Versatile Coordination Modes of Triphospha-1,4-pentadiene-2,4

2 days ago - The Cu1···Cu1′ distance (2.7708(6) Å) is shorter than the sum of the van der Waals radii (rCu = 1.96 ± 0.10 Å),(8,18a−f) and th...
0 downloads 8 Views 1MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

Versatile Coordination Modes of Triphospha-1,4-pentadiene-2,4diamine Anup K. Adhikari, Toni Grell, Peter Lönnecke, and Evamarie Hey-Hawkins* Faculty of Chemistry and Mineralogy, Institute of Inorganic Chemistry, Johannisallee 29, D-04103 Leipzig, Germany S Supporting Information *

ABSTRACT: 1,3,5-Triphospha-1,4-pentadiene-2,4-diamine reacts with [M(CO)4L] (M = Mo, L = nbd (norbornadiene); M = W, L = 2 CH3CN) to give the chelate complexes [M(CO)4(PMes{C(NHCy)PMes}2-κP1,P3)]. In contrast, an unusual intramolecular rearrangement occurred with [Cu(CH3CN)4]PF6 leading to the dimeric copper(I) complex [Cu(CNCy){PHMesPMesC(NHCy)PMes-κP1,P3}]2(PF6)2. The mechanism of the rearrangement was supported by quantum-mechanical calculations. The transition-metal complexes were characterized by multinuclear NMR spectroscopy, mass spectrometry, infrared spectroscopy, and X-ray crystallography.





INTRODUCTION Ligands containing soft1 phosphorus and hard nitrogen donor atoms have been investigated intensively over the past decades.2 Among the variety of P,N ligand systems, pincer ligands and their metal complexes have received interest due to their high stability, versatility, and activity.3 The introduction of a PC functionality in a pincer ligand enhances the reactivity remarkably.4 The chemistry of low-coordinate phosphorus compounds, such as phosphinidenes and phosphaalkenes, is an active research area,5 as compounds with PC double bonds possess low-lying π* orbitals, which can act as effective π acceptors in the coordination of transition metals.5g,6 The η1 coordination mode of phosphaalkenes toward metal atoms or ions is common, but the η 2 coordination mode of phosphaalkenes is also known (Figure 1).5v,6 Chelating

RESULTS AND DISCUSSION The reaction of 1,3,5-triphospha-1,4-pentadiene-2,4-diamine (1) with 1 equiv of [M(CO)4L] (M = Mo, L = nbd; M = W, L = 2 CH3CN) gave the corresponding molybdenum(0) and tungsten(0) complexes [M(CO)4(PMes{C(NHCy)PMes}2κP1,P3)] (M = Mo (2), M = W (3)), in which 1 coordinates via the two terminal phosphorus atoms in a cis arrangement resulting in a six-membered ring (Scheme 1). Scheme 1. Synthesis of Mo Complex 2 and W Complex 3

Complexes 2 and 3 were isolated as air- and moisture-stable crystalline orange (2) or dark red (3) solids suitable for singlecrystal X-ray diffraction studies. In both complexes the metal center has an almost perfect octahedral coordination sphere (Figure 2 for 2; Supporting Information for 3). The coordination of the phosphorus atoms is in accordance with the hard-soft acid-base (HSAB) principle, as both the phosphorus atom and the group 6 metal atoms are soft;1 the six-membered rings have a boat conformation (Figure 3). The P−C bond lengths of the central P2 atom from 1.852(2) to 1.872(2) Å correspond to single bonds.5k,8 In contrast, the

Figure 1. Different bonding modes of phosphaalkenes in metal complexes.

phosphaalkene complexes have also been described.7 Recently, we reported the synthesis of 1,3,5-triphospha-1,4-pentadiene2,4-diamine (1) with a PC(N)−P−C(N)P arrangement, which contains P, N, and PC double bonds as potential donors.8 Here, we report the coordination behavior of 1 toward molybdenum(0), tungsten(0), and copper(I). © XXXX American Chemical Society

Received: January 10, 2018

A

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

shorter than a carbon−nitrogen single bond.9 Both observations indicate a conjugation of the lone pair of electrons on the nitrogen atoms with the terminal PC bonds. The M−P bond lengths in 2 and 3 are comparable with other reported phosphaalkene complexes, such as cis-[Mo(CO)4(1-PPh2-2CHPMes*-C6H4-κP,P′)] (Mes* = 2,4,6-tBu3C6H2), cis[W(CO) 4(PhPCPh2-κP)2], [W(CO)5(RPCHCHMe2κP)] (R = Ph, tBu, Men (Men = menthyl)), [W(CO)5(RP C(2-C6H4OMe)OEt-κP)] (R = CH(NMe2)2, tBu), [W(CO)5(Me3SiPC(2-C6H4Me)OEt-κP)].10,11 In comparison, the M−P bonds in (diphosphazane)molybdenum(0) tetracarbonyl complexes12 are shorter and in (diphosphinidenecyclobutene)molybdenum(0) tetracarbonyl complexes13 longer than observed for 2 and 3. The M−CO bonds trans to the phosphorus atoms of the bidentate ligand 1 are shorter (∼1.98 Å) than the other two M−CO bonds (∼2.04 Å), which is due to the weaker π-acid character of the phosphaalkene group compared to the carbonyl group. The 31P{1H} NMR spectra display a triplet (14.2 ppm for 2, 17.5 ppm for 3) for the central phosphorus atom (P2) (−10.9 ppm in 1)8 and a doublet (80.5 ppm for 2, 55.8 ppm for 3) for the terminal phosphorus atoms (A2B spin system) (97.0 and 102.4 ppm in 1).8 A similar effect was also observed in inversely polarized phosphaalkenes reported by L. Weber and coworkers.11f The 2JPP coupling constants in 2 and 3 (142 and 141 Hz, respectively) are larger than those in the free ligand 1 (2JPP 9 and 59 Hz).8 For complex 3, satellites due to coupling to 183 W (I = 1/2; natural abundance: 14.31%) are observed for the doublet. The 1JPW coupling constant is 236 Hz, also suggesting a cis coordination of the ligand to the W(CO)4 fragment.14 For local C2v symmetry, four absorptions would be expected in the infrared spectra of complexes 2 and 3; only three intense ν(CO) bands in the region of 1860−2013 cm−1 are observed indicating some overlap of absorptions.15 When [Cu(CH3CN)4]PF6 was reacted with 1, the unusual dimeric copper(I) complex 4 was obtained (Scheme 2). In this reaction, a new P−P bond is created between the two neighboring phosphorus atoms of the starting material 1. As 1,3,5-triphospha-1,4-pentadiene-2,4-diamine (1) is obtained as one of three products by insertion of cyclohexyl isonitrile into the P−P bonds of [P4Mes4]2−,8 the reaction observed here between 1 and [Cu(CH3CN)4]PF6 can be regarded as a partway reverse reaction. The molecular structure of 4 (Figure 4) reveals that a rearrangement of 1 occurred during the reaction. Density functional theory (DFT) calculations were performed to gain

Figure 2. Molecular structure of 2. Thermal ellipsoids are set at 50% probability (H atoms other than N−H and disordered diethyl ether molecule are omitted for clarity). Selected bond lengths [Å] and bond angles [deg] of 2 (values of 3 (molecular structure not shown) in []): P2−C1 1.872(2) [1.872(2)], P2−C2 1.858(2) [1.852(2)], P1−C1 1.709(2) [1.712(2)], P3−C2 1.715(2) [1.714(2)], C1−N1 1.358(2) [1.352(2)], C2−N2 1.353(2) [1.356(2)], P1−M1 2.5121(5) [2.5117(4)], P3−M1 2.5116(5) [2.4921(4)], M1−C43 1.987(2) [1.975(2)], M1−C44 1.972(2) [1.985(2)], M1−C42 2.032(2) [2.028(2)], M1−C45 2.041(2) [2.037(2)], C42−O1 1.144(3) [1.148(2)], C43−O2 1.148(3) [1.155(2)], C44−O3 1.155(3) [1.155(2)], C42−O4 1.145(3) [1.146(2)], N1−C1−P2 111.8(1) [109.9(1)], N2−C2−P2 115.0(1) [114.4(2)], N1−C1−P1 132.2(1) [132.5(1)], N2−C2−P3 130.5(1) [130.0(1)], P2−C1−P1 115.47(9) [117.20(8)], P2−C2−P3 113.59(9) [114.47(8)], C2−P2−C1 109.29(8) [110.65(7)], C1−P1−M1 121.74(6) [119.86(5)], C2− P3−M1 121.61(6) [122.37(5)], P3−M1−P1 86.37(2) [86.47(1)], M1−C42−O1 178.1(2) [177.3(2)], M1−C43−O2 178.5(2) [177.5(2)], M1−C44−O3 178.8(2) [177.4(2)], M1−C45−O4 176.5(2) [178.3(2)].

Figure 3. Boat conformation of the six-membered rings in 2 (left) and 3 (right).

terminal P−C bond lengths (1.709(2) to 1.715(2) Å) indicate double bond character but are yet longer than in reported phosphaalkenes (ca. 1.68 Å).5,6 Furthermore, the C−N bond lengths (ca. 1.35 Å) of the sp2 carbon atoms (C1/C2) are Scheme 2. Synthesis of Copper(I) Complex (4)

B

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

proton to a terminal P atom) to give the intermediate I2 (ΔE = +3.9 kJ/mol). Subsequently, an intramolecular rearrangement takes place that includes the formation of the P−P bond and the elimination of a cyclohexyl isocyanide, which coordinates at the CuI cation leading to I3 (ΔE = +28.4 kJ/mol). Interestingly, the sum of all energies in this process going from 1 to I3 is almost zero. Dimerization of I3 finally gives product 4 (ΔE = −117.6 kJ/mol). However, no transition state was found for the transformation from I2 to I3. We expect that at least two transition states are involved in this reaction step. The copper atom in complex 4 is coordinated in a distorted tetrahedral fashion. The P1−Cu1−P3 bond angle of the fivemembered ring is much smaller (89.04(3)°) than the other bond angles (C1−P3−Cu1 (112.67(9)°), P2−C1−P3 (118.8(1)°), C1−P2−P1 (102.21(9)°) and P2−P1−Cu1 (105.53(3)°). The five-membered ring formed by P2−P1− Cu1−P3−C1 exhibits a twist conformation. The P−P (2.194(1) Å) and the P2−C1 (1.829(3) Å) bonds are in the range of single bonds (P−P 2.21 Å,16 P−C 1.85 Å5k,8), while the P3−C1 bond length of 1.747(3) Å indicates multiple bond character but is longer than reported for phosphaalkenes.5,6 The C1−N1 bond length is 1.328(3) Å, which is shorter than a carbon−nitrogen single bond,9 suggesting conjugation of the lone pair of electrons at nitrogen with the PC bond. The Cu1−P1 (2.2932(7) Å) and the Cu1−P3/Cu1−P3′ bond lengths (2.3380(7) and 2.3756(8) Å) are comparable to other copper(I) phosphine complexes.17 The molecule resides on a crystallographic inversion center, which is located in the center of the central planar Cu2P2 ring. The Cu1···Cu1′ distance (2.7708(6) Å) is shorter than the sum of the van der Waals radii (rCu = 1.96 ± 0.10 Å),8,18a−f and the Cu1−P3−Cu1′ bond angle is very small (72.00(2)°), indicating a weak cuprophilic interaction,8,17,18 which is also supported by the calculated Löwdin and Mayer Cu−Cu bond order of 0.1904 and 0.1947, respectively.8 31 1 P{ H} NMR spectrum of 4 in CDCl3 displays a broad signal (−80.3 ppm at room temperature (RT) and −80.8 ppm at −50 °C), a broad doublet (−45.5 ppm at RT and −47.6 ppm at −50 °C) and a broad triplet (−24.6 ppm at RT and −28.4 ppm at −50 °C) with one sharp septet (−144.3 ppm at RT and −144.5 ppm at −50 °C) for the counterion hexafluorophosphate, at room temperature as well as at lower temperature

Figure 4. Molecular structure of the dication of 4. Thermal ellipsoids are set at 50% probability (two counterions (PF6−), H atoms other than N−H, P−H, and disordered THF molecules are omitted for clarity). Selected bond lengths [Å] and bond angles [deg]: P1−P2 2.194(1), P2−C1 1.829(3), P3−C1 1.747(3), C1−N1 1.328(3), P1− Cu1 2.2932(7), P3−Cu1 2.3380(7), P3−Cu1′ 2.3756(8), Cu1−C35 1.926(3), N2−C35 1.152(3), N1−C1−P2 117.6(2), N1−C1−P3 122.4(2), P2−C1−P3 118.8(1), C1−P2−P1 102.21(9), P2−P1−Cu1 105.53(3), P1−Cu1−P3 89.04(3), C1−P3−Cu1 112.67(9), Cu1− P3−Cu1′ 72.00(2), P3−Cu1−P3′ 108.00(2), P3−Cu1−C35 117.69(8), P1−Cu1−C35 120.26(3), P3′−Cu1−C35 119.09(9), P3′−Cu1−P1 97.45(3).

insight into the mechanism of the unusual rearrangement (for details see the Supporting Information). The formation of 4 can be rationalized as follows (Scheme 3): The first step of the reaction is the coordination of the CuI cation by the terminal phosphorus atoms of 1 leading to the formation of I1 (ΔE = −32.8 kJ/mol), which is analogous to the chelate complexes 2 and 3. The CuI ion facilitates a tautomerization reaction (N−H

Scheme 3. Proposed Mechanism of Formation of Copper(I) Complex 4

C

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

mmol) in THF (15 mL) at −78 °C. The color of the solution changed to dark red after 15 min. The red solution was warmed to room temperature over several hours and stirred overnight. The solvent was removed under vacuum, and the resulting red solid was extracted with Et2O (15 mL). Dark red prisms of 3 appeared after 1 d at room temperature. Yield: 0.255 g (95%). 1H NMR (C6D6, 298 K): δ = 0.26−1.50 (m, 22 H, cyclohexyl protons), 1.95 (s, 3 H, p-CH3 in Mes of P2), 2.02 (s, 6 H, p-CH3 in Mes of P1/P3), 2.72 (br, 6 H, o-CH3 in Mes of P2), 2.83−2.86 (bs, 12 H, o-CH3 in Mes of P1/P3), 4.49 (m, 2 H, NH), 6.75 (s, 4 H, m-H in Mes of P1/P3), 6.78 (s, 2 H, m-H in Mes of P2) ppm. 13C{1H} NMR (C6D6, 298 K): δ = 21.1−23.0 (MesCH3), 25.0−33.8 (cyclohexyl carbons), 56.7 (cyclohexyl carbon, NCH), 129.0−130.5 (m-CH in Mes), 134.6 (ipso-C in Mes), 140.0− 142.2 (C−CH3), 177.6 (t, 2JPC = 6.5 Hz, CO), 178.0 (t, 2JPC = 6.5 Hz, CO), 200.8 (m, PC), 206.3 (m, PC) ppm. 31P{1H} NMR (C6D6, 298 K): δ = 17.5 (t, 1 P, 2JPP = 141 Hz, P2), 55.8 (d, 2 P, 2JPP = 141 Hz, 1JPW = 236 Hz, P1/P3) ppm. FT-IR (KBr, cm−1): ν̃ = 3321 (s, NH), 2922 (m, CH, sp3), 2008 (s, CO), 1891 (s, CO), 1857 (s, CO), 1601 (m, PC). MS (ESI-LRMS, pos, n-hexane/CH3CN): m/z = 966.3 [M]+, 938.3 [M − CO]+. Anal. Calcd: C45H57N2O4P3W (966.72), calcd. C 55.91, H 5.94, N 2.90; found C 56.29, H 6.16, N 2.61%. mp 184−186 °C. 4: A suspension of [Cu(CH3CN)4]PF6 (0.125 g, 0.335 mmol) in THF (15 mL) was added dropwise to a solution of 1 (0.220 g, 0.328 mmol) in THF (15 mL) at −78 °C. The color of the solution changed to bright orange after 15 min. The orange solution was warmed to room temperature over several hours and stirred overnight. The solution was concentrated to 15 mL and layered with n-hexane (5 mL). Orange prisms of 4 appeared after 1 d at room temperature. Yield: 0.272 g (94%). 1H NMR (CDCl3, 298 K): δ = 0.31−3.14 (m, 44 H, cyclohexyl protons), 2.24−2.62 (54 H, CH3 in Mes), 4.27 (bs, 2 H, NH), 6.68 (br, 2 H, PH, 1JPH = 157 Hz), 6.46−7.01 (12 H, m-H in Mes) ppm. 1H NMR (CDCl3, 223 K): δ = 0.36−3.10 (m, 44 H, cyclohexyl protons), 0.95−2.76 (54 H, CH3 in Mes), 4.10 (br, 2 H, NH), 6.45 (bd, 2 H, PH), 6.39−7.04 (12 H, m-H in Mes) ppm. 13 C{1H} NMR (CDCl3, 298 K): δ = 21.1−24.7 (Mes-CH3), 25.6−32.2 (cyclohexyl carbons), 54.4−60.4 (cyclohexyl carbon, NCH), 115.3− 128.5 (ipso-C in Mes), 129.9−131.2 (m-CH in Mes), 137.8−145.3 (C−CH3, CyNC), 195.7 (PC) ppm. 31P{1H} NMR (CDCl3, 298 K): δ = −24.6 (bt, 1 P, P2), −45.5 (bd, 1 P, P1), −80.3 (br, 1 P, P3), −144.3 (sept, 1 P, 1JPF = 712.7 Hz, PF6) ppm. 31P{1H} NMR (CDCl3, 223 K): δ = −28.4 (bt, 1 P, P2), −47.6 (bd, 1 P, P1), −80.8 (br, 1 P, P3), −144.5 sept, 1 P, PF6, 1JPF = 713 Hz) ppm. FT-IR (KBr, cm−1): ν̃ = 3301 (s, NH), 2934 (m, CH, sp3), 2190 (s, CN), 2378 (w, P−H), 1603 (s, PC). MS (ESI-HRMS, pos, CHCl3/CH3CN): m/z = 1185.4956 [M − 2 PF6 − Cu − 2 CyNC]+, 733.3011 [monomer − PF6]+, 624.2120 [monomer − PF6 − CyNC]+. Anal. Calcd: C82H114Cu2F12N4P8·4C4H8O (2047.02), calcd. C 57.51, H 7.19, N 2.74; found C 56.25, H 6.90, N 2.73%. mp 177−180 °C. X-ray data were collected with a Gemini CCD diffractometer (Rigaku Inc.), λ(Mo Kα) = 0.710 73 Å, T = 130(2) K, empirical absorption corrections with SCALE3 ABSPACK.24 The structures were solved by dual space methods with SHELXT-2014. Structure refinement was done with SHELXL-201725 by using full-matrix leastsquares routines against F2. With the exception of NH and PH fragments all hydrogen atoms were calculated on idealized positions. The pictures were generated with the program Diamond.26 CCDC 1814283 (2), CCDC 1814284 (3), and CCDC 1814285 (4) contain the supplementary crystallographic data for this paper. Additional crystallographic information is available in the Accession Codes. Crystal Data for 2. C49H67MoN2O5P3, M = 952.89, triclinic, space group P1̅, a = 12.3053(4) Å, b = 12.7264(4) Å, c = 18.0638(4) Å, α = 99.842(2)°, β = 95.997(2)°, γ = 114.441(3)°, V = 2488.6(1) Å3, Z = 2, ρcalcd = 1.272 Mg·m−3, μ(Mo Kα) = 0.405 mm−1, θmax = 32.68°, R = 0.0589, Rw = 0.1082 (all data), 16 729 independent reflections, 558 parameters, 7 restraints, residual electron density 1.658 and −0.833 e· Å−3. Crystal Data for 3. C45H57N2O4P3W, M = 966.68, monoclinic, space group P21/n, a = 15.4814(1) Å, b = 20.1559(1) Å, c = 15.4985(1) Å, β = 111.699(1)°, V = 4493.48(5) Å3, Z = 4, ρcalcd =

(−50 °C). The 1H NMR spectrum in CDCl3 also displays broad signals for the methyl protons and the aromatic protons at room temperature; however, these signals are getting sharper at lower temperature (−50 °C), which indicates dynamic behavior (monomer−dimer) of complex 4 in solution. The infrared spectrum of 4 shows sharp bands at 3302 cm−1 (ν(N− H)) and 2109 cm−1 (ν(CN)), for the coordinated cyclohexyl isocyanide ligand,19 and at 2378 cm−1 (ν(P−H)) for the secondary phosphine.



CONCLUSION 1,3,5-Triphospha-1,4-pentadiene-2,4-diamine (1) reacts with group 6 metal(0) carbonyls (metal = molybdenum and tungsten) to give chelate complexes in which only the terminal phosphorus atoms coordinate to the metal center. In contrast, the reaction with copper(I) leads to an intramolecular rearrangement of 1 with P−P bond formation and elimination of cyclohexyl isocyanide. Further reactivity studies of 1 are underway.



EXPERIMENTAL SECTION

General Remarks. All experiments were performed under dry nitrogen. Only the addition of metal complexes to a solution of 1 was performed in the dark. Solvents (tetrahydrofuran (THF) and Et2O) were dried and freshly distilled under nitrogen and kept over a potassium mirror or molecular sieves (4 Å). The NMR spectra were recorded at 25 °C with a Bruker Avance DRX 400 spectrometer (1H NMR: 400.13 MHz, 13C NMR: 100.16 MHz, 31P NMR: 161.97 MHz). Tetramethylsilane (TMS) was used as internal standard for 1H NMR spectra. 13C and 31P NMR experiments were referenced to TMS on the Ξ scale.20 In all 31P NMR experiments, 85% H3PO4 in water was used as external standard. Electrospray ionization mass spectrometry (ESI-MS) measurements were performed with a Bruker Daltonics FT-ICR-MS spectrometer (Type APEX II, 7 T). Elemental analyses were performed with a Heraeus VARIO EL oven. Infrared spectra were recorded with a spectrometer from PerkinElmer (System 2000) in the range from 4000 to 400 cm−1. The melting points (Gallenkamp) were determined in a capillary tube. Ligand 1,8 [Mo(CO)4(nbd)],21 and [Cu(CH3CN)4]PF622 were synthesized according to the literature. [W(CO)4(CH3CN)2] was synthesized by modification of the known literature procedure.23 2: A solution of [Mo(CO)4(nbd)] (0.091 g, 0.303 mmol) in THF (15 mL) was added dropwise to a solution of 1 (0.200 g, 0.298 mmol) in THF (15 mL) at −78 °C. The color of the solution changed to bright orange after 15 min. The orange solution was warmed to room temperature over several hours and stirred overnight. The solvent was removed under vacuum, and the resulting orange solid was extracted with Et2O (15 mL). Dark orange prisms of 2 appeared after 1 d at room temperature. Yield: 0.255 g (93%). 1H NMR (C6D6, 298 K): δ = 0.26−1.51 (m, 22 H, cyclohexyl protons), 1.95 (s, 3 H, p-CH3 in Mes of P2), 2.02 (s, 6 H, p-CH3 in Mes of P1/P3), 2.71 (br, 6 H, o-CH3 in Mes of P2), 2.83−2.87 (bs, 12 H, o-CH3 in Mes of P1/P3), 4.56 (m, 2 H, NH), 6.76 (s, 4 H, m-H in Mes of P1/P3), 6.78 (s, 2 H, m-H in Mes of P2) ppm. 13C{1H} NMR (C6D6, 298 K): δ = 21.1−23.2 (MesCH3), 25.0−33.8 (cyclohexyl carbons), 56.7 (cyclohexyl carbon, NCH), 129.0−130.4 (m-CH in Mes), 135.4 (m, ipso-C in Mes) 139.8−142.1 (C−CH3), 179.5 (br, CO), 180.0 (br, CO), 208.7 (m, PC), 216.0 (m, PC) ppm. 31P{1H} NMR (C6D6, 298 K): δ = 14.2 (t, 1 P, 2JPP = 142 Hz, P2), 80.5 (d, 2 P, 2JPP = 142 Hz, P1/P3) ppm. FT-IR (KBr, cm−1): ν̃ = 3323 (s, NH), 2923 (m, CH, sp3), 2013 (s, CO), 1902 (s, CO), 1860 (s, CO), 1601 (m, PC). MS (ESI-HRMS, pos, CH3Cl/CH3CN): m/z = 881.3 [M + H]+, 1781.5196 [2 M + Na]+. Anal. Calcd: C45H57MoN2O4P3·C4H10O (952.97), calcd. C 61.76, H 7.09, N 2.94; found C 61.76, H 6.77, N 2.94%. mp 178−181 °C. 3: A solution of [W(CO)4(CH3CN)2] (0.097 g, 0.256 mmol) in THF (15 mL) was added dropwise to a solution of 1 (0.170 g, 0.253 D

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry 1.429 Mg·m−3, μ(Mo Kα) = 2.720 mm−1, θmax = 37.38°, R = 0.0463, Rw = 0.0632 (all data), 22 671 independent reflections, 513 parameters, residual electron density 1.790 and −1.694 e·Å−3. Crystal Data for 4. C98H146Cu2F12N4O4P8, M = 2047.02, monoclinic, space group P21/n, a = 19.0776(5) Å, b = 13.5866(4) Å, c = 20.0700(5) Å, β = 102.835(3)°, V = 5072.2(2) Å3, Z = 2, ρcalcd = 1.340 Mg·m−3, μ(Mo Kα) = 0.618 mm−1, θmax = 30.20°, R = 0.0976, Rw = 0.1445 (all data), 13 935 independent reflections, 635 parameters, 73 restraints, residual electron density 0.754 and −0.583 e·Å−3.



hetero- and homobimetallic complexes. Coord. Chem. Rev. 2009, 253, 1793−1832. (e) Fanfoni, L.; Meduri, A.; Zangrando, E.; Castillon, S.; Felluga, F.; Milani, B. New Chiral P-N Ligands for the Regio- and Stereoselective Pd-Catalyzed Dimerization of Styrene. Molecules 2011, 16, 1804−1824. (f) Nareddy, P.; Mantilli, L.; Guénée, L.; Mazet, C. Atropoisomeric (P,N) Ligands for the Highly Enantioselective PdCatalyzed Intramolecular Asymmetric α-Arylation of α-Branched Aldehydes. Angew. Chem., Int. Ed. 2012, 51, 3826−3831. (g) Mazuela, J.; Paptchikhine, A.; Tolstoy, P.; Pámies, O.; Diéguez, M.; Andersson, P. G. A New Class of Modular P,N-Ligand Library for Asymmetric PdCatalyzed Allylic Substitution Reactions: A Study of the Key Pd−πAllyl Intermediates. Chem. - Eur. J. 2010, 16, 620−638. (h) Lin, C.-H.; Nesterov, V. N.; Richmond, M. G. 2-[(Diphenylphosphino)methyl]-6methylpyridine (PN) coordination chemistry at triosmium clusters: Regiospecific ligand activation and DFT evaluation of the isomeric Os3(CO)10(PN) clusters. J. Organomet. Chem. 2013, 744, 24−34. (i) Mata, Y.; Dieguez, M. Pyranoside Phosphite-Oxazoline Ligand Library: Highly Efficient Modular P,N Ligands for PalladiumCatalyzed Allylic Substitution Reactions. A Study of the Key Palladium Allyl Intermediates. Adv. Synth. Catal. 2009, 351, 3217−3234. (j) Willms, H.; Frank, W.; Ganter, C. Coordination Chemistry and Catalytic Application of Bidentate Phosphaferrocene−Pyrazole and − Imidazole Based P,N-Ligands. Organometallics 2009, 28, 3049−3058. (k) Benito-Garagorri, D.; Kirchner, K. Modularly Designed Transition Metal PNP and PCP Pincer Complexes based on Aminophosphines: Synthesis and Catalytic Applications. Acc. Chem. Res. 2008, 41, 201− 213. (l) Roseblade, S. J.; Pfaltz, A. Iridium-Catalyzed Asymmetric Hydrogenation of Olefins. Acc. Chem. Res. 2007, 40, 1402−1411. (m) Braunstein, P. Bonding and Organic and Inorganic Reactivity of Metal-Coordinated Phosphinoenolates and Related Functional Phosphine-Derived Anions. Chem. Rev. 2006, 106, 134−159. (n) Guiry, P. J.; Saunders, C. P. The Development of Bidentate P,N Ligands for Asymmetric Catalysis. Adv. Synth. Catal. 2004, 346, 497− 537. (o) Chelucci, G.; Orru, G.; Pinna, G. A. Chiral P,N-ligands with pyridine-nitrogen and phosphorus donor atoms. Syntheses and applications in asymmetric catalysis. Tetrahedron 2003, 59, 9471− 9515. (p) Pfaltz, A.; Blankenstein, J.; Hilgraf, R.; Hörmann, E.; McIntyre, S.; Menges, F.; Schönleber, M.; Smidt, S. P.; Wüstenberg, B.; Zimmermann, N. Iridium-Catalyzed Enantioselective Hydrogenation of Olefins. Adv. Synth. Catal. 2003, 345, 33−43. (q) Espinet, P.; Soulantica, K. Phosphine-pyridyl and related ligands in synthesis and catalysis. Coord. Chem. Rev. 1999, 193−195, 499−556. (r) Slone, C. S.; Weinberger, D. A.; Mirkin, C. A. In Progress in Inorganic Chemistry; Karlin, K. D., Ed.; Wiley: New York, 1999; Vol. 48, pp 233−350. (s) Zhang, S.; Pattacini, R.; Braunstein, P. In Advances in Organometallic Chemistry and Catalysis: The Silver/Gold Jubilee International Conference on Organometallic Chemistry Celebratory Book; Pombeiro, A. J. L., Ed.; Wiley: Hoboken, NJ, 2013; pp 185− 198. (t) Carroll, M. P.; Guiry, P. J. P,N ligands in asymmetric catalysis. Chem. Soc. Rev. 2014, 43, 819−833. (3) (a) Morales-Morales, D.; Jensen, C. M. The Chemistry of Pincer Compounds; Elsevier: Amsterdam, Netherlands, 2007. (b) van Koten, G.; Milstein, D. Organometallic Pincer Chemistry; Springer: Heidelberg, Germany, 2013. (c) van der Boom, M. E.; Milstein, D. Cyclometalated Phosphine-Based Pincer Complexes: Mechanistic Insight in Catalysis, Coordination, and Bond Activation. Chem. Rev. 2003, 103, 1759− 1792. (d) Choi, J.; MacArthur, A. H. R.; Brookhart, M.; Goldman, A. S. Dehydrogenation and Related Reactions Catalyzed by Iridium Pincer Complexes. Chem. Rev. 2011, 111, 1761−1779. (e) Selander, N.; Szabó, K. J. Catalysis by Palladium Pincer Complexes. Chem. Rev. 2011, 111, 2048−2076. (f) Gunanathan, C.; Milstein, D. Applications of Acceptorless Dehydrogenation and Related Transformations in Chemical Synthesis. Science 2013, 341, 1229712. (g) Gunanathan, C.; Milstein, D. Bond Activation and Catalysis by Ruthenium Pincer Complexes. Chem. Rev. 2014, 114, 12024−12087. (h) Haenel, M. W.; Oevers, S.; Angermund, K.; Kaska, W. C.; Fan, H.-J.; Hall, M. B. Thermally Stable Homogeneous Catalysts for Alkane Dehydrogenation. Angew. Chem., Int. Ed. 2001, 40, 3596−3600. (i) Gossage, R. A.; van de Kuil, L. A.; van Koten, G. Diaminoarylnickel(II) “Pincer”

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b00067. Molecular structure of 3, detailed theoretical calculations of the reaction mechanism for the formation of complex 4, multinuclear NMR and ESI mass spectra for 2, 3, and 4 (PDF) Reactivity to forming chelate complexes; characterization of transition-metal complexes (PDF) Accession Codes

CCDC 1814283−1814285 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Evamarie Hey-Hawkins: 0000-0003-4267-0603 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Support from the Deutscher Akademischer Austauschdienst (DAAD-GSSP doctoral grant for A.K.A.), the Studienstiftung des deutschen Volkes (doctoral grant for T.G.), and the Graduate School BuildMoNa is gratefully acknowledged. We thank M.Sc. B. Schwarze for measurement of several NMR spectra.

■ ■

DEDICATION Dedicated to Prof. Dr. Wolfgang Weigand on the occasion of his 60th birthday. REFERENCES

(1) Pearson, R. G. Hard and Soft Acids and Bases. J. Am. Chem. Soc. 1963, 85, 3533−3539. (2) (a) Gavrilov, K. N.; Polosukhin, A. I. Chiral P,N-bidentate ligands in coordination chemistry and organic catalysis involving rhodium and palladium. Russ. Chem. Rev. 2000, 69, 661−682. (b) Müller, G.; Klinga, M.; Osswald, P.; Leskelä, M.; Rieger, B. Palladium Complexes with Bidentate P,N Ligands: Synthesis, Characterization and Application in Ethene Oligomerization. Z. Naturforsch. B 2002, 57, 803−809. (c) Kermagoret, A.; Tomicki, F.; Braunstein, P. Nickel and Iron Complexes with N,P,N-type Ligands: Synthesis, Structure and Catalytic Oligomerization of Ethylene. Dalton Trans. 2008, 2945− 2955. (d) Maggini, S. Classification of P,N-binucleating ligands for E

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Complexes: Mechanistic Considerations in the Kharasch Addition Reaction, Controlled Polymerization, and Dendrimeric Transition Metal Catalysts. Acc. Chem. Res. 1998, 31, 423−431. (j) Albrecht, M.; van Koten, G. Platinum Group Organometallics Based on “Pincer” Complexes: Sensors, Switches, and Catalysts. Angew. Chem., Int. Ed. 2001, 40, 3750−3781. (k) Murugesan, S.; Kirchner, K. Non-precious metal complexes with an anionic PCP pincer architecture. Dalton Trans. 2016, 45, 416−439. (l) Schröder-Holzhacker, C.; Stöger, B.; Pittenauer, E.; Allmaier, G.; Veiros, L. F.; Kirchner, K. High-spin iron(II) complexes with mono-phosphorylated 2,6-diaminopyridine ligands. Monatsh. Chem. 2016, 147, 1539−1545. (m) Shih, W.-C.; Ozerov, O. V. One-Pot Synthesis of 1,3-Bis(phosphinomethyl)arene PCP/PNP Pincer Ligands and Their Nickel Complexes. Organometallics 2015, 34, 4591−4595. (4) (a) Chang, Y.-H.; Nakajima, Y.; Tanaka, H.; Yoshizawa, K.; Ozawa, F. Facile N−H Bond Cleavage of Ammonia by an Iridium Complex Bearing a Noninnocent PNP-Pincer Type Phosphaalkene Ligand. J. Am. Chem. Soc. 2013, 135, 11791−11794. (b) Chang, Y.-H.; Nakajima, Y.; Tanaka, H.; Yoshizawa, K.; Ozawa, F. Mechanism of N− H Bond Cleavage of Aniline by a Dearomatized PNP-Pincer Type Phosphaalkene Complex of Iridium(I). Organometallics 2014, 33, 715−721. (c) Nakajima, Y.; Nakao, Y.; Sakaki, S.; Tamada, Y.; Ono, T.; Ozawa, F. Electronic Structure of Four-Coordinate Iron(I) Complex Supported by a Bis(phosphaethenyl)pyridine Ligand. J. Am. Chem. Soc. 2010, 132, 9934−9936. (d) Lin, Y.-F.; Ichihara, N.; Nakajima, Y.; Ozawa, F. Disproportionation of Bis(phosphaethenyl)pyridine Iron(I) Bromide Induced by tBuNC. Organometallics 2014, 33, 6700−6703. (e) Taguchi, H.-O.; Sasaki, D.; Takeuchi, K.; Tsujimoto, S.; Matsuo, T.; Tanaka, H.; Yoshizawa, K.; Ozawa, F. Unsymmetrical PNP-Pincer Type Phosphaalkene Ligands Protected by a Fused-Ring Bulky Eind Group: Synthesis and Applications to Rh(I) and Ir(I) Complexes. Organometallics 2016, 35, 1526−1533. (f) Serin, S. C.; Pick, F. S.; Dake, G. R.; Gates, D. P. Copper(I) Complexes of Pyridine-Bridged Phosphaalkene-Oxazoline Pincer Ligands. Inorg. Chem. 2016, 55, 6670−6678. (g) Miura-Akagi, P. M.; Nakashige, M. L.; Maile, C. K.; Oshiro, S. M.; Gurr, J. R.; Yoshida, W. Y.; Royappa, A. T.; Krause, C. E.; Rheingold, A. L.; Hughes, R. P.; Cain, M. F. Synthesis of a Tris(phosphaalkene)phosphine Ligand and Fundamental Organometallic Reactions on Its Sterically Shielded Metal Complexes. Organometallics 2016, 35, 2224−2231. (5) (a) Appel, R.; Knoll, F.; Ruppert, I. Phospha-alkenes and Phospha-alkynes, Genesis and Properties of the (p-p)π-Multiple Bond. Angew. Chem., Int. Ed. Engl. 1981, 20, 731−744. (b) Appel, R.; Knoll, F. Double Bonds Between Phosphorus and Carbon. Adv. Inorg. Chem. 1989, 33, 259−361. (c) Weber, L. Phosphaalkenes with Inverse Electron Density. Eur. J. Inorg. Chem. 2000, 2000, 2425−2441. (d) Yoshifuji, M. Sterically protected organophosphorus compounds in low co-ordination states. J. Chem. Soc., Dalton Trans. 1998, 3343− 3350. (e) Gaumont, A. C.; Denis, J. M. Preparation, Characterization, and Synthetic Potential of Unstable Compounds Containing Phosphorus-Carbon Multiple Bonds. Chem. Rev. 1994, 94, 1413− 1439. (f) Mathey, F. Expanding the Analogy between PhosphorusCarbon and Carbon-Carbon Double Bonds. Acc. Chem. Res. 1992, 25, 90−96. (g) Mathey, F. Phospha-Organic Chemistry: Panorama and Perspectives. Angew. Chem., Int. Ed. 2003, 42, 1578−1604. (h) Orthaber, A.; Ö berg, E.; Jane, R. T.; Ott, S. Alternative Synthesis and Structures of C-monoacetylenic Phosphaalkenes. Z. Anorg. Allg. Chem. 2012, 638, 2219−2224. (i) Bates, J. I.; Patrick, B. O.; Gates, D. P. A Lewis acid-mediated synthesis of P-alkyl-substituted phosphaalkenes. New J. Chem. 2010, 34, 1660−1666. (j) Yam, M.; Chong, J. H.; Tsang, C.; Patrick, B. O.; Lam, A. E.; Gates, D. P. Scope and Limitations of the Base-Catalyzed Phospha-Peterson PC BondForming Reaction. Inorg. Chem. 2006, 45, 5225−5234. (k) Wang, Y.; Xie, Y.; Wei, P.; King, R. B.; Schaefer, H. F., III; Schleyer, P. v. R.; Robinson, G. H. Carbene-Stabilized Diphosphorus. J. Am. Chem. Soc. 2008, 130, 14970−14971. (l) Back, O.; Kuchenbeiser, G.; Donnadieu, B.; Bertrand, G. Nonmetal-Mediated Fragmentation of P4: Isolation of P1 and P2 Bis(carbene) Adducts. Angew. Chem., Int. Ed. 2009, 48, 5530−5533. (m) Back, O.; Donnadieu, B.; Parameswaran, P.;

Frenking, G.; Bertrand, G. Isolation of crystalline carbene-stabilized P2-radical cations and P2-dications. Nat. Chem. 2010, 2, 369−373. (n) Dyker, C. A.; Bertrand, G. Soluble Allotropes of Main-Group Elements. Science 2008, 321, 1050−1051. (o) 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. (p) 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. (q) Back, O.; Henry-Ellinger, M.; Martin, C. D.; Martin, D.; Bertrand, G. 31P NMR Chemical Shifts of Carbene−Phosphinidene Adducts as an Indicator of the πAccepting Properties of Carbenes. Angew. Chem., Int. Ed. 2013, 52, 2939−2943. (r) Arduengo, A. J.; Calabrese, J. C.; Cowley, A. H.; Dias, H. V. R.; Goerlich, J. R.; Marshall, W. J.; Riegel, B. Carbene− Pnictinidene Adducts. Inorg. Chem. 1997, 36, 2151−2158. (s) Arduengo, A. J., III; Dias, H. V. R.; Calabrese, J. C. A Carbene· Phosphinidene Adduct: “Phosphaalkene”. Chem. Lett. 1997, 26, 143− 144. (t) Rodrigues, R. R.; Dorsey, C. L.; Arceneaux, C. A.; Hudnall, T. W. Phosphaalkene vs. phosphinidene: the nature of the P−C bond in carbonyl-decorated carbene → PPh adducts. Chem. Commun. 2014, 50, 162−164. (u) Schneider, H.; Schmidt, D.; Radius, U. The reductive P−P coupling of primary and secondary phosphines mediated by N-heterocyclic carbenes. Chem. Commun. 2015, 51, 10138−10141. (v) Dillon, K. B.; Mathey, F.; Nixon, J. F. Phosphorus: The Carbon Copy; Wiley, Chichester, UK, 1998. (w) Adhikari, A. K.; Grell, T.; Lönnecke, P.; Hey-Hawkins, E. Formation of a Carbene− Phosphinidene Adduct by NHC-Induced P−P Bond Cleavage in Sodium Tetramesityltetraphosphanediide. Eur. J. Inorg. Chem. 2016, 2016, 620−622. (x) Pietschnig, R.; Orthaber, A. In Advances in the Chemistry of Phosphaalkenes, Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier: Amsterdam, Netherlands, 2016. (y) Gudat, D. In Low-Coordinate Main Group Compounds− Group 15, Comprehensive Inorganic Chemistry II; Reedijk, J., Poeppelmeier, K., Eds.; Elsevier: Amsterdam, Netherlands, 2013; pp 587−621. (z) Weigand, J. J.; Burford, N. In Catenated Phosphorus Compounds, Comprehensive Inorganic Chemistry II; Reedijk, J., Poeppelmeier, K., Eds.; Elsevier: Amsterdam, Netherlands, 2013; pp 119−149. (6) (a) Weber, L. Metallophosphaalkenesfrom Exotics to Versatile Building Blocks in Preparative Chemistry. Angew. Chem., Int. Ed. Engl. 1996, 35, 271−288. (b) Weber, L. Recent developments in the chemistry of metallophosphaalkenes. Coord. Chem. Rev. 2005, 249, 741−763. (c) Nixon, J. F. The Coordination Chemistry of Compounds Containing Phosphorus-Carbon Multiple Bonds. Chem. Rev. 1988, 88, 1327−1362. (d) Scherer, O. J. Phosphorus, Arsenic, Antimony, and Bismuth Multiply Bonded Systems with Low Coordination Number  Their Role as Complex Ligands. Angew. Chem., Int. Ed. Engl. 1985, 24, 924−943. (f) Floch, P. L. Phosphaalkene, phospholyl and phosphinine ligands: New tools in coordination chemistry and catalysis. Coord. Chem. Rev. 2006, 250, 627−681. (g) Dugal-Tessier, J.; Conrad, E. D.; Dake, G. R.; Gates, D. P. In Phosphorus(III) Ligands in Homogeneous Catalysis: Design and Synthesis; Kamer, P. C. J., van Leeuwen, P. W. N. M., Eds.; Wiley: Chichester, UK, 2012; pp 321−341. (h) van der Knaap, T. A.; Bickelhaupt, F.; Kraaykamp, J. G.; van Koten, G.; Bernards, J. P. C.; Edzes, H. T.; Veeman, W. S.; de Boer, E.; Baerends, E. J. η1- and η2coordination in a (phosphaalkene)platinum(0) complex. Organometallics 1984, 3, 1804−1811. (i) Apfel, R.; Casser, C.; Knoch, F. Ü ber niederkoordinierte phosphorverbindungen: XXXX. 2,4,6-tri-tButylphenylmethylenphosphan, ein vielseitiger ligand in übergangsmetall-komplexen. J. Organomet. Chem. 1985, 293, 213−217. (j) Marinetti, A.; Mathey, F. A Novel Entry to the PC-Double Bond: the “Phospha-Wittig” Reaction. Angew. Chem., Int. Ed. Engl. 1988, 27, 1382−1384. (k) Regitz, M.; Scherer, O. J. Multiple Bonds and Low Coordination in Phosphorus Chemistry; Thieme: Stuttgart, Germany, 1990. (l) Le Floch, P.; Marinetti, A.; Ricard, L.; Mathey, F. Synthesis, Structure, and Reactivity of (Phosphoranylidenephosphine)pentacarbonyltungsten Complexes. Another Access to the Phosphorus-Carbon Double Bond. J. Am. F

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Chem. Soc. 1990, 112, 2407−2410. (m) Lorenz, I.-P.; Pohl, W.; Nöth, H.; Schmidt, M. P-Funktionalisierte Diferriophosphonium Salze des Typs [{CpFe(CO)2}2PPhR]+X− und [{CpFe(CO)2}2PClR]+X−. J. Organomet. Chem. 1994, 475, 211−221. (n) Takeuchi, K.; Minami, A.; Nakajima, Y.; Ozawa, F. Synthesis and Structures of Nickel Complexes with a PN-Chelate Phosphaalkene Ligand. Organometallics 2014, 33, 5365−5370. (o) Greenacre, V. K.; Trathen, N.; Crossley, I. R. Ruthenaphosphaalkenyls: Synthesis, Structures, and Their Conversion to η2-Phosphaalkene Complexes. Organometallics 2015, 34, 2533− 2542. (7) (a) Bréque, A.; Santini, C. C.; Mathey, F.; Fischer, J.; Mitschler, A. 4,5-Dimethyl-2-(2-pyridyl)phosphorin as a Chelating Ligand. Synthesis and X-ray Crystal Structure Analysis of (4,5-dimethyl-2-(2pyridyl)phosphorin)tetracarbonylchromium. Inorg. Chem. 1984, 23, 3463−3467. (b) Le Floch, P.; Mathey, F. Transition metals in phosphinine chemistry. Coord. Chem. Rev. 1998, 178−180, 771−791. (c) Klein, M.; Albrecht, C.; Schnakenburg, G.; Streubel, R. A New Route to Phosphaalkene Chelate Complexes: SET Deoxygenation of Oxaphosphirane Complexes Followed by Intramolecular CO Substitution. Organometallics 2013, 32, 4938−4943. (d) Magnuson, K. W.; Oshiro, S. M.; Gurr, J. R.; Yoshida, W. Y.; Gembicky, M.; Rheingold, A. L.; Hughes, R. P.; Cain, M. F. Streamlined Preparation and Coordination Chemistry of Hybrid Phosphine−Phosphaalkene Ligands. Organometallics 2016, 35, 855−859. (e) Ekici, S.; Gudat, D.; Nieger, M.; Nyulaszi, L.; Niecke, E. Kinetically Controlled Protonation of a Cyclic Phosphamethanide Complex to a PHPhosphonium Ylide. Angew. Chem., Int. Ed. 2002, 41, 3367−3371. (f) Ekici, S.; Nieger, M.; Glaum, R.; Niecke, E. A Strategy for the Synthesis of Phosphorus-Containing Macrocycles  Ligands for Exceptional Coordination Geometries. Angew. Chem., Int. Ed. 2003, 42, 435−438. (g) Freytag, M.; Ito, S.; Yoshifuji, M. Coordination Behavior of Sterically Protected Phosphaalkenes on the AuCl Moiety Leading to Catalytic 1,6-Enyne Cycloisomerization. Chem. - Asian J. 2006, 1, 693−700. (h) Orthaber, A.; Belaj, F.; Pietschnig, R. Synthesis, structure and π-delocalization of a phosphaalkenyl based neutral PNP-pincer. Inorg. Chim. Acta 2011, 374, 211−215. (8) Adhikari, A. K.; Sárosi, M. B.; Grell, T.; Lönnecke, P.; HeyHawkins, E. Unusual Reactivity of Sodium Tetramesityltetraphosphanediide towards Cyclohexyl Isocyanide. Chem. - Eur. J. 2016, 22, 15664−15668. (9) (a) Allen, F. H.; Kennard, O.; Watson, D. G.; Brammer, L.; Orpen, A. G.; Taylor, R. Tables of Bond Lengths determined by X-ray and Neutron Diffraction. Part 1. Bond Lengths in Organic Compounds. J. Chem. Soc., Perkin Trans. 2 1987, S1−S19. (b) Feilchenfeld, H. A Relation Between the Lengths of Single, Double and Triple Bonds. J. Phys. Chem. 1959, 63, 1346−1346. (10) Brauer, D. J.; Liek, C.; Stelzer, O. Chelating phosphines with low coordinated phosphorus atoms: Part I. Phosphaalkenes associated with phosphine moieties  X-ray structure of cis-(CO)4Mo[orthoPh2P−C6H4−CHP(t-Bu3C6H2)]·C7H16. J. Organomet. Chem. 2001, 626, 106−112. (11) (a) Wang, S.; Samedov, K.; Serin, S. C.; Gates, D. P. PhP CPh2 and Related Phosphaalkenes: A Solution Equilibrium between a Phosphaalkene and a 1,2-Diphosphetane. Eur. J. Inorg. Chem. 2016, 2016, 4144−4151. (b) Marinetti, A.; Bauer, S.; Ricard, L.; Mathey, F. The ″Phospha-Wittig″ Reaction: A new Method for Building Phosphorus-Carbon Double and Single Bonds from Carbonyl Compounds. Organometallics 1990, 9, 793−798. (c) de Vaumas, R.; Marinetti, A.; Ricard, L.; Mathey, F. Use of Prochiral Phosphaalkene Complexes in the Synthesis of Optically Active Phosphines. J. Am. Chem. Soc. 1992, 114, 261−266. (d) Albrecht, C.; Shi, L.; Pérez, J. M.; van Gastel, M.; Schwieger, S.; Neese, F.; Streubel, R. Deoxygenation of Coordinated Oxaphosphiranes: A New Route to PC Double-Bond Systems. Chem. - Eur. J. 2012, 18, 9780−9783. (e) Seidl, M.; Stubenhofer, M.; Timoshkin, A. Y.; Scheer, M. Reaction of Pentelidene Complexes with Diazoalkanes: Stabilization of Parent 2,3-Dipnictabutadienes. Angew. Chem., Int. Ed. 2016, 55, 14037−14040. (f) Weber, L.; Meyer, M.; Stammler, H.-G.; Neumann, B. Reactivity of the Inversely Polarized Phosphaalkenes RPC(NMe2)2 (R = tBu, Me3Si, H)

towards Arylcarbene Complexes [(CO)5MC(OEt)Ar] (Ar = Ph, M = Cr, W; Ar = 2-MeC6H4, 2-MeOC6H4, M = W). Chem. - Eur. J. 2001, 7, 5401−5408. (12) Balakrishna, M. S.; Prakasha, T. K.; Krishnamurthy, S. S.; Siriwardane, U.; Hosmane, N. S. Organometallic derivatives of diphosphinoamines, X2PN(R)PX2. Reactions with carbonyl derivatives of group 6 metals and iron pentacarbonyl. The crystal structures of [Mo(CO)4PhN(P(OPh)2)2] and [W(CO)4iPrN(PPh2)2]. J. Organomet. Chem. 1990, 390, 203−216. (13) Toyota, K.; Tashiro, K.; Yoshifuji, M.; Miyahara, I.; Hayashi, A.; Hirotsu, K. X-Ray structure analyses of diphosphinidenecyclobutene and its chelate type tetracarbonylmolybdenum(0) complex. J. Organomet. Chem. 1992, 431, C39−C41. (14) (a) Andrews, G. T.; Colquhoun, I. J.; McFarlane, W. Fourier Transform Heteronuclear Magnetic Triple Resonance in Complex Spin SystemsIII: Symmetrical Ditertiary Phosphine Complexes of Group VI Metal Carbonyls. Polyhedron 1983, 2, 783−790. (b) Grim, S. O.; Wheatland, D. A. A Phosphorus-31 Nuclear Magnetic Resonance Study of Tertiary Phosphine Derivatives of Group VI Metal Carbonyls. III. Disubstituted Compounds. Inorg. Chem. 1969, 8, 1716−1719. (c) Crumbliss, A. L.; Topping, R. J. In Phosphorus-31 NMR Spectroscopy in Stereochemical Analysis; Verkade, J. G., Quin, L. D., Eds.; VCH: Hoboken, NJ, 1987; pp 531−55510.1002/ pauz.19880170607. (15) (a) Lindner, E.; Mohr, M.; Nachtigal, C.; Fawzi, R.; Henkel, G. Preparation, properties and reactions of metal-containing heterocycles: Part C: tetraazatetraphosphadimolybdacyclophanes: synthesis, isolation, characterization, and X-ray crystal structures. J. Organomet. Chem. 2000, 595, 166−177. (b) Ly, T. Q.; Slawin, A. M. Z.; Woollins, J. D. Synthesis, co-ordination chemistry and crystallographic studies of some bis(aminophosphines). J. Chem. Soc., Dalton Trans. 1997, 1611− 1616. (16) Burck, S.; Götz, K.; Kaupp, M.; Nieger, M.; Weber, J.; Schmedt auf der Günne, J.; Gudat, D. Diphosphines with Strongly Polarized P− P Bonds: Hybrids between Covalent Molecules and Donor−Acceptor Adducts with Flexible Molecular Structures. J. Am. Chem. Soc. 2009, 131, 10763−10774. (17) (a) van Koten, G.; Noltes, J. G.; Spek, A. L. Group IB Organometallic Chemistry: XXXVII. Complex Forming Reactions of Polynuclear Arylcopper Compounds: Calk−P Bond Cleavage in 1,2bis(diphenylphosphino)ethane (diphos) by (2-Me2NCH2C6H4)4Cu4 and Crystal Structure of [(C6H5)2PCu·diphos]2·2C6H6. J. Organomet. Chem. 1978, 159, 441−463. (b) Greiser, T.; Weiss, E. Darstellung und Kristallstruktur von Bis(μ-diphenylphosphido)-bis[ethylenbis(diphenylphosphin)kupfer(I)]-Benzol(1/2), [(C 6 H 5 ) 2 PCu(Ph2PCH2CH2PPh2)]2·2 C6H6. Chem. Ber. 1978, 111, 516−522. (c) Ghalib, M.; Jones, P. G.; Schulzke, C.; Sziebert, D.; Nyulászi, L.; Heinicke, J. W. π-Rich σ2P-Heterocycles: Bent η1-P- and μ2-PCoordinated 1,3-Benzazaphosphole Copper(I) Halide Complexes. Inorg. Chem. 2015, 54, 2117−2127. (d) Mankad, N. P.; Rivard, E.; Harkins, S. B.; Peters, J. C. Structural Snapshots of a Flexible Cu2P2 Core that Accommodates the Oxidation States CuICuI, Cu1.5Cu1.5, and CuIICuII. J. Am. Chem. Soc. 2005, 127, 16032−16033. (18) (a) Batsanov, S. S. Van der Waals Radii of Elements. Inorg. Mater. 2001, 37, 871−885. (b) Nag, S.; Banerjee, K.; Datta, D. Estimation of the van der Waals radii of the d-block elements using the concept of bond valence. New J. Chem. 2007, 31, 832−834. (c) Hu, S.Z.; Zhou, Z.-H.; Robertson, B. E. Consistent approaches to van der Waals radii for the metallic elements. Z. Kristallogr. 2009, 224, 375− 376. (d) Batsanov, S. S. Thermodynamic determination of van der Waals radii of metals. J. Mol. Struct. 2011, 990, 63−66. (e) Ghalib, M.; Könczöl, L.; Nyulászi, L.; Palm, G. J.; Schulzke, C.; Heinicke, J. W. πExcess Aromatic σ2-P Ligands: Synthesis and Structure of an unprecedented μ2-P-1,3-benzazaphosphole bridged Tetranuclear Copper(I) Acetate Complex. Dalton Trans. 2015, 44, 1769−1774. (f) Welsch, S.; Lescop, C.; Balazs, G.; Réau, R.; Scheer, M. Simultaneous End-On/Side-On Coordination Modes of a Diphosphorus Tetrahedral Complex Imposed by Pre-organization of Oligometallic CuI Acceptors. Chem. - Eur. J. 2011, 17, 9130−9144. G

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (g) Stamp, L.; Dieck, H. T. Copper(I) Complexes with Unsaturated Nitrogen Ligands. Part IV. Synthesis and Structure of Copper(I) Monoazadiene Complexes. Inorg. Chim. Acta 1988, 147, 199−206. (h) Hwang, W.-S.; Hwang, C.-J.; Wang, D.-L.; Lee, L.-S.; Chiang, M. Y. Synthesis and Structures of 1-(2-Thienylmethyl)-2-(2-thienyl)bezimidazole and Its Cu(I) and Ag(I) Complexes. J. Chin. Chem. Soc. 1998, 45, 555−561. (i) Oakley, S. H.; Soria, D. B.; Coles, M. P.; Hitchcock, P. B. Structural diversity in the coordination of amidines and guanidines to monovalent metal halides. Dalton Trans. 2004, 537−546. (j) Oakley, S. H.; Coles, M. P.; Hitchcock, P. B. Structural Consequences of the Prohibition of Hydrogen Bonding in Copper− Guanidine Systems. Inorg. Chem. 2004, 43, 5168−5172. (k) Samai, S.; Biradha, K. Coordination Polymers of Flexible Bis(benzimidazole) Ligand: Halogen Bridging and Metal···Arene Interactions. Cryst. Growth Des. 2011, 11, 5723−5732. (l) Jana, S.; Chattopadhyay, S. Design and construction of copper(I) complexes based on flexidentate cyclic N2-donor Schiff bases via in situ reduction of copper(II) precursors. Polyhedron 2014, 81, 298−307. (19) Boorman, P. M.; Craig, P. J.; Swaddle, T. W. Vibrational spectra of some isonitrile complexes of cobalt(I) and (II). Can. J. Chem. 1970, 48, 838−844. (20) Harris, R. K.; Becker, E. D.; Cabral de Menezes, S. M.; Goodfellow, R.; Granger, P. NMR Nomenclature: Nuclear SpinProperties and Conventions for Chemical Shifts (IUPAC Recommendations 2001). Concepts Magn. Reson. 2002, 14, 326−346. (21) Bennett, M. A.; Pratt, L.; Wilkinson, G. Transition-metal Complexes of Seven-membered Ring Systems. Part IV. Proton Resonance Spectra of Cycloheptatriene Complexes of Group VI Metals. J. Chem. Soc. 1961, 2037−2044. (22) Kubas, G. J.; Monzyk, B.; Crumbliss, A. L. Tetrakis(Acetonitrile)Copper(I) Hexafluorophosphate. Inorg. Synth. 2007, 19, 90−92. (23) A suspension of trimethylamine N-oxide (0.704 g, 9.373 mmol) in CH3CN (10 mL) was added to a suspension of W(CO)6 (1.50 g, 4.26 mmol) in CH3CN (25 mL) at room temperature. The greenishyellow clear solution was stirred for 1.5 h. The solvent was removed under vacuum, and the resulting greenish-yellow solid was washed with n-hexane (2 × 15 mL) to give [W(CO)4(CH3CN)2] as greenishyellow solid (1.34 g, yield 83%). (24) Empirical absorption correction. CrysAlis-Pro Software package; Oxford Diffraction Ltd., 2014. (25) (a) SHELX includes SHELXL-2017: Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71 (71), 3−8. (b) SHELXT: Sheldrick, G. M. SHELXT − Integrated space-group and crystal-structure determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, 71, 3−8. (26) Brandenburg, K. Diamond; Crystal Impact GbR: Bonn, Germany, 2014.

H

DOI: 10.1021/acs.inorgchem.8b00067 Inorg. Chem. XXXX, XXX, XXX−XXX