Int. Ed. 2018, 57, 1874–1878. (19) (a) Masuda, J. D.; Schoeller, W. W.; Donnadieu, B.; Bertrand, G. NHC-mediated aggregation of P4: isolation of a P12 cluster. J. Am. Chem. Soc. 2007, 129, 14180–14181. (b) Cossairt, B. M.; Cummins, C. C.

ACS Paragon Plus Environment

Page 5 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Shuttling P3 from niobium to rhodium: the synthesis and use of Ph3SnP3(C6H8) as a P3– synthon. Angew. Chem., Int. Ed. 2010, 49, 1595– 1598. (20) Robertson, A. P. M.; Dyker, C. A.; Gray, P. A.; Patrick, B. O.; Decken, A.; Burford, N. Diverse reactivity of the cyclodiphosphinophosphonium cation [(PtBu)3Me]+: parallels with epoxides and new catena-phosphorus frameworks. J. Am. Chem. Soc. 2014, 136, 14941−14950. (21) Weigand, W.; Cordes, A. W.; Swepston, P. N. Tetra-tertbutylcyclotetraphosphane, [PC(CH3)3]4. Acta Cryst. 1981, B37, 1631−1634. (22) Geissler, B.; Barth, S.; Bergsträsser, U.; Slany, M.; Durkin, J.; Hitchcock, P. B.; Hofmann, M.; Binger, P.; Nixon, J. F.; Schleyer, P. v. R.;

Regitz, M. Tetraphosphatricyclo[4.2.0.02,5]octadienes; new phosphaalkyne cyclotetramers derived from λ3σ2-diphosphete. Angew. Chem., Int. Ed. Engl. 1995, 34, 484−487. (23) Smith, R. C.; Urnezius, E.; Lam, K.-C.; Rheingold, A. L.; Protasiewicz, J. D. Syntheses and structural characterizations of the unsymmetrical diphosphene DmpP = PMes* (Dmp = 2,6-Mes2C6H3, Mes* = 2,4,6-tBu3C6H2) and the cyclotetraphosphane [DmpPPPh]2. Inorg. Chem. 2002, 41, 5296−5299. (24) (a) Mathey, F. The organic chemistry of phospholes. Chem. Rev. 1988, 88, 429–453. (b) Mathey, F. Phospha-organic chemistry: panorama and perspectives. Angew. Chem., Int. Ed. 2003, 42, 1578–1604.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 6

Insert Table of Contents artwork here

ACS Paragon Plus Environment

6