Next Article in Journal / Special Issue
Stabilization of 2-Pyridyltellurium(II) Derivatives by Oxidorhenium(V) Complexes
Previous Article in Journal
Room Temperature Surfactant-Free Synthesis of Gold Nanoparticles in Alkaline Ethylene Glycol
Previous Article in Special Issue
Coinage Metal Complexes Containing Perfluorinated Carboxylates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phosphine Functionalized CpC Ligands and Their Metal Complexes

Fachbereich Chemie, RPTU Kaiserslautern-Landau, Erwin-Schrödinger-Straße 54, 67663 Kaiserslautern, Germany
*
Author to whom correspondence should be addressed.
Chemistry 2023, 5(2), 912-933; https://doi.org/10.3390/chemistry5020062
Submission received: 15 March 2023 / Revised: 4 April 2023 / Accepted: 6 April 2023 / Published: 18 April 2023
(This article belongs to the Special Issue Commemorating 150 Years of Justus von Liebig’s Legacy)

Abstract

:
Simple nucleophilic aliphatic substitution gives access to mono- and diphosphine ligands with a CpC group in the backbone. The monophosphine ligand coordinates to gold(I) via the phosphine site, to thallium(I) via the cyclopentadienyl site and to ruthenium(II) via a combination of both, resulting in an ansa-type structure. Coordination with the cyclopentadiene site is not possible for the diphosphine ligand. In this case, monodentate coordination to gold(I) and bidentate coordination to the [PdCl(μ2-Cl)]2, the [Rh(CO)(μ2-Cl)]2, and the Rh(CO)Cl fragment is observed, showing the variability in coordination modes possible for the long-chain diphosphine ligand. Ligands and complexes were characterized by means of NMR and IR spectroscopy, elemental analysis and X-ray structure analysis.

Graphical Abstract

1. Introduction

Both cyclopentadienyl derivatives and phosphines are among the most frequently used ligands in organometallic chemistry. It is therefore not surprising that Charrier and Mathey published the first example of a phosphine-functionalized cyclopentadiene as early as 1978. They obtained it by reacting sodium cyclopentadienide and chloromethyldiphenylphosphine [1]. The combination of a cyclopentadiene-derived nucleophile with an electrophile bearing a phosphine site remains the most important route to phosphine-functionalized cyclopentadienes. Alternatively, these compounds are accessible by nucleophilic addition of a phosphide to a fulvene or to a spiro [2.4]hepta-4,6-diene derivative [2,3,4,5]. Based on these strategies, a large number of such compounds have become accessible in recent decades.
When both the cyclopentadienyl and the phosphine site coordinate to one transition metal center, the resulting compounds exhibit increased stability compared to classical complexes with unfunctionalized cyclopentadienyl ligands, because the η5→η3→η1 shift of the cyclopentadienyl ligand is suppressed by the ansa-bridging of the two donors. In addition, the phosphine donor remains in the immediate vicinity of the metal center after dissociation, which has a further stabilizing effect.
Among all phosphine functionalized cyclopentadienyl derivatives, ferrocenyl phosphines have probably attracted the most attention, in particular due to their use as mono- or bidentate ligands in transition metal catalyzed reactions [6,7,8,9,10,11,12,13,14,15,16]
Recently, we presented the C2-symmetric and thus chiral cyclopentadiene derivative (CpC)−1 (1−1) as a ligand for early and late transition metal complexes [17,18,19]. The corresponding cyclopentadiene CpCH (1) is accessible in a few steps with high yield from dibenzosuberenone. In parallel to the transition metal chemistry of compound 1, we searched for new derivatives functionalized at its cyclopentadiene core. By reacting the anion (CpC)−1 (1−1) with oxygen, the corresponding ketone is accessible in good yield. With this η4-coordinating dienone, we were able to synthesize ruthenium and iron compounds whose structures and reactivities correspond to those of the Shvo and Knölker catalysts, respectively [20]. Here, we now report on novel phosphine-functionalized derivatives of 1 and their use as ligands in transition metal chemistry.

2. Materials and Methods

Chemicals were purchased from the following suppliers: ABCR GmbH, Fisher Scientific GmbH, Merck KGaA and Strem Chemicals GmbH. All commercially available starting materials were used without further purification. Compounds sensitive to air or moisture were handled and reacted under exclusion of oxygen and water in suitable Schlenk tubes. The solvents dichloromethane, diethyl ether, n-pentane and toluene were dried with a MBraun MB-SPS solvent drying system and degassed by passing nitrogen for 10 min. Acetonitrile and tetrahydrofuran were dried according to standard methods [21]. Glassware was heated three times with a heat gun under vacuum and refilled with dry nitrogen before use. When necessary, the purification of compounds was carried out by column chromatography on the CombiFlash Rf200 instrument from Teledyne Isco with pre-packed RediSept® columns. NMR spectra were recorded using Avance 400 and 600 devices from Bruker Corporation at a temperature of 293 K (20 °C). The resonances were only partly assignable (numbering according to Scheme 1). Air- or moisture-sensitive compounds were measured under an atmosphere of nitrogen using an NMR tube with a Teflon cap from VWR International GmbH. The anhydrous deuterated solvents CD3CN, C6D6, CDCl3 and CD2Cl2 were dried according to standard methods, re-condensed and stored under an atmosphere of nitrogen in Schlenk tubes. The evaluation of the NMR spectra was carried out with the software MestReNova 6.0.2-5475 © of Mestrelab Research. The infrared spectra were recorded on a Perkin-Elmer FT-ATR-IR 100 spectrometer with an ATR cell with diamond-coated zinc selenide windows. The IR spectra were processed using the Perkin-Elmer Spectrum 6.3.5 software from and the OriginLab Corporation software OriginPro 8G. Elemental analyses were measured out in the Analytical Laboratory of the Department of Chemistry at the TU Kaiserslautern. Compounds sensitive to air or moisture were filled in a glove box into tin capsules and sealed under an atmosphere of argon. The measurement of the elemental analyses was carried with a Vario Micro Cube analyser from Elementar-Analysetechnik. CpCH (1) was synthesized according to a published procedure [17]. (tht)AuCl, TlOEt, [(η6-C6H6)RuCl2]2, (CH3CN)2PdCl2 and [(CO)2Rh(μ2-Cl)]2 were obtained commercially.
CpCHBzPPh2 (2): A total of 2.17 mL (3.47 mmol) of n-butyllithium (1.6 M solution in n-hexane) was slowly added at 0 °C to a solution of 1.24 g of CpCH (1) [17] (3.14 mmol) in 10 mL of toluene. The reaction mixture was stirred for 2 h at this temperature and then was slowly warmed to room temperature. During this time, the precipitation of a colorless solid was observed, which was dissolved by the addition of 1 mL of tetrahydrofurane. Then, 976 mg (3.14 mmol) of [2-(chloromethyl)phenyl](diphenyl)phosphane [22,23], dissolved in 10 mL of toluene, was slowly added at 0 °C. The reaction mixture was stirred for additional 2 h at this temperature. During this time, a color change from orange to dark red was observed and the reaction mixture was stirred for another 18 h at room temperature. A total of 15 mL of water was added, and the organic phase was separated, washed three times with 5 mL of water and dried over MgSO4. The solvent was removed, and the crude red–brown product was dried under vacuum. Purification was carried out by column chromatography (MPLC) with a mixture of n-hexane and ethyl acetate (99:1). Compound 2 was obtained slightly contaminated as a light pink solid in the first fraction. As a minor product (16%), compound 3 can be isolated from the second fraction. Compound 2 was finally purified by precipitation from a saturated dichloromethane solution with n-pentane. Yield: 1.23 g (59%) of a colorless solid. Elemental analysis calcd. for C50H37P (668.82 g/mol): C 89.79, H 5.58; found: C 89.87, H 5.70%. 1H NMR (400 MHz, CDCl3): δ 7.55 (d, 3JHH = 7.9 Hz, 1H, HAr), 7.34–7.28 (m, 2H, HAr), 7.25–7.05 (m, 15H, HAr), 6.98 (dt, 3JHH = 15.5, 7.2 Hz, 5H, HAr), 6.89–6.84 (m, 1H, HAr), 6.84–6.79 (m, 2H, HAr), 6.76 (t, 3JHH = 6.9 Hz, 2H, HAr), 6.60 (d, 3JHH = 7.1 Hz, 2H, HAr), 4.69 (t, 3JHH = 5.1 Hz, 1H, H1), 4.07 (dt, 2JHH = 16.7, 3JHH = 3.3 Hz, 4JPH = 3.3 Hz, 1H, H7), 3.80 (d, 2JHH = 12.7 Hz, 1H, H6 or H6′), 3.75–3.63 (m, 3H, H6 or H6′ and H7), 3.33 (d, 3JHH = 12.6 Hz, 1H, H6 or H6′). 13C NMR (101 MHz, CDCl3): δ 147.9, 144.9, 143.9, 143.1, 142.9, 141.6, 141.0, 139.8, 139.6, 136.8 (d, JPC = 7.4 Hz), 136.7 (d, JPC = 7.8 Hz), 136.2, 136.0, 134.2, 134.2, 134.1, 134.0, 133.9, 133.1, 132.9, 132.66, 132.1, 129.2, 128.9, 128.9, 128.8, 128.7, 128.6, 128.0, 128.0, 127.9, 127.4, 127.0, 126.6, 126.5, 126.3, 125.1, 124.8, 55.1 (C1), 42.2 (C6 or C6′), 41.8 (C6 or C6′), 34.6 (d, 3JPC = 24.1 Hz, C7). 31P NMR (162 MHz, CDCl3): δ −15.44 (s). IR (ATR, cm−1): ṽ 3059 m, 2962 m, 2921 m, 2872 w, 1584 m, 1485 m, 1464 m, 1432 s, 1333 m, 1304 m, 1277 m, 1213 m, 1193 m, 1144 m, 1124 m, 1090 m, 1027 m, 940 m, 779 m, 765 s, 746 vs, 717 s, 697 vs, 674 s.
CpC(BzPPh2)2 (3): A total of 0.74 mL (1.18 mmol) of n-butyllithium (1.6 M in n-hexane) wasadded at 0 °C to a solution of 716 mg (1.07 mmol) of 2 in (10 mL) of toluene. The reaction mixture was stirred for 2 h at this temperature and afterwards slowly warmed to room temperature. During this time, the precipitation of a colorless solid was observed, which was dissolved by the addition of 1 mL of tetrahydrofurane. Then, 333 mg (1.07 mmol) of [2-(chloromethyl)phenyl](diphenyl)phosphane [22,23] dissolved in 5 mL of toluene wasadded slowly at 0 °C and the reaction mixture was stirred for another 2 h at this temperature, whereby its color changed from pink to dark red. After warming to room temperature, the mixture was stirred for further 18 h. A total of 5 mL of water was added, the organic phase was separated, washed three times with 5 mL of water, and then dried over MgSO4. The solvent was removed, and the light pink crude product was dried under vacuum. Finally, the crude product was purified by precipitation from a saturated dichloromethane solution with n-pentane. Colorless prismatic single crystals were obtained by slow diffusion of n-pentane into a saturated toluene solution. Yield: 840 mg (83%) of a colorless solid. Elemental analysis calcd. For C69H52P2 (943.12 g/mol): C 87.87, H 5.56; found: C 87.86, H 5.9%. 1H NMR (400 MHz, CDCl3): δ 7.81 (d, 3JHH = 7.9 Hz, 2H, HAr), 7.42–7.30 (m, 18H, HAr), 7.25 (dd, 3JHH = 10.3, 5.5 Hz, 4H, HAr), 7.17 (t, 3 3JHH = 7.4 Hz, 2H, HAr), 7.09 (dd, 3JHH = 10.6, 4.4 Hz, 4H, HAr), 7.02–6.94 (m, 6H, HAr), 6.90 (t, 3JHH = 7.6 Hz, 2H, HAr), 6.85 (t, 3JHH = 7.6 Hz, 2H, HAr), 6.71 (dd, 3JHH = 7.3, 3.8 Hz, 2H, HAr), 6.50 (d, 3JHH = 7.7 Hz, 2H, HAr), 4.33 (dd, 2JHH = 16.1, 4JPH = 8.0 Hz, 2H, H7), 3.84 (d, 2JHH = 16.1 Hz, 2H, H7), 3.39 (d, 2JHH = 12.5 Hz, 2H, H6), 2.48 (d, 2JHH = 12.4 Hz, 2H, H6). 13 C NMR (101 MHz, CDCl3): δ 147.6, 145.0, 142.4, 141.7, 141.5, 141.1, 137.8, 137.7, 137.3, 137.2, 137.0, 134.4 (d, JPC = 20.2 Hz), 133.9 (d, JPC = 20.1 Hz), 133.3, 133.0, 132.3, 128.9, 128.7, 128.5 (d, JPC = 19.4 Hz), 128.3, 128.1, 127.8, 127.5 (d, JPC = 2.1 Hz), 126.8, 126.6, 126.5, 126.1, 124.7, 65.6 (t, 4JPC = 2.8 Hz, C1), 40.9 (br, C6), 40.1 (dt, 3JPC = 10.7, 5.7 Hz, C7). 31P NMR (162 MHz, CDCl3): δ −16.22 (s). IR (ATR, cm−1): ṽ = 3053 m, 3023 w, 2999 w, 2927 w, 2905 w, 2867 w, 2843 w, 1586 m, 1480 m, 1461 m, 1434 s, 1264 w, 1157 w, 1120 w, 1092 m, 1065 m, 1026 m, 947 w, 904 m, 754 s, 738 vs, 694 vs, 680 m, 656 m.
[CpCHBzPPh2(AuCl)] (4): The reaction was carried out in the absence of light. A total of 222 mg (0.33 mmol) of compound 2 and 136 mg (0.42 mmol) of (tetrahydrothiophene)gold(I) chloride was stirred for 18 h at room temperature in 15 mL of dichloromethane. A total of 30 mL of n-pentane was added to precipitate a colorless solid, which was filtered off, washed three times with 15 mL of diethyl ether and three times with 15 mL of n-pentane, and finally, dried under vacuum. Yield: 298 mg (78%) of a colorless solid. Prismatic single crystals containing one equivalent of toluene were obtained by slow diffusion of n-pentane into a saturated toluene solution. Elemental analysis calcd. for C50H37AuClP (901.22 g/mol): C 66.64, H 4.14, found: C 66.53, H 4.10%. 1H NMR (400 MHz, CDCl3): δ 7.66–7.57 (m, 3H, HAr), 7.55–7.48 (m, 2 × 3JHH = 8.8 and 7.6 Hz, 1.9 Hz, 3H, HAr), 7.42–7.27 (m, 9H, HAr), 7.25–7.14 (m, 6H, HAr), 7.11–6.92 (m, 6H, HAr), 6.91–6.85 (m, 1H, HAr), 6.75–6.67 (m, 2H, HAr), 4.65 (t, 3JHH = 4.6 Hz, 1H, H1), 4.51 (dd, 2JHH = 18.3 Hz, 4.7 Hz, 1H, H7), 4.05 (dd, 2JHH = 18.7 Hz, 3.5 Hz, 1H, H7), 3.86 (d, 2JHH = 12.7 Hz, 1H, H6 or H6′), 3.79 (dd, 2JHH = 12.6 Hz, 7.2 Hz, 2H, H6 or H6′), 3.60 (d, 2JHH = 12.5 Hz, 1H, H6 or H6′). 13C NMR (101 MHz, CDCl3): δ 146.1, 144.2, 143.5, 143.3, 143.2, 142,0 141.6, 140.7, 139.7 (d, JPC = 5.1 Hz), 135.2 (d, JPC = 14.0 Hz), 134.2 (d, JPC = 13.8 Hz), 133.4, 132.4 (d, JPC = 2.5 Hz), 132.3, 132.2, 132.1 (d, JPC = 1.9 Hz), 132.1, 131.8, 129.6 (d, JPC = 12.0 Hz), 129.3, 129.2, 129.0, 129.0, 128.9, 128.4, 128.1, 128.0, 128.0, 127.9, 127.9, 127.7, 127.7, 127.5, 127.4, 127.0, 126.9, 126.9, 126.8, 126.7, 126.6 (d, JPC = 9.8 Hz), 126.2, 125.0 (d, JPC = 10.0 Hz), 53.8 (C1), 41.9 (C6 or C6′), 41.8 (C6 or C6′), 32.4 (d, JPC = 15.8 Hz, C7). 31P NMR (162 MHz, CDCl3): δ 24.4. IR (ATR, cm−1): ṽ 3054 w, 3014 w, 2980 m, 2886 w, 2831 w, 1589 w, 1480 m, 1436 s, 1351 w, 1250 w, 1185 w, 1158 w, 1121 w, 1099 m, 1039 w, 998 w, 945 w, 908 w, 771 s, 744 vs, 711 s, 691 vs.
[(η5-CpCHBzPPh2)Tl] (5): The reaction was carried out in the absence of light. A total of 1.33 g (1.99 mmol) of 2 was dissolved in 30 mL o dry benzene. A total of 874 mg (350 mmol) of thallium(I) ethanolate was added and the mixture was stirred for 30 min at room temperature. A bright yellow solid precipitated, which was filtered off, washed three times with 40 mL of benzene and dried under vacuum. Yield 1.44 g (83%) of a bright yellow microcrystalline solid, which was recrystallized by cooling a hot saturated solution in THF to room temperature. Elemental analysis calcd. for C50H36PTl (872.18 g/mol): C 68.85, H 4.16; found: C 68.82, H 4.26%. 1H NMR (600 MHz, DMSO-d6): δ 7.58–7.53 (m, 1H, HAr), 7.46–7.38 (m, 6H, HAr), 7.30 (d, 3JHH = 7.2 Hz, 3H, HAr), 7.26 (t, 3JHH = 7.1 Hz, 3H HAr), 7.17 (t, 3JHH = 7.2 Hz, 2H, HAr), 7.15–7.12 (m, 1H, HAr), 7.10–7.05 (m, 2H, HAr), 6.98–6.83 (m, 9H, HAr), 6.78–6.74 (m, 2H, HAr), 6.61 (dd, 2 × 3JHH = 6.9, 4.8 Hz, 1H, HAr), 5.05 (d, 2JHH = 18.4 Hz, 1H, H7), 4.39 (d, 2JHH = 18.4 Hz, 1H, H7), 4.14 (d, 3JHH = 11.6 Hz, 1H, H6 or H6′), 3.83 (br, 2H, H6 or H6′), 3.74 (d, 3JHH = 10.8 Hz, 1H, H6 or H6′). 13C NMR (151 MHz, DMSO-d6): δ 148.0, 147.8, 139.8, 139.36, 138.6, 136.4, 136.3, 136.2, 135.6, 134.4, 134.3, 133.6 (d, JPC = 19.7 Hz), 133.4 (d, JPC = 19.5 Hz), 131.8, 130.9, 130.2, 128.9, 128.8, 128.8, 128.8, 128.7, 128.5, 128.2 (d, JPC = 3.9 Hz), 128.0, 127.6, 127.1, 126.9, 126.3, 125.4, 125.1, 124.8, 124.5, 124.4, 124.4, 123.8, 123.8, 123.7, 123.5, 123.0, 122.8, 121.1, 118.7, 115.5 (d, JPC = 3.7 Hz, C1), 43.1 (C6 or C6′), 41.3 (C6 or C6′), 30.1 (d, JPC = 24.3 Hz, C7). 31P NMR (243 MHz, DMSO-d6): δ −15.6.
[(η5-CpCBzPPh2)Ru(NCMe)2]PF6 (6): A total of 579 mg (0.66 mmol) of 5, 150 mg (0.30 mmol) of [(η6-C6H6)RuCl2]2 and 140 mg (0.76 mmol) of KPF6 were dissolved in 20 mL of acetonitrile and was stirred for 18 h at room temperature, while the color of the mixture changed from yellow to orange. The precipitated solid (TlCl, KCl) was filtered off and the volume of the solution was reduced to about 5 mL by removing the solvent under vacuum. By the addition of 30 mL of diethyl ether, an orange colored solid precipitated, which was filtered off, washed three times with 15 mL of diethyl ether and three times with 15 mL of n-pentane, and finally, dried under vacuum. Yield: 530 mg (80%) of an orange colored solid. Elemental analysis calcd. for C54H42F6N2P2Ru (1036.98 g/mol): C 65.12, H 4.25, N 2.81; found: C 64.89, H 4.48, N 2.74.1H NMR (400 MHz, CD3CN): δ 7.63–7.58 (m, 2H, HAr), 7.52–7.49 (m, 3H, HAr), 7.48–7.42 (m, 3H, HAr), 7.42–7.34 (m, 4H, HAr), 7.31 (d, 3JHH = 7.7 Hz, 1H, HAr), 7.27–7.21 (m, 3H, HAr), 7.19–7.13 (m, 3H, HAr), 7.11–6.99 (m, 6H, HAr), 6.96–6.91 (m, 1H, HAr), 6.90–6.82 (m, 2H, HAr), 6.63 (td, 3JHH = 7.7, 4JHH = 0.8 Hz, 1H, HAr), 6.17 (d, 3JHH = 7.7 Hz, 1H, HAr), 4.47 (d, 2JHH = 13.1 Hz, 1H, H6 or H6′), 4.38 (dd, 2JHH = 18.0, 4 J PH = 6.8 Hz, 1H, H7), 4.18 (d, 2JHH = 12.8 Hz, 1H, H6 or H6′), 4.12 (d, 2JHH = 17.9 Hz, 1H, H7), 3.78 (t, 2JHH = 13.2 Hz, 2H, H6 or H6′), 1.96 (s, 6H, CH3CN). 13C NMR (101 MHz, CD3CN): δ = 147.0, 146.8, 144.5, 143.9, 143.7 (d, JPC = 2.5 Hz), 142.1, 135.1 (d, JPC = 3.2 Hz), 134.9, 134.8 (d, JPC = 3.2 Hz), 134.4, 133.8 (d, JPC = 10.1 Hz), 133.2, 132.9 (d, JPC = 10.7 Hz), 132.6, 132.5 (d, JPC = 17.0 Hz), 132.0 (d, JPC = 1.8 Hz), 131.9 (d, JPC = 1.8 Hz), 131.2, 131.1 (d, JPC = 1.9 Hz), 130.7 (d, JPC = 1.7 Hz), 130.5, 130.3, 130.2, 129.5, 129.4, 129.3, 129.2, 129.1, 128.9, 128.5 (d, JPC = 8.7 Hz), 128.4, 128.3, 128.3, 127.1, 126.5, 126.5, 126.3 (d, JPC = 9.0 Hz), 126.0, 105.6 (d, JPC = 3.5 Hz, C1), 92.4 (d, JPC = 9.6 Hz, C2-C5), 92.0 (C2-C5), 87.8 (C2-C5), 75.0 (C2-C5), 41.9 (C6 or C6′), 41.0 (C6 or C6′), 30.6 (d, JPC = 5.3 Hz, C7). 31P NMR (162 MHz, CD3CN): δ 51.7 (s, PPh2), −144.6 (hept., 1JPF = 706.3 Hz, PF6-). 19F NMR (376 MHz, CD3CN): δ −72.9 (d, 1JPF = 706.1 Hz, PF6-). ESI-MS (CD3CN): m/z found (calcd.) 810.13 (810.19, [C50H36PRu(CH3CN)]+), 769.09 (769.16, [C50H36PRu]+). IR (ATR, cm−1): ṽ 3056 w, 3005 w, 2963 w, 1926 w, 2853 w, 2038 vw, 1599 w, 1483 m, 1435 s, 1371 w, 1261 w, 1162 w, 1093 m, 1039 w, 1028 w, 950 w, 919 w, 876 m, 834 vs, 754 s, 747 s, 738 s, 719 m, 696 s, 674 m.
[CpC(BzPPh2(AuCl)2)] (7): The reaction was carried analogous to the synthesis of the gold(I) complex 4 using 239 mg (0.25 mmol) of the diphosphine derivative 3 and 171 mg (0.53 mmol) of (tht)AuCl. After stirring the mixture for 18 h at room temperature in 15 mL of CH2Cl2, the colorless precipitate was filtered of, washed four times with 15 mL of diethyl ether and then dried under vacuum. Prismatic single crystals containing one equivalent of toluene were obtained by slow diffusion of n-pentane into a saturated toluene solution. Yield: 313 mg (83%) of a colorless solid. Due to the poor solubility of compound 7 in all organic solvents, it was impossible to obtain a 13C NMR spectrum. Elemental analysis calcd. for C69H52Au2Cl2P2 (1407.94 g/mol): C 58.86, H 3.72; found: C 58.62, H 3.83%. 1H NMR (400 MHz, CDCl3): δ 7.61–7.56 (m, 3JHH = 7.0 Hz, 2H, HAr), 7.52–7.45 (m, 8H, HAr), 7.40–7.27 (m, 20H, HAr), 7.22–7.17 (m, 3JHH = 9.6, 3.9 Hz, 4H, HAr), 7.08–7.04 (m, 4H, HAr), 6.91–6.83 (m, 4H, HAr), 6.60 (d, 3JHH = 7.6 Hz, 2H, HAr), 4.01 (d, 2JHH = 18.2 Hz, 2H, H7 or H7′), 3.65 (d, 2JHH = 17.1 Hz, 2H, H7 or H7′), 3.51 (d, 2JHH = 12.6 Hz, 2H, H6 or H6′), 2.89 (d, 2JHH = 12.5 Hz, 2H, H6 or H6′). 31P NMR (162 MHz, CDCl3): δ 24.9. IR (ATR, cm−1): ṽ 3058 w, 3015 w, 2948 w, 2925 w, 2842 w, 1588 w, 1480 m, 1438 s, 1310 w, 1261 m, 1179 w, 1101 s, 1068 m, 1029 m, 999 m, 903 w, 847 w, 793 m, 757 vs, 743 vs, 689 m.
[CpC(BzPPh2)2(PdCl(μ2-Cl))2] (8): 183 mg (0.19 mmol) of ligand 3 and 101 mg (0.39 mmol) of (CH3CN)2Pd(Cl)2 were dissolved in 6 mL of CH2Cl2. The mixture was stirred for 18 h at room temp. Then, 10 mL of n-pentane were added. An orange colored solid precipitated, which was filtered off, washed three times with 5 mL of diethyl ether and three times with 5 mL of n-pentane and dried under vacuum. Yield: 197 mg (78%) of a temperature-sensitive orange colored solid, which was stored at T < 10 °C. Single crystals were obtained by slow diffusion of n-pentane into a saturated solution of the compound in toluene. Elemental analysis calcd. for C69H52Cl4P2Pd2 (1297.77 g/mol): C 63.86, H 4.04; found: C 63.56, H 3.93%. 1H NMR (400 MHz, CDCl3): δ 7.87 (dd, 3JHH = 12.5, 7.6 Hz, 4H, HAr), 7.83–7.73 (m, 4H, 2 × HAr and 2×H7 or H7′), 7.60 (d, 3JHH = 8.2 Hz, 2H, HAr), 7.55–7.50 (m, 2H, HAr), 7.45 (d, 3JHH = 7.5 Hz, 2H, HAr), 7.42–7.37 (m, 6H, HAr), 7.33–7.29 (m, 2H, HAr), 7.25–7.21 (m, 2H, HAr), 7.16–7.07 (m, 4H, HAr), 7.05–7.00 (m, 2H, HAr), 6.98–6.88 (m, 9H, HAr), 6.85 (dd, 3JHH = 6.1, 0.7 Hz, 1H, HAr), 6.83–6.78 (m, 2H, HAr), 6.62 (dd, 3JHH = 12.5, 7.9 Hz, 4H, HAr), 4.80 (d, 2JHH = 17.7 Hz, 2H, H7 or H7′), 3.89 (d, 2JHH = 12.3 Hz, 2H, H6 or H6′), 3.82 (d, 2JHH = 12.5 Hz, 2H, H6 or H6′). 13C NMR (101 MHz, CDCl3): δ 147.9, 144.3, 141.6 (d, JPC = 10.2 Hz), 141.0, 140.6, 137.8 (d, JPC = 11.6 Hz), 133.0 (d, JPC = 9.6 Hz), 132.3, 131.8, 130.4, 129.1, 128.8, 128.7, 128.6, 128.5, 128.4, 128.3, 128.0, 127.7, 127.4, 127.3, 127.2, 126.9 (d, JPC = 21.3 Hz), 126.7, 126.6, 126.4, 126.3 (d, JPC = 3.5 Hz), 125.1, 63.8 (C1), 48.2 (d, JPC = 11.4 Hz, C7 or C7′), 41.88 (C6 or C6′). 31P NMR (162 MHz, CDCl3): δ 25.4 (s). IR (ATR, cm−1): ṽ 3056 m, 2945 w, 2892 w, 2836 w, 1589 w, 1480 m, 1461 w, 1435 s, 1327 w, 1188 w, 1161 w, 1092 s, 999 w, 902 w, 850 w, 757 s, 741 s, 706 s, 690 s.
[CpC(BzPPh2)2)(Rh(μ2-Cl(CO))2] (9): A total of 107 mg (0.11 mmol) of ligand 3 and 45.3 mg (0.11 mmol) of [(CO)2Rh(μ2-Cl)]2 was dissolved in 6 mL of dichloromethane. The mixture was stirred for 2 h at room temperature. Then, 10 mL of n-pentane was added. A bright yellow solid precipitated, which was filtered off, washed three times with 5 mL of n-pentane and dried under vacuum. Yield: 122 mg (85%) of a bright yellow microcrystalline solid, which was stored at temperatures below 10 °C to prevent decomposition. Yellow prismatic single crystals were obtained by slow diffusion of n-pentane into a saturated solution of the compound in toluene. Elemental analysis calcd. for C71H52Cl2O2P2Rh2 (1275.84 g/mol): C 66.84, H 4.11; found: C 67.10, H 4.42%. 1H NMR (600 MHz, CDCl3): δ 8.47 (d, 3JHH = 7.8 Hz, 1H, HAr), 7.80–7.70 (m, 2H, HAr), 7.65 (dd, 3JHH = 7.0, 3JHH = 5.6 Hz, 1H, HAr), 7.55–7.51 (m, 3JHH = 11.1, 3JHH = 4.1 Hz, 1H, HAr), 7.47–7.34 (m, 12H, HAr), 7.32 (t, 3JHH = 7.5 Hz, 1H, HAr), 7.31–7.25 (m, 6H, HAr), 7.24–7.21 (m, 2H, HAr), 7.18 (t, 3JHH = 7.7 Hz, 1H, HAr), 7.12–7.04 (m, 5H, 3 × HAr and 2 × H7 or H7′), 7.02–6.95 (m, 5H, HAr), 6.90 (d, 3JHH = 7.6 Hz, 1H, HAr), 6.89–6.82 (m, 3JHH = 8.3 Hz, 2H, HAr), 6.81 (d, 3JHH = 7.9 Hz, 1H, HAr), 6.71 (dd, 3JHH = 11.8, 3JHH = 7.7 Hz, 1H, HAr), 6.69–6.61 (m, 4H, HAr), 4.03 (d, 2JHH = 12.7 Hz, 1H, H6 or H6′), 3.87 (d, 2JHH = 12.8 Hz, 1H, H6 or H6′), 3.77–3.69 (m, 2H, H6 or H6′ and H7 or H7′), 3.64 (d, 2JHH = 12.7 Hz, 1H, H6 or H6′), 3.45 (d, 2JHH = 18.9 Hz, 1H, H7 + H7′). 13C NMR (151 MHz, CDCl3): δ 183.2 (dd, 1JRhC = 78.0, 2JPC = 20.6 Hz, CO), 182.3 (dd, 1JRhC = 78.8, 2JPC = 21.1 Hz, CO), 148.0 (d, JPC = 38.4 Hz), 144.6 (d, JPC = 32.7 Hz), 142.7 (d, JPC = 10.0 Hz), 142.2 (d, JPC = 9.6 Hz), 141.9, 141.1, 141.0, 140.7, 136.5 (b), 135.6 (d, JPC = 11.6 Hz), 133.8 (d, JPC = 10.3 Hz), 132.8, 132.6, 132.5, 132.5, 132.2, 132.1, 132.0, 132.0, 131.8, 131.5, 131.0 (d, JPC = 1.5 Hz), 131.0, 130.9, 130.8 (d, JPC = 1.8 Hz), 130.2, 130.0, 129.9, 129.4, 129.2, 129.1, 128.9, 128.88, 128.8, 128.8, 128.6, 128.5, 128.3, 128.2, 128.1, 128.0, 127.8 (d, JPC = 4.3 Hz), 127.7 (d, JPC = 4.2 Hz), 127.5, 127.5, 127.4, 126.9, 126.7, 126.7, 126.1, 126.0, 126.0, 125.7, 125.6, 125.3, 124.9, 63.5 (C1), 46.5 (d, JPC = 15.5 Hz, C7 or C7′), 44.9 (d, JPC = 13.8 Hz, C7 or C7′), 42.1 (C6 or C6′), 41.7 (C6 or C6′). 31P NMR (162 MHz, CDCl3): δ 39.8 (d, 1JRhP = 171.5 Hz), 35.9 (d, 1JRhP = 175.0 Hz). IR (ATR, cm−1): ṽ 3055 w, 3017 w, 2948 w, 2895 w, 2863 w, 2839 w, 1991 vs, 1975 vs, 1590 w, 1482 m, 1462 w, 1435 s, 1327 w, 1310 w, 1187 m, 1160 m, 1119 m, 1093 s, 999 m, 902 m, 855 w, 764 m, 753 s, 737 vs, 720 m, 690 vs, 677 s, 662 m.
[CpC(BzPPh2)2(RhCl(CO))] (10): A total of 217 mg (0.23 mmol) of ligand 3 and 46.1 mg (0.12 mmol) of [(CO)2Rh(μ2-Cl)]2 were dissolved in 10 mL of dichloromethane. The mixture was stirred for 18 h at 45 °C. Then, 10 mL of n-pentane was added. A shiny yellow solid precipitated, which was filtered off, was washed three times with 5 mL of diethyl ether and three times with 5 mL of n-pentane and was dried under vacuum. Yield: 157 mg (62%) of a shiny yellow solid. Yellow prismatic single crystals were obtained by slow diffusion of n-pentane into a saturated solution of the compound in toluene. Elemental analysis calcd. for C70H52ClOP2Rh (1109.49 g/mol): C 75.78, H 4.72; found: C 75.48, H 4.68. 1H NMR (400 MHz, CDCl3): δ 8.11–8.05 (m, 2H, HAr), 7.87 (s, 1H, HAr), 7.58–7.52 (m, 2H, HAr), 7.48–7.44 (m, 4H, HAr), 7.41–7.28 (m, 9H, HAr), 7.24–7.18 (m, 3H, HAr), 7.17–7.05 (m, 5H, HAr), 7.04–6.86 (m, 7H, HAr), 6.85–6.71 (m, 5H, HAr), 6.65–6.59 (m, 2H, HAr), 6.57–6.48 (m, 4H, HAr), 6.23 (d, 2JHH = 16.8 Hz, 1H, H7 or H7′), 5.56 (d, 2JHH = 16.7 Hz, 1H, H7 or H7′), 4.94 (d, 3JHH = 13.5 Hz, 1H, H7 or H7′), 4.09 (q, 3JHH = 12.3 Hz, 2H, H6 or H6′), 3.55 (d, 2JHH = 12.4 Hz, 1H, H7 or H7′), 3.37 (d, 2JHH = 12.4 Hz, 2H, 1 × H6 or H6′ and 1 × H7 or H7′). 13C NMR (101 MHz, CDCl3): δ 190.0 (d, 1JRhC = 78.5 Hz), 149.3, 147.7, 143.7 (d, xxJPC = 15.8 Hz), 143.6 (d, JPC = 3.5 Hz), 142.3, 141.2, 140.8, 139.8, 138.0, 137.9, 137.6, 137.2, 136.8, 136.0 (d, JPC = 11.9 Hz), 135.4, 135.1, 135.0 (d, JPC = 1.8 Hz), 134.5, 133.9 (d, JPC = 3.7 Hz), 133.7 (d, JPC = 2.6 Hz), 133.3, 133.2, 132.9, 132.8 (d, JPC = 2.0 Hz), 132.4 (d, JPC = 1.6 Hz), 132.2, 132.1, 132.0, 131.9, 131.4 (d, JPC = 3.8 Hz), 131.0, 130.9, 130.5, 130.2 (d, JPC = 4.0 Hz), 129.8 (d, JPC = 1.9 Hz), 129.2, 128.7, 128.6, 128.5 (d, JPC = 2.5 Hz), 128.4, 128.1, 128.0, 128.0, 127.8, 127.6, 127.6, 127.2, 127.1, 126.9 (d, JPC = 2.2 Hz), 126.5, 126.3, 126.2, 126.1, 125.7, 125.0, 124.8, 66.9 (C1), 49.3 (C7 or C7′), 46.8 (d, JPC = 8.7 Hz, C7 or C7′), 42.9 (C6 or C6′), 41.8 (C6 or C6′). 31P NMR (162 MHz, CDCl3): δ 35.6 (dd, 2JPP = 340.2, 1JRhP = 127.5 Hz), 26.2 (dd, 2JPP = 339.7, 1JRhP = 123.2 Hz). ESI-MS (CD2Cl2): m/z found (calcd.) 1073.39 (1073.25, [C70H52OP2Rh]+). IR (ATR, cm−1): ṽ 3056 w, 3010 w, 2947 w, 2837 w, 1961 vs, 1587 w, 1573 w, 1480 m, 1466 m, 1449 m, 1434 s, 1339 w, 1312 w, 1279 w, 1187 m, 1157 w, 1120 w, 1090 s, 1044 w, 1028 w, 999 m, 947 w, 908 w, 872 w, 841 w, 781 m, 748 vs, 690 vs, 673 m.
X-rax structure analysis: Crystal data and refinement parameters were collected and are presented in Table 1 and Table 2. All structures were solved using direct method of SIR92 [24], completed by subsequent difference Fourier syntheses, and refined by full-matrix least-squares procedures [25]. Semi-empirical absorption correction from equivalents (Multiscan) was applied to ligand 3, while analytical numeric absorption correction was carried out to all other complexes [26]. All non-hydrogen atoms were refined with anisotropic displacement parameters. All hydrogen atoms were placed in calculated positions and refined using a riding model. In the original solved structures of complexes 5, 7, 8 and 9, severely disordered and/or partially occupied solvents were also found. The quality of the measured samples and the corresponding collected raw data were limited. Even with the help of lots of restraints, these disorders could not be treated satisfyingly. To obtain a better understanding of the main structures, the SQUEEZE process integrated in PLATON was used. Additionally, detailed information has been posted in the corresponding final CIF files. CCDC 2247305-2247311 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif (accessed on 4 April 2023).

3. Results

3.1. Ligand Synthesis and Characterization

Phosphine functionalized derivatives of CpCH (1) were obtained by the deprotonation of the compound with n-butyl lithium and treatment of the intermediately generated LiCpC with 2-[(chloromethyl)phenyl](diphenyl)phosphine (Scheme 2) [22,23] The process follows a protocol published by Li et al. in 2019 for the synthesis of a phosphine-functionalized tetraethylcyclopentadiene [27]. Astonishingly, there are almost no other reports on cyclopentadienyl compounds with a 2-diphenylphosphanylphenylmethyl unit. Knochel et al. reported the synthesis of chiral 1-(SFc)-diphenylphosphanyl-2-(o-diphenylphosphanylphenylmethyl)ferrocene via a multi-step process starting with the Friedel–Crafts acylation of ferrocene with 2-bromobenzoyl chloride [28].
Reacting the substrates in a 1:1 ratio with a slight excess of n-butyl lithium allowed us to isolate the mono-substituted derivative CpCHBzPPh2 (2) in 59% yield after purification by column chromatography. As a minor product, the bi-functionalized compound CpC(BzPPh2)2 (3) was obtained in 16% yield. This speaks for a protonation/deprotonation equilibrium between CpCH (1), LiCpC, CpCHBzPPh2 (2) and LiCpCBzPPh2 and a rather high reactivity of LiCpCBzPPh2 towards the electrophile. Accordingly, the treatment of phosphine 2 with another equivalent of n-butyl lithium and the electrophile gave the diphosphine 3 in 83% yield.
While the 31P NMR spectra of 2 and 3 show rather similar chemical shifts of expected values [27,29], their 1H and 13C NMR spectra differ largely according to the symmetry of the compounds. In the 1H NMR spectrum of C2-symmetric 3, there are four doublets in the aliphatic region, which are assigned to the diastereotopic protons of the two chemically inequivalent methylene groups of the molecule. In contrast, the 1H and 13C NMR spectra of C1-symmetric 2 are more complicated. In the aliphatic region of the 1H NMR spectrum, for example, there are four resonances for the two, now magnetically inequivalent methylene groups bridging the phenyl groups in the “wings” of the Cp backbone. Three further resonances (ABB’ spin system) are assigned to the proton at the Cp ring and the protons of the methylene group linking the Cp ring and the triphenyl phosphine moiety. These methylene protons can easily be identified by their coupling pattern, including a small long-range coupling to the phosphorous atom (4JPH = 3.3 Hz) [29].
In the literature, there are only a few reports of 1,1-diphosphine-functionalized cyclopentadiene derivatives. In 1965, Keough and Grayson succeeded in adding fluorene to vinylphosphonium salts, resulting in dicationic phosphonium species [30]. Inagaki et al. published fluorene derivatives with two phosphinyl moieties directly bound to the sp3 carbon atom of the fluorene backbone [31]. In addition, there are two patents dealing with such compounds [32,33].
By slow diffusion of n-pentane into a saturated toluene solution of compound 3, single crystals suitable for an X-ray structure analysis could be obtained. Compound 3 crystallizes as colorless prisms of the triclinic space group P 1 ¯ with two equivalents of toluene in the unit cell. The two phosphine units are found arranged between the wings of the ligand, in an almost orthogonal position to the central cyclopentadiene ring. Figure 1 shows the structure of 3 and summarizes relevant structural parameters.

3.2. Transition Metal Complexes

While ligand 3 is expected to act like a typical diphosphine, the coordination chemistry of mono-phosphine 2 allows more variations: it may either act as a typical monodentate phosphine, as an η5-coordinating Cp-type ligand or may exhibit a combination of these two modes of metal–ligand interactions. To approve the first type of coordination, ligand 2 was treated with [(tht)AuCl] (tht = tetrahydrothiophene, Scheme 3). Gold(I) prefers linear coordination, which avoids involving the Cp ring of 2 in its coordination environment. The according gold(I) complex 4 was obtained in 78% yield as a colorless solid, which is stable under ambient conditions.
While for typical 31P NMR resonances of triarylposphine gold(I) chloride, complexes chemical shift of around 32 ppm are reported [34,35], the 31P NMR resonance of compound 4 appears at 24.4 ppm. We assign this to some shielding effects of the aromatic environment of the CpC backbone. The coordination of the gold(I) site in addition leads to a better separation of the aliphatic 1H NMR resonances compared to the metal-free ligand 2 (see the Supplementary Materials). Here, a triplet at 4.65 ppm (3JHH = 4.6 Hz) is assigned to the proton located at the five-membered Cp ring.
By the slow diffusion of n-pentane into a saturated toluene solution of compound 4, single crystals suitable for an X-ray structure analysis could be obtained. Compound 4 crystallizes as colorless prisms of the triclinic space group P 1 ¯ with one equivalent of toluene in the unit cell. Figure 2 shows the structure of 4 and summarizes relevant structural parameters.
The crystal structure complex 4 shows the typical linear coordination geometry of gold(I) with an the P1-Au1-Cl1 angle of 179.42(4)°. The measured P1-Au1 (2.2427(10) Å) and Au1-Cl1 (2.2865(10) Å) distances are comparable to the data of (triphenylphosphine)gold(I) chloride and related gold(I) complexes [34,36]. In contrast to many other gold(I) complexes [37,38,39], there is no hint for a d10–d10 interaction between the gold(I) sites of neighboring molecules. The phosphine substituent occupies an almost orthogonal position to the central five-membered ring of the CpC backbone, minimizing steric repulsions.
We recently published the synthesis of a series of late transition metal complexes wherein TlCpC is used as the transferring reagent for the [CpC] ligand [19]. To achieve similar chemistry with a phosphine-functionalized CpC derivative, compound 2 was deprotonated in dry benzene with stoichiometric amounts of thallium(I) ethanolate in the absence of light (Scheme 4).
The thallium(I) compound 5 precipitated immediately after the two reactants were combined and was isolated by filtration with a yield of 83% as a pale yellow solid that can be handled in the air. The 1H NMR spectrum of 5 differs clearly from the protonated ligand 2. In addition to two doublets for the protons of the methylene group linking the Cp ring and the phosphine site (2JHH = 18.4 Hz), there are four partially superimposed and slightly broadened resonances that are assigned to the four diastereotopic protons methylene groups in the “CpC-wings”. This speaks for a C1 symmetric structure on the NMR time scale and thus for a localization of the thalium(I) cation on one side of the Cp ring. The singlet at −15.56 ppm in the 31P NMR spectrum is only very weakly shifted towards the lower field compared to the signal of the metal-free ligand 2.
Crystallization from a saturated solution in the provided yellow, needle-like crystals with the monoclinic space group P21/n. Figure 3 shows the structure of 5 and summarizes relevant structural parameters.
There are numerous examples in the literature describing the coordination of thallium(I) to aromatic compounds [40,41,42,43,44,45,46], while only a few structurally characterized systems with thallium(I)–phosphorus coordination have been published. In 1974, Nakayama et al. published a gas-phase study on a series of thallium(I) complexes and stated that “No thallium(I) complexes with bismuth, arsenic nor phosphorus ligands has been isolated from a liquid phase so far.” [47]. This changed during the past two decades, in particular due to the use of weakly coordinating anions [48,49]. The fact that thallium(I) phosphine complexes are hardly accessible as long as there is an alternative for the metal cation is confirmed by the solid-state structure of compound 5. Compound 5 forms chains with thallium(I) centres bridging two (CpC)−1 moieties, as shown in Figure 3, and there is no interaction with the phosphine site. In addition to the coordination to the Cp site, thallium(I) also coordinates to two of the six-membered rings, the phenylene ring bound to the phosphorus atom and one of the phenylene rings of the CpC moiety. Since the covalency of the bonds between the thallium(I) and the π-donating moieties is low, the Tl-C bond lengths differ largely as documented in the caption of Figure 3 exemplarily for the Cp unit. This allows the adoption of the positioning of the cation to the steric requirements of the solid-state structure. According to the electrostatic interaction between the Cp site and the thallium(I) cation, the Cp-Tl distance (2.6131(4) Å) is much shorter than the Ar-Tl distances (3.3139(4) and 3.2030(4) Å).
In 2003, Salzer et al. developed a protocol for the preparation of chiral cyclopentadienide ruthenium(II) complexes with an ansa-bridging phosphine group and two labile bound acetonitrile ligands) [50,51]. They started from cyclopentadiene derivatives bearing a chiral phosphine substituent and [RuCl(μ2-Cl)(η33-C10H16)]2, used Li2CO3 as the base and KPF6 to provide a weakly coordinating counter anion. Due to our good experiences with TlCpC as a [CpC]−1 transfer reagent [19], we applied a combination of compound 5 and the ruthenium(II) precursor [(η6-C6H6)RuCl2]2 to generate an analogous CpC derivative (Scheme 5).
The reaction was carried out at room temperature in acetonitrile in the presence of a stoichiometric amount of KPF6. After work-up, compound 6 was isolated in 80% yield as an orange-colored solid, which was characterized by means of NMR and IR spectroscopy, mass spectrometry and elemental analysis. Tiny crystals of insufficient quality were obtained by slow diffusion of diethylether into a solution of compound 6 in acetonitrile. The presence of the two acetonitrile ligands was confirmed by a singlet resonance at 1.96 ppm in the 1H NMR spectrum and by an absorption band in the IR spectrum at 2038 cm−1. The resonances of the four diastereotopic protons of the two methylene units in the “wings” of the CpC ligand are identified by their 2JHH coupling constants of about 13 Hz. For the two diastereotopic protons of the third methylene unit, a 2JHH coupling constant of 18 Hz is observed. One of these two protons shows an additional coupling (4JPH = 6.8 Hz) to the phosphorus atom. The chemical shift of the 31P resonance of 51.7 ppm unambiguously proves the coordination of the phosphine site to the ruthenium(II) center. A peak at m/z = 810 in the ESI mass spectrum corresponds to the mass of the fragment [M-CH3CN]+ a second peak at m/z = 769 is assigned to the fragment [M-2 × CH3CN]+.
As mentioned above, the diphosphine functionalized CpC derivative 3 can only act as a typical diphosphine donor. However, the two phosphine sites are linked by a chain of overall seven carbon atoms allowing either a chelating coordination of the diphosphine to one metal site or a bridging coordination to two metal sites. The simplest way of dealing with these alternatives is by coordination of two equivalents of gold(I) to compound 3. The according gold(I) complex 7 was thus obtained in 83% yield in an analogous way as compound 4 using [(tht)AuCl] as the gold source (Scheme 6).
In contrast to the mono nuclear gold(I) complex 4, the dinuclear compound 7 is C2 symmetric, which largely simplifies the 1H NMR spectrum, leading to only four resonances for the diastereotopic protons of the two chemically inequivalent types of methylene groups with coupling constants of about 12.5 resp. 17.5 Hz. The 31P NMR resonance at 24.9 ppm is close to the value obtained for compound 4 (24.4 ppm).
By the slow diffusion of n-pentane into a saturated toluene solution of compound 7, single crystals suitable for an X-ray structure analysis could be obtained. Compound 7 crystallizes as colorless prisms of the triclinic space group P 1 ¯ . Figure 4 shows the structure of 7 and summarizes relevant structural parameters.
Typical for gold(I) complexes, the two gold sites are coordinated almost linearly. To minimize steric interactions, in particular of the diphenylphosphinyl units, the two P-Au-Cl units point into opposite directions, which excludes an intramolecular Au···Au interaction. The bond parameters are in the same range as measured for complex 4.
In 2006 Gelman et al. obtained a mono nuclear chelate complex as well as a dinuclear, chlorido-bridged palladium(II) complex by the treatment of the diphosphine ligand 1,8-bis(diisopropylphosphino)triptycen with different amounts of [PdCl2(CH3CN)2] [52,53]. In a series of analogue reactions, ligand 3 was reacted with different amounts of [PdCl2(CH3CN)2]. In contrast to the results of Gelman et al., only the dinuclear palladium complex 8 could be isolated, independent from the ligand-to-metal ratio and from solvent and temperature. By applying a slight excess of the metal source, the best yield was obtained (78%, Scheme 7).
1H and 13C NMR spectra of complex 8 clearly prove the presence of a C2 symmetric compound. There are, e.g., only two doublets of doublets for the diastereotopic protons of the overall four methylene groups. While the methylene groups bridging the phenylene rings in the “wings” of the ligand give rise to two doublets at 3.82 und 3.89 ppm with a typical 2JHH coupling constant of about 12.4 Hz, the two methylene units linking the phosphine sites to the CpC core are observed at 4.80 and at about 7.78 ppm. The value of the latter chemical shift is rather unexpected. It was identified by means of a H,H-COSY NMR experiment and is explained by the fact that of two of these methylene proton direct towards the central Pd2Cl4 unit. According to the C2 symmetry of compound 8, there is only one resonance in the 31P NMR spectrum, observed at 25.4 ppm, which is a typical value for this type of palladium(II) compounds [54].
By slow diffusion of n-pentane into a saturated toluene solution of complex 8, single crystals suitable for an X-ray structure analysis could be obtained. Compound 8 crystallizes as colorless prisms of the monoclinic space group P2/n with two crystallographically independent molecules in the unit cell. Figure 5 shows the structure of one of these units of compound 8 and summarizes relevant structural parameters.
As expected on the basis of the NMR spectra of compound 8, the phosphine donors coordinate to the central [(μ2-Cl)Pd(Cl)]2 unit in an anti-conformation that preserves the C2-symmetry of the CpC moiety. The two palladium(II) sites occupy a distorted square-planar geometry. In comparison to structurally similar [(L)Pd(μ2-Cl)(Cl)]2 (L = σ-donor ligand) compounds, the Pd···Pd distance (3.3098(6) Å) is by about 0.15–0.20 Å shorter, which we assign to the steric restrictions of the diphosphine ligand (see [55] and references cited therein). Due to the strongly σ-donating phosphine ligands the bond distances between the palladium(II) centers and the bridging chlorido ligands in trans-position to the phosphine donors are slightly longer than those in the trans-position to the apical chlorido ligands.
The fragment ClRh(CO) is well-known to prefer trans- over cis-coordination with chelating diphosphines as long as the linker between the phosphine sites allows this [56,57,58,59,60,61,62,63]. In ligand 3, the two phosphine sites are linked by a chain of overall seven carbon atoms. Therefore, the formation of a trans-coordinated rhodium(I) complex seemed likely, when ligand 3 was reacted first with half an equivalent of [(μ2-Cl)Rh(CO)2]2 at room temperature. However, spectroscopic characterization of the product unambiguously proved the generation of the dinuclear rhodium(I) complex 9 (Scheme 8). Reaction with one equivalent of the precursor [(μ2-Cl)Rh(CO)2]2 provided complex 9 in excellent yields.
The molecular structure of compound 9 resembles the structure of palladium(II) complex 8. In contrast, in the case of 9 the phosphine donors are coordinated in a syn- and not in an anti-arrangement to the metal sites, which breaks the C2-symmetry of the chelating ligand and leads to an overall C1-symmetric structure.
This decrease in symmetry results in the typical pattern of resonances for the methylene protons as described above for other complexes with C1-symmetry. As already observed for palladium complex 8, the 1H NMR resonances of two of the methylene protons are shifted largely towards lower field (7.00 ppm). The two phosphorus atoms are magnetically not equivalent and show couplings to the neighboring 103Rh centers (δ = 35.9 ppm, 1JRhP = 171.5 Hz and δ = 39.8 ppm, 1JRhP = 175.0 Hz), which are close to values measured for other binuclear rhodium(I) complexes of similar structure [52,64,65]. In the 13C NMR spectrum two doublets of doublets at δ = 183.2 (dd, 1JRhC = 78.0, 2JPC = 20.6 Hz) and δ = 182.3 (dd, 1JRhC = 78.8, 2JPC = 21.1 Hz) are assigned to the rhodium-bound carbonyl ligands. In the IR spectrum of compound 9, there are two carbonyl absorptions at 1991 and 1975 cm−1 [52,64,65].
By the slow diffusion of n-pentane into a saturated toluene solution of complex 9, single crystals suitable for an X-ray structure analysis could be obtained. Compound 9 crystallizes as yellow prisms of the triclinic space group P 1 ¯ . Figure 6 shows the structure of 9 and summarizes relevant structural parameters.
Both rhodium(I) sites in compound 9 occupy a distorted square-planar geometry, which is typical for d8-configured transition metal ions and has already been observed for the palladium(II) complex 8. In contrast to 8, however, the phosphines are coordinating the two rhodium(I) centers in a syn-configuration. This is in agreement with the NMR data of the complex 9 that correspond to a C1-symmetric structure. The Rh-Cl distances in trans-position to the π-accepting carbonyl ligands are slightly shorter than those in trans-position to the phosphine donors. The large interplanar angle of 132.28° between the chlorido-bridged rhodium(I) sites, the P-P distance of 6.378 Å, which is somewhat shorter with respect to the metal-free ligand and the relative long Rh-Rh distance of 3.28 Å confirm calculations carried out by López-Valbuena et al. on structural features of diphosphane ligands with large bite-angles [64].
As mentioned above, the fragment ClRh(CO) tends to form trans-coordinated mononuclear complexes with chelating diphosphines possessing long linker units. Therefore, ligand 3 was reacted with half an equivalent of the precursor [(μ2-Cl)Rh(CO)2]2. In 2006, Gelman et al. reported the preferential formation of a monometalic chelate complex using shorter reaction times [52,53]. In our case, however, the selective formation of the monorhodium(I) complex 10 was achieved by combining a prolonged reaction time and an elevated reaction temperature (Scheme 9).
The C2 symmetric ligand backbone in combination with a local CS symmetry of the coordination environment at the rhodium(I) site led to an overall C1 symmetry and thus, as expected, to eight doublets for the methylene protons in the 1H NMR spectrum of compound 10. Three of these resonances that are assigned according to their coupling constants to the methylene groups linking the CpC backbone with the phosphine sites are strongly shifted towards lower field. The X-ray structure (see below) of compound 10 shows that these three protons are oriented towards the central (trans-P)2Rh(CO)Cl unit. The 31P NMR spectrum of compound 10 reflects its C1 symmetry: there are two doublets of doublets with a chemical shift of 26.2 und 35.6 ppm, each showing a large 2JPP coupling constant of 340 Hz and a smaller 1JRhP coupling constant (123.2 resp. 127.5 Hz). The large 2JPP coupling constant is typical for the trans-orientation of the two phosphine donors [66]. Compared to the chlorido-bridged complex 9, the low field shift of the resonances of the two phosphorus atoms is not as pronounced. ESI mass spectrometry reveals a peak at m/z = 1073.39, which corresponds to the cationic fragment [M-Cl]+ (calcd.: m/z = 1073.25). In the IR spectrum of compound 10, there is an intense absorption ṽ(CO) at 1961 cm−1, which is typical for trans-coordinated, monomeric diphosphinerhodium carbonyl complexes [52,60,61,62,67,68].
By the slow diffusion of n-pentane into a saturated toluene solution of complex 10, single crystals suitable for an X-ray structure analysis could be obtained. Compound 10 crystallizes as yellow prisms of the triclinic space group P21/n; Figure 7 shows the structure of 10 and summarizes relevant structural parameters.
As reported for other mononuclear diphosphine carbonyl(chlorido)rhodium(I) complexes with long-chained diphosphine ligands, the two phosphine donors are oriented in trans-position [53,61,62,63]. The central rhodium(I) site is coordinated in a distorted square-planar geometry. Due to the steric influence of the CpC moiety, the P1-Rh1-P2 angle is significantly smaller than 180° (167.27(3)°) as is the Cl1Rh1-C70 angle (169.08(9)°).

4. Discussion

The novel chiral cyclopentadienyl-type ligand CpCH (1) is available following a high yield protocol of overall five steps starting from cheap dibenzosuberenone. We have shown in this manuscript, that CpCH (1) has turned out to be an ideal precursor for the synthesis of Cp-functionalized phosphine ligands. Both the monophosphine ligand 2 and the diphosphine ligand 3 are obtained by high yield nucleophilic substitution reactions with 2-[(chloromethyl)phenyl](diphenyl)phosphine using deprotonated compound 1. In particular, the simplicity of the synthesis of 3 providing a long-chained and chiral diphosphine is of interest, since syntheses of similar systems usually suffer from poor yields.
With compounds 2 and 3 overall six different types of coordination could be realized. The monophosphine 2 still has a cyclopentadiene backbone with one proton attached to the five-membered ring and is thus able to coordinate like a typical phosphine or a typical cyclopentadienide or by combining these two modes. Examples of all three modes of coordination could be obtained by selecting the appropriate (transition) metal precursor. In contrast to monophosphine 2, diphosphine 3 cannot be deprotonated anymore at the cyclopentadiene site. It therefore reacts like a typical diphosphine. Bridging coordination to two gold(I) sites, as well as the trans-coordination to one rhodium(I) in a square-planar environment, are expected, in terms of the coordinative behavior of these two metal sites. In the latter case, a ten-membered ring is formed. However, we also found two examples for a coordination to two d8 metal sites (palladium(II) and rhodium(I)) that are part of dinuclear chloride-bridged complexes. In these two cases, the ligand forms a twelve-membered ring including a M(μ2-Cl)2M moiety. We assign this rather outstanding flexibility in coordination on one side to the long chain (seven carbon atoms) that connects the two phosphine centers. On the other side, there are four aromatic and one quaternary sp3 hybridized carbon atoms in this chain, which reduce the number of possible conformations and thus stabilize the discussed modes of coordination.
In this manuscript, we have focused on aspects of the synthesis, spectroscopic data and structural elucidation of a series of complexes containing the new chiral phosphines 2 and 3. The evaluation of, for example, the catalytical properties of selected compounds from this series is presently under investigation. Here, we concentrate on transformations that are known to require diphosphine ligands having a large bite angle and on transformations catalyzed by the ansa-bridged ruthenium complex 6.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry5020062/s1. It contains figures of NMR and IR spectra of all compounds discussed in this manuscript.

Author Contributions

Investigation, F.N. and Y.S.; supervision and writing, W.R.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Carl Zeiss Stiftung who donated a grant to F.N.

Data Availability Statement

Data is contained within the article and supplementary materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Charrier, C.; Mathey, F. La diphenyl-cyclopentadienylmethyl-phosphine et ses complexesferroceniques et cymantreniques. Tetrahedron Lett. 1978, 19, 2407–2410. [Google Scholar] [CrossRef]
  2. Kettenbach, R.T.; Bonrath, W.; Butenschön, H. [ω-(Phosphanyl)alkyl]cyclopentadienyl Complexes. Chem. Ber. 1993, 126, 1657–1669. [Google Scholar] [CrossRef]
  3. Butenschön, H. Cyclopentadienylmetal Complexes Bearing Pendant Phosphorus, Arsenic, and Sulfur Ligands. Chem. Rev. 2000, 100, 1527–1564. [Google Scholar] [CrossRef] [PubMed]
  4. Bensley, D.M.; Mintz, E.A.; Sussangkarn, S.J. Synthesis of [C5(CH3)4H]CH2CH2CH2P(C6H5)2: A novel heterodifunctional ligand possessing both a tetramethylcyclopentadiene and a remote diphenylphosphine functionality. J. Org. Chem. 1988, 53, 4417–4419. [Google Scholar] [CrossRef]
  5. Bensley, D.M.; Mintz, E. 1,2,3,4,6-Pentamethylfulvene: A convenient precursor to substituted tetramethylcyclopentadienyl transition metal complexes. J. Organomet. Chem. 1988, 353, 93–102. [Google Scholar] [CrossRef]
  6. Hayashi, T.; Kumada, M. Asymmetric synthesis catalyzed by chiral ferrocenylphosphine-transition metal complexes. 2. Nickel- and palladium-catalyzed asymmetric Grignard cross-coupling. Acc. Chem. Res. 1982, 15, 395–401. [Google Scholar] [CrossRef]
  7. Fihri, A.; Meunier, P.; Hierso, J.-C. Performances of symmetrical achiral ferrocenylphosphine ligands in palladium-catalyzed cross-coupling reactions: A review of syntheses, catalytic applications and structural properties. Coord. Chem. Rev. 2007, 251, 2017–2055. [Google Scholar] [CrossRef]
  8. Kataoka, N.; Shelby, Q.; Stambuli, J.P.; Hartwig, J.F. Air Stable, Sterically Hindered Ferrocenyl Dialkylphosphines for Palladium-Catalyzed C−C, C−N, and C−O Bond-Forming Cross-Couplings. J. Org. Chem. 2002, 67, 5553–5566. [Google Scholar] [CrossRef]
  9. Ito, Y.; Sawamura, M.; Hayashi, T. Catalytic asymmetric aldol reaction: Reaction of aldehydes with isocyanoacetate catalyzed by a chiral ferrocenylphosphine-gold(I) complex. J. Am. Chem. Soc. 1986, 108, 6405–6406. [Google Scholar] [CrossRef]
  10. Hayashi, T.; Mise, T.; Fukushima, M.; Kagotani, M.; Nagashima, N.; Hamada, Y.; Matsumoto, A.; Kawakami, S.; Konishi, M. Asymmetric Synthesis Catalyzed by Chiral Ferrocenylphosphine–Transition Metal Complexes. I. Preparation of Chiral Ferrocenylphosphines. Bull. Chem. Soc. Jpn. 1980, 53, 1138–1151. [Google Scholar] [CrossRef]
  11. Hayashi, T.; Yamamoto, A.; Ito, Y.; Nishioka, E.; Miura, H.; Yanagi, K. Asymmetric Synthesis Catalyzed by Chiral Ferrocenylphosphine-Transition-Metal Complexes. 8. Palladium-Catalyzed Asymmetric Allylic Amination. J. Am. Chem. Soc. 1989, 111, 6301–6311. [Google Scholar] [CrossRef]
  12. Hayashi, T.; Hayashizaki, K.; Kiyoi, T.; Ito, Y. Asymmetric synthesis catalyzed by chiral ferrocenylphosphine-transition-metal complexes. 6. Practical asymmetric synthesis of 1,1′-binaphthyls via asymmetric cross-coupling with a chiral [(alkoxyalkyl)ferrocenyl]monophosphine/nickel catalyst. J. Am. Chem. Soc. 1988, 110, 8153–8156. [Google Scholar] [CrossRef]
  13. Chen, Y.; Yi, X.; Cheng, Y.; Huang, A.; Yang, Z.; Zhao, X.; Ling, F.; Zhong, W. Rh-Catalyzed Highly Enantioselective Hydrogenation of Functionalized Olefins with Chiral Ferrocenylphosphine-Spiro Phosphonamidite Ligands. J. Org. Chem. 2022, 87, 7864–7874. [Google Scholar] [CrossRef] [PubMed]
  14. Barth, E.L.; Davis, R.M.; Mohadjer Beromi, M.; Walden, A.G.; Balcells, D.; Brudvig, G.W.; Dardir, A.H.; Hazari, N.; Lant, H.M.C.; Mercado, B.Q.; et al. Bis(dialkylphosphino)ferrocene-Ligated Nickel(II) Precatalysts for Suzuki–Miyaura Reactions of Aryl Carbonates. Organometallics 2019, 38, 3377–3387. [Google Scholar] [CrossRef]
  15. Škoch, K.; Schulz, J.; Císařová, I.; Štěpnička, P. Pd(II) Complexes with Chelating Phosphinoferrocene Diaminocarbene Ligands: Synthesis, Characterization, and Catalytic Use in Pd-Catalyzed Borylation of Aryl Bromides. Organometallics 2019, 38, 3060–3073. [Google Scholar] [CrossRef]
  16. Škoch, K.; Císařová, I.; Štěpnička, P. Synthesis of a Polar Phosphinoferrocene Amidosulfonate Ligand and Its Application in Pd-Catalyzed Cross-Coupling Reactions of Aromatic Boronic Acids and Acyl Chlorides in an Aqueous Medium. Organometallics 2016, 35, 3378–3387. [Google Scholar] [CrossRef]
  17. Chung, J.-Y.; Schulz, C.; Bauer, H.; Sun, Y.; Sitzmann, H.; Auerbach, H.; Pierik, A.J.; Schünemann, V.; Neuba, A.; Thiel, W.R. Cyclopentadienide Ligand CpC– Possessing Intrinsic Helical Chirality and Its Ferrocene Analogues. Organometallics 2015, 34, 5374–5382. [Google Scholar] [CrossRef]
  18. Chung, J.-Y.; Sun, Y.; Thiel, W.R. Titanium(IV) complexes bearing the (CpC) ligand. J. Organometal. Chem. 2017, 829, 31–36. [Google Scholar] [CrossRef]
  19. Nährig, F.; Gemmecker, G.; Chung, J.-Y.; Hütchen, P.; Lauk, S.; Klein, M.; Sun, Y.; Niedner-Schatteburg, G.; Sitzmann, H.; Thiel, W.R. Complexes of Platinum Group Elements Containing the Intrinsically Chiral Cyclopentadienide Ligand (CpC)−1. Organometallics 2020, 39, 1934–1944. [Google Scholar] [CrossRef]
  20. Nährig, F.; Nunheim, N.; Salih, K.S.M.; Chung, J.-Y.; Gond, D.; Sun, Y.; Becker, S.; Thiel, W.R. A Novel Cyclopentadienone and its Ruthenium and Iron Tricarbonyl Complexes. Eur. J. Inorg. Chem. 2021, 4832–4841. [Google Scholar] [CrossRef]
  21. Armarego, W.L.F.; Chai, C.L.L. Purification of Laboratory Chemicals, 6th ed.; Butterworth-Heinemann: Oxford, UK, 2009. [Google Scholar]
  22. Herrmann, J.; Pregosin, P.S.; Salzmann, R.; Albinati, A. Palladium π-Allyl Chemistry of New P,S Bidentate Ligands. Selective but Variable Dynamics in the Isomerization of the η3-C3H5 and η3-PhCHCHCHPh π-Allyl Ligands. Organometallics 1995, 14, 3311–3318. [Google Scholar] [CrossRef]
  23. Rauchfuss, T.B.; Patino, F.T.; Roundhill, D.M. Platinum Metal Complexes of Amine- and Ether-Substituted Phosphines. Inorg. Chem. 2002, 14, 652–656. [Google Scholar] [CrossRef]
  24. Altomare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A.; Burla, M.C.; Polidori, G.; Camalli, M. SIR92—A program for automatic solution of crystal structures by direct methods. J. Appl. Cryst. 1994, 27, 435. [Google Scholar] [CrossRef]
  25. Sheldrick, G.M. A short history of SHELX. Acta Cryst. 2008, A64, 112–122. [Google Scholar] [CrossRef]
  26. Rigaku Oxford Diffraction; Version 1.171.38.46; CrysAlisPro: Seattle, WA, USA, 2015.
  27. Li, J.; Yin, J.; Wang, G.-X.; Yin, Z.-B.; Zhang, W.-X.; Xi, Z. Synthesis and reactivity of asymmetric Cr(i) dinitrogen complexes supported by cyclopentadienyl–phosphine ligands. Chem. Commun. 2019, 55, 9641–9644. [Google Scholar] [CrossRef] [PubMed]
  28. Ireland, T.; Tappe, K.; Grossheimann, G.; Knochel, P. Synthesis of a New Class of Chiral 1,5-Diphosphanylferrocene Ligands and Their Use in Enantioselective Hydrogenation. Chem. Eur. J. 2002, 8, 843–852. [Google Scholar] [CrossRef]
  29. Trampert, J.; Sun, Y.; Thiel, W.R. The reactivity of [{2-(diphenylphosphino)phenyl}methyl]-3-imidazol-2-ylidenes towards group VIII element precursors. J. Organometal. Chem. 2020, 915, 121222. [Google Scholar] [CrossRef]
  30. Keough, P.T.; Grayson, M. Phosphonioethylation. Michael Addition to Vinylphosphonium Salts. J. Org. Chem. 1964, 29, 631–635. [Google Scholar] [CrossRef]
  31. Matsusaka, Y.; Shitaya, S.; Nomura, K.; Inagaki, A. Synthesis of Mono-, Di-, and Trinuclear Rhodium Diphosphine Complexes Containing Light-Harvesting Fluorene Backbones. Inorg. Chem. 2017, 56, 1027–1030. [Google Scholar] [CrossRef]
  32. Xie, L.; Zhao, S.; Zhang, M.; Wu, C.; Li, T.; Gao, M. Compound of Fluorene Class Containing Dual Hetero Atoms as well as Its Synthetic Method and Application. CN Patent 1480453A, 10 March 2004. [Google Scholar]
  33. Liu, H.; Xie, L.; Jing, J.M. Solid Catalyst Composition for Olefinic Polymerization and Catalyst Thereof. CN Patent 1508159A, 30 June 2004. [Google Scholar]
  34. Eger, T.R.; Munstein, I.; Steiner, A.; Sun, Y.; Niedner-Schatteburg, G.; Thiel, W.R. New cationic organometallic phosphane ligands and their coordination to gold(I). J. Organomet. Chem. 2016, 810, 51–56. [Google Scholar] [CrossRef]
  35. Batchelor, L.K.; Păunescu, E.; Soudani, M.; Scopelliti, R.; Dyson, P.J. Influence of the Linker Length on the Cytotoxicity of Homobinuclear Ruthenium(II) and Gold(I) Complexes. Inorg. Chem. 2017, 56, 9617–9633. [Google Scholar] [CrossRef] [PubMed]
  36. Borissova, A.O.; Korlyukov, A.A.; Antipin, M.Y.; Lyssenko, K.A. Estimation of Dissociation Energy in Donor−Acceptor Complex AuCl·PPh3 via Topological Analysis of the Experimental Electron Density Distribution Function. J. Phys. Chem. A 2008, 112, 11519–11522. [Google Scholar] [CrossRef] [PubMed]
  37. Zi, W.; Dean Toste, F. Recent advances in enantioselective gold catalysis. Chem. Soc. Rev. 2016, 45, 4567–4589. [Google Scholar] [CrossRef]
  38. Johnson, B.F.G. The chemistry of gold. Gold Bull. 1971, 4, 9–11. [Google Scholar] [CrossRef]
  39. Scherbaum, F.; Grohmann, A.; Huber, B.; Krüger, C.; Schmidbaur, H. Use of the CH Acidity of 2,4,4-Trimethyl-4,5-dihydrooxazole to Synthesize Triauriomethanes and Novel Gold Clusters. Angew. Chem. Int. Ed. Engl. 1988, 27, 1544–1546. [Google Scholar] [CrossRef]
  40. Janiak, C. (Organo)thallium (I) and (II) chemistry: Syntheses, structures, properties and applications of subvalent thallium complexes with alkyl, cyclopentadienyl, arene or hydrotris(pyrazolyl)borate ligands. Coord. Chem. Rev. 1997, 163, 107–216. [Google Scholar] [CrossRef]
  41. Schumann, H.; Janiak, C.; Khani, H. Cyclopentadienylthallium(I) compounds with bulky cyclopentadienyl ligands. J. Organomet. Chem. 1987, 330, 347–355. [Google Scholar] [CrossRef]
  42. Schumann, H.; Janiak, C.; Khan, M.A.; Zuckerman, J.J. Eine zweite ungewöhnliche Kristallmodifikation von pentabezylcyclopentadienylthallium(I), (PhCH2)5C5Tl. J. Organomet. Chem. 1988, 354, 7–13. [Google Scholar] [CrossRef]
  43. Frasson, E.; Menegus, F.; Panattoni, C. Chain Structure of the Cyclopentadienily of Monovalent Indium and Thallium. Nature 1963, 199, 1087–1089. [Google Scholar] [CrossRef]
  44. Werner, H.; Otto, H.; Kraus, H.J. Die Kristallstruktur von TlC5Me5. J. Organomet. Chem. 1986, 315, C57–C60. [Google Scholar] [CrossRef]
  45. Schmidbaur, H.; Bublak, W.; Riede, J.; Müller, G. [{1,3,5-(CH3)3H3C6}6TI4]·[GaBr4]4—Synthese und Struktur eines gemischten Mono- und Bis(aren)thallium-Komplexes. Angew. Chem. 1985, 97, 402–403. [Google Scholar] [CrossRef]
  46. Schmidbaur, H. Arenkomplexe von einwertigem Gallium, Indium und Thallium. Angew. Chem. 1985, 97, 893–904. [Google Scholar] [CrossRef]
  47. Nakayama, H.; Nishijima, C.; Tachiyashiki, S. Ion-Molecule Reactions Betwee Thallium(I) and Various Ligands: Formation of 1:1 Complexes in Gas Phase. Chem. Lett. 1974, 3, 733–736. [Google Scholar] [CrossRef]
  48. Betley, T.A.; Peters, J.C. The Strong-Field Tripodal Phosphine Donor, [PhB(CH2PiPr2)3], Provides Access to Electronically and Coordinatively Unsaturated Transition Metal Complexes. Inorg. Chem. 2003, 42, 5074–5084. [Google Scholar] [CrossRef]
  49. Szlosek, R.; Ackermann, M.T.; Marquardt, C.; Seidl, M.; Timoshkin, A.Y.; Scheer, M. Coordination of Pnictogenylboranes Towards Tl(I) Salts and a Tl- Mediated P-P Coupling. Chem. Eur. J. 2023, 29, e202202911. [Google Scholar] [CrossRef]
  50. Doppiu, A.; Englert, U.; Salzer, A. Cationic half-sandwich ruthenium(II) complexes with cyclopentadienyl–phosphine ligands. Inorg. Chim. Acta 2003, 350, 435–441. [Google Scholar] [CrossRef]
  51. Doppiu, A.; Salzer, A. A New Route to Cationic Half-Sandwich Ruthenium(II) Complexes with Chiral Cyclopentadienylphos phane Ligands. Eur. J. Inorg. Chem. 2004, 2244–2252. [Google Scholar] [CrossRef]
  52. Azerraf, C.; Cohen, S.; Gelman, D. Roof-Shaped Halide-Bridged Bimetallic Complexes via Ring Expansion Reaction. Inorg. Chem. 2006, 45, 7010–7017. [Google Scholar] [CrossRef] [PubMed]
  53. Azerraf, C.; Grossman, O.; Gelman, D. Rigid trans-spanning triptycene-based ligands: How flexible they can be? J. Organomet. Chem. 2007, 692, 761–767. [Google Scholar] [CrossRef]
  54. Noskowska, M.; Śliwińska, E.; Duczmal, W. Simple fast preparation of neutral palladium (II) complexes with SnCl3 and Cl ligands. Trans. Met. Chem. 2003, 28, 756–759. [Google Scholar] [CrossRef]
  55. Aullón, G.; Ujaque, G.; Lledós, A.; Alvarez, S.; Alemany, P. To Bend or Not To Bend:  Dilemma of the Edge-Sharing Binuclear Square Planar Complexes of d8 Transition Metal Ions. Inorg. Chem. 1998, 37, 804–813. [Google Scholar] [CrossRef]
  56. Bunten, K.A.; Farrar, D.H.; Poë, A.J.; Lough, A. Stoichiometric and Catalytic Oxidation of BINAP by Dioxygen in a Rhodium(I) Complex. Organometallics 2002, 21, 3344–3350. [Google Scholar] [CrossRef]
  57. Mondal, J.U.; Young, K.G.; Blake, D.M. Enthalpy changes in oxidative addition reactions of iodine with monomeric and dimeric rhodium(I) complexes. J. Organomet. Chem. 1982, 240, 447–451. [Google Scholar] [CrossRef]
  58. Dyer, G.; Wharf, R.M.; Hill, W.E. 31P NMR studies of cis-[RhCl(CO)(bis-phosphine)] complexes. Inorg. Chim. Acta 1987, 133, 137–140. [Google Scholar] [CrossRef]
  59. Mann, B.E.; Masters, C.; Shaw, B.L. Nuclear magnetic resonance studies on metal complexes. Part VII. The 31P n.m.r. spectra of some complexes of the type trans-RhCl-(CO)L2 [L = tertiary phosphine or P(OMe3)]. J. Chem. Soc. A 1971, 1104–1106. [Google Scholar] [CrossRef]
  60. Vallarino, L. Carbonyl complexes of rhodium. Part I. Complexes with triarylphosphines, triarylarsines, and triarylstibines. J. Chem. Soc. 1957, 2287–2292. [Google Scholar] [CrossRef]
  61. Marty, W.; Kapoor, P.N.; Bürgi, H.-B.; Fischer, E. Complexes of 3,3′-Oxybis[(diphenylphosphino)methylbenzene] with Ni(II), Pd(II), Pt(II), Rh(I), and Ag(I). How Important is Backbone Rigidity in the Formation of trans-Spanning Bidenatate Chelates? Helv. Chim. Acta 1987, 70, 158–170. [Google Scholar] [CrossRef]
  62. Eberhard, M.R.; Heslop, K.M.; Orpen, A.G.; Pringle, P.G. Nine-Membered Trans Square-Planar Chelates Formed by a bisbi Analogue. Organometallics 2005, 24, 335–337. [Google Scholar] [CrossRef]
  63. Reed, F.J.S.; Venanzi, L.M. Transition metal complexes with bidentate ligands spanning trans-positions. IV. Preparation and properties of some rhodium and iridium complexes of 2,11-bis(diphenylphosphinomethyl)benzo[c]phenanthrene. Helv. Chim. Acta 1977, 60, 2804–2814. [Google Scholar] [CrossRef]
  64. López-Valbuena, J.M.; Escudero-Adan, E.C.; Benet-Buchholz, J.; Freixa, Z.; van Leeuwen, P.W.N.M. An approach to bimetallic catalysts by ligand design. Dalton Trans. 2010, 39, 8560–8574. [Google Scholar] [CrossRef]
  65. Hierso, J.-C.; Lacassin, F.; Broussier, R.; Amardeil, R.; Meunier, P. Synthesis and characterisation of a new class of phosphine-phosphonite ferrocenediyl dinuclear rhodium complexes. J. Organomet. Chem. 2004, 689, 766–769. [Google Scholar] [CrossRef]
  66. Meeuwissen, J.; Sandee, A.J.; de Bruin, B.; Siegler, M.A.; Spek, A.L.; Reek, J.N.H. Phosphinoureas: Cooperative Ligands in Rhodium-Catalyzed Hydroformylation? On the Possibility of a Ligand-Assisted Reductive Elimination of the Aldehyde. Organometallics 2010, 29, 2413–2421. [Google Scholar] [CrossRef]
  67. Sanger, A.R.J. Reactions of di-µ-chloro-bis[cyclo-octa-1,5-dienerhodium(I)] with carbon mono-oxide and mono-, di-, or tri-tertiary phosphines, or 1,2-bis(diphenylarsino)ethane. Chem. Soc. Dalton Trans. 1977, 120–129. [Google Scholar] [CrossRef]
  68. Thurner, C.L.; Barz, M.; Spiegler, M.; Thiel, W.R. Ligands with cycloalkane backbones II. Chelate ligands from 2-(diphenylphosphinyl)cyclohexanol: Syntheses and transition metal complexes. J. Organomet. Chem. 1997, 541, 39–49. [Google Scholar] [CrossRef]
Scheme 1. Numbering for the assignment of NMR resonances (see below) of the CpC complexes.
Scheme 1. Numbering for the assignment of NMR resonances (see below) of the CpC complexes.
Chemistry 05 00062 sch001
Scheme 2. Synthesis of ligands 2 and 3.
Scheme 2. Synthesis of ligands 2 and 3.
Chemistry 05 00062 sch002
Figure 1. Molecular structure of compound 3 in the solid state. The two co-crystallizing molecules of toluene and all hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: C1-C2 1.525(2), C1-C5 1.532(2), C2-C3 1.358(2), C3-C4 1.476(2), C4-C5 1.359(3), P1···C1···P2 99.25(3), C3-C2-C6-C7 40.5(3), C2-C3-C10-C9 -48.2(3), C4-C5-C23-C22 36.8(3), C5-C4-C19-C20 -45.7(3).
Figure 1. Molecular structure of compound 3 in the solid state. The two co-crystallizing molecules of toluene and all hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: C1-C2 1.525(2), C1-C5 1.532(2), C2-C3 1.358(2), C3-C4 1.476(2), C4-C5 1.359(3), P1···C1···P2 99.25(3), C3-C2-C6-C7 40.5(3), C2-C3-C10-C9 -48.2(3), C4-C5-C23-C22 36.8(3), C5-C4-C19-C20 -45.7(3).
Chemistry 05 00062 g001
Scheme 3. Synthesis of the mononuclear gold(I) complex 4.
Scheme 3. Synthesis of the mononuclear gold(I) complex 4.
Chemistry 05 00062 sch003
Figure 2. Molecular structure of compound 4 in the solid state. The co-crystallizing molecule of toluene and all hydrogen atoms except the hydrogen atom bound to C1 are omitted for clarity. Ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Au1-P1 2.2427(10), Au1-Cl1 2.2865(10), C1-C2 1.516(5), C1-C5 1.515(5), C2-C3 1.356(6), C3-C4 1.486(5), C4-C5 1.369(6), P1-Au1-Cl1 179.42(4), C3-C2-C6-C7 42.61(7), C4-C5-C23-C22 36.71(7), C2-C3-C10-C9 −45.98(7), C5-C4-C19-C20 −42.35(7).
Figure 2. Molecular structure of compound 4 in the solid state. The co-crystallizing molecule of toluene and all hydrogen atoms except the hydrogen atom bound to C1 are omitted for clarity. Ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Au1-P1 2.2427(10), Au1-Cl1 2.2865(10), C1-C2 1.516(5), C1-C5 1.515(5), C2-C3 1.356(6), C3-C4 1.486(5), C4-C5 1.369(6), P1-Au1-Cl1 179.42(4), C3-C2-C6-C7 42.61(7), C4-C5-C23-C22 36.71(7), C2-C3-C10-C9 −45.98(7), C5-C4-C19-C20 −42.35(7).
Chemistry 05 00062 g002
Scheme 4. Synthesis of the CpC thallium(I) complex 5.
Scheme 4. Synthesis of the CpC thallium(I) complex 5.
Chemistry 05 00062 sch004
Figure 3. Molecular structure of compound 5 in the solid state. All hydrogen atoms are omitted for clarity, ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: C1-Tl1 2.8405(18), C2-Tl1 2.8137(18), C3-Tl1 2.8799(17), C4-Tl1 2.9341 (17), C5-Tl1 2.9342 (17), Cp-Tl1 2.6131(3), Ar1-Tl2 3.3139(4), Ar2-Tl2 3.2030(4), C1-C2 1.419(3), C1-C5 1.418(3), C2-C3 1.433(3), C3-C4 1.428(3), C4-C5 1.431(3), Cp-Tl1-Ar1 124.770(9), Cp-Tl1-Ar2 132.793(11), Ar1-Tl-Ar2 95.279(10), C3-C2-C6-C7 41.8(3), C4-C5-C23-C22 44.0(3), C2-C3-C10-C9 −37.3(2), C5-C4-C19-C20 −42.8(3); Cp denotes the centroid of the five-membered cyclopentadienyl ring. Ar1 and Ar2 denote the centroids of the two six-membered rings, which undergo coordination to the thallium(I) cation.
Figure 3. Molecular structure of compound 5 in the solid state. All hydrogen atoms are omitted for clarity, ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: C1-Tl1 2.8405(18), C2-Tl1 2.8137(18), C3-Tl1 2.8799(17), C4-Tl1 2.9341 (17), C5-Tl1 2.9342 (17), Cp-Tl1 2.6131(3), Ar1-Tl2 3.3139(4), Ar2-Tl2 3.2030(4), C1-C2 1.419(3), C1-C5 1.418(3), C2-C3 1.433(3), C3-C4 1.428(3), C4-C5 1.431(3), Cp-Tl1-Ar1 124.770(9), Cp-Tl1-Ar2 132.793(11), Ar1-Tl-Ar2 95.279(10), C3-C2-C6-C7 41.8(3), C4-C5-C23-C22 44.0(3), C2-C3-C10-C9 −37.3(2), C5-C4-C19-C20 −42.8(3); Cp denotes the centroid of the five-membered cyclopentadienyl ring. Ar1 and Ar2 denote the centroids of the two six-membered rings, which undergo coordination to the thallium(I) cation.
Chemistry 05 00062 g003
Scheme 5. Synthesis of the CpC ruthenium(II) complex 6.
Scheme 5. Synthesis of the CpC ruthenium(II) complex 6.
Chemistry 05 00062 sch005
Scheme 6. Synthesis of the dinuclear gold(I) complex 7.
Scheme 6. Synthesis of the dinuclear gold(I) complex 7.
Chemistry 05 00062 sch006
Figure 4. Molecular structure of compound 7 in the solid state. All hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Au1-Cl1 2.2878(13), Au2-Cl2 2.2807(13), Au1-P1 2.2352(11), Au2-P2 2.2295(13), C1-C2 1.532(6), C2-C3 1.358(6), C3-C4 1.467(7), C4-C5 1.367(6), C1-C5 1.530(6), P1-Au1-Cl1 171.50(4), P2-Au2-Cl2 173.53(5), C3-C2-C6-C7 33.7(7), C4-C5-C23-C22 −34.8(6), C2-C3-C10-C9 49.9(10), C5-C4-C19-C20 47.5(6).
Figure 4. Molecular structure of compound 7 in the solid state. All hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Au1-Cl1 2.2878(13), Au2-Cl2 2.2807(13), Au1-P1 2.2352(11), Au2-P2 2.2295(13), C1-C2 1.532(6), C2-C3 1.358(6), C3-C4 1.467(7), C4-C5 1.367(6), C1-C5 1.530(6), P1-Au1-Cl1 171.50(4), P2-Au2-Cl2 173.53(5), C3-C2-C6-C7 33.7(7), C4-C5-C23-C22 −34.8(6), C2-C3-C10-C9 49.9(10), C5-C4-C19-C20 47.5(6).
Chemistry 05 00062 g004
Scheme 7. Synthesis of the dinuclear, chlorido-bridged palladium(II) complex 8.
Scheme 7. Synthesis of the dinuclear, chlorido-bridged palladium(II) complex 8.
Chemistry 05 00062 sch007
Figure 5. Molecular structure of compound 8 in the solid state. All hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Pd···Pd 3.3098(6), Pd1-Cl1 2.2880(16), Pd1-Cl2 2.3144(16), Pd1-P1 2.2499(14), C1-C2 1.533(6), C2-C3 1.360(7), C3-C4 1.475(7), Cl1-Pd1-Cl2 173.99(6), Cl1-Pd1-P1 89.96(5), Cl1-Pd1-Cl2_a 90.27(5), Cl2-Pd1-P1 95.71(5), Cl2-Pd1-Cl2_a 83.83(6), Cl2_a-Pd1-P1 171.23(5), C3-C2-C8-C9 35.5(5), C2-C3-C4-C13 47.1(6).
Figure 5. Molecular structure of compound 8 in the solid state. All hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Pd···Pd 3.3098(6), Pd1-Cl1 2.2880(16), Pd1-Cl2 2.3144(16), Pd1-P1 2.2499(14), C1-C2 1.533(6), C2-C3 1.360(7), C3-C4 1.475(7), Cl1-Pd1-Cl2 173.99(6), Cl1-Pd1-P1 89.96(5), Cl1-Pd1-Cl2_a 90.27(5), Cl2-Pd1-P1 95.71(5), Cl2-Pd1-Cl2_a 83.83(6), Cl2_a-Pd1-P1 171.23(5), C3-C2-C8-C9 35.5(5), C2-C3-C4-C13 47.1(6).
Chemistry 05 00062 g005
Scheme 8. Synthesis of the dinuclear, chlorido-bridged rhodium(I) carbonyl complex 9.
Scheme 8. Synthesis of the dinuclear, chlorido-bridged rhodium(I) carbonyl complex 9.
Chemistry 05 00062 sch008
Figure 6. Molecular structure of compound 9 in the solid state. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. All hydrogen atoms are omitted for clarity. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Rh···Rh 3.2797(5), Rh1-Cl1 2.3926(7), Rh1-Cl2 2.4127(7), Rh1-P1 2.2486(7), Rh1-C70 1.794(3), Rh2-Cl1 2.3880(7), Rh2-Cl2 2.4053(7), Rh2-P2 2.2431(7), Rh2-C71 1.805(3), Cl1-Rh1-Cl2 83.42(2), Cl1-Rh1-P1 96.09(2), Cl1-Rh1-C70 172.39(10), Cl2-Rh1-P1 176.67(3), Cl2-Rh1-C70 91.26(9), P1-Rh1-C70 89.53(9), Cl1-Rh2-Cl2 83.68(2), Cl1-Rh2-P2 97.62(3), Cl1-Rh2-C71 173.89(9), Cl2-Rh2-P2 177.68(3), Cl2-Rh2-C71 91.67(9), P2-Rh2-C71 86.90(9), Rh1-Cl1-Rh2 86.63(2), Rh1-Cl2-Rh2 85.80(2), C3-C2-C6-C7 38.2(3), C2-C3-C10-C9 46.6(4), C5-C4-C19-C20 45.7(4), C4-C5-C23-C22-37.7(4).
Figure 6. Molecular structure of compound 9 in the solid state. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. All hydrogen atoms are omitted for clarity. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Rh···Rh 3.2797(5), Rh1-Cl1 2.3926(7), Rh1-Cl2 2.4127(7), Rh1-P1 2.2486(7), Rh1-C70 1.794(3), Rh2-Cl1 2.3880(7), Rh2-Cl2 2.4053(7), Rh2-P2 2.2431(7), Rh2-C71 1.805(3), Cl1-Rh1-Cl2 83.42(2), Cl1-Rh1-P1 96.09(2), Cl1-Rh1-C70 172.39(10), Cl2-Rh1-P1 176.67(3), Cl2-Rh1-C70 91.26(9), P1-Rh1-C70 89.53(9), Cl1-Rh2-Cl2 83.68(2), Cl1-Rh2-P2 97.62(3), Cl1-Rh2-C71 173.89(9), Cl2-Rh2-P2 177.68(3), Cl2-Rh2-C71 91.67(9), P2-Rh2-C71 86.90(9), Rh1-Cl1-Rh2 86.63(2), Rh1-Cl2-Rh2 85.80(2), C3-C2-C6-C7 38.2(3), C2-C3-C10-C9 46.6(4), C5-C4-C19-C20 45.7(4), C4-C5-C23-C22-37.7(4).
Chemistry 05 00062 g006
Scheme 9. Synthesis of the mononuclear rhodium(I) carbonyl complex 10.
Scheme 9. Synthesis of the mononuclear rhodium(I) carbonyl complex 10.
Chemistry 05 00062 sch009
Figure 7. Molecular structure of compound 10 in the solid state. All hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. All hydrogen atoms are omitted for clarity. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Rh1-Cl1 2.3953(7), Rh1-P1 2.3081(7), Rh1-P2 2.3356(7), Rh1-C70 1.813(3), Cl1-Rh1-P1 87.04(3), Cl1-Rh1-P2 87.15(2), Cl1-Rh1-C70 169.08(9), P1-Rh1-P2 167.27(3), P1-Rh1-C70 90.38(8), P2-Rh1-C70 97.37(8), C3-C2-C6-C7 35.9(4), C2-C3-C10-C9 47.3(3), C5-C4-C19-C20 45.3(3), C4-C5-C23-C22 39.5(3).
Figure 7. Molecular structure of compound 10 in the solid state. All hydrogen atoms are omitted for clarity. The central CpC system is represented in stick-style for clarity, ellipsoids are drawn at the 50% level. All hydrogen atoms are omitted for clarity. Characteristic bond lengths [Å], angles [°] and dihedral angles [°]: Rh1-Cl1 2.3953(7), Rh1-P1 2.3081(7), Rh1-P2 2.3356(7), Rh1-C70 1.813(3), Cl1-Rh1-P1 87.04(3), Cl1-Rh1-P2 87.15(2), Cl1-Rh1-C70 169.08(9), P1-Rh1-P2 167.27(3), P1-Rh1-C70 90.38(8), P2-Rh1-C70 97.37(8), C3-C2-C6-C7 35.9(4), C2-C3-C10-C9 47.3(3), C5-C4-C19-C20 45.3(3), C4-C5-C23-C22 39.5(3).
Chemistry 05 00062 g007
Table 1. Crystallographic data, data collection and refinement.
Table 1. Crystallographic data, data collection and refinement.
345
empirical formulaC83H68PC57H45AuClPC50H36PTl
formula weight1127.31993.31872.13
crystal size [mm]0.408 × 0.307 × 0.2390.244 × 0.106 × 0.0690.313 × 0.093 × 0.063
T [K]293(2)150(2)150(2)
λ [Å]1.541841.541841.54184
crystal systemtriclinictriclinicmonoclinic
space group P 1 ¯ P 1 ¯ P21/n
a [Å]14.1471(5)10.3433(3)18.7154(2)
b [Å]14.5119(4)15.2162(6)12.2606(1)
c [Å]17.7514(6)15.5558(7)20.4489(2)
α [°]107.920(3)106.743(4)90
β [°]109.661(3)107.515(3)97.058(1)
γ [°]93.028(3)96.040(3)90
V [Å3]3215.3(2)2185.84(16)4656.69(8)
Z224
ρ calcd. [g·cm−3]1.1641.5091.244
μ [mm−1]0.9507.5197.198
Θ-range [°]3.251–62.7363.608–62.7293.417–62.775
refl. coll.24,39115,04835,775
indep. refl.10,246
[Rint = 0.0229]
6947
[Rint = 0.0481]
7447
[Rint = 0.0257]
data/restr./param.10246/0/7686947/0/5427447/0/469
final R indices
[I > 2σ(I)] a
0.0522, 0.14550.0367, 0.09460.0171, 0.0402
R indices (all data)0.0573, 0.15240.0381, 0.09610.0179, 0.0407
GooF b1.0361.0821.029
Δρmax/min (e·Å−3)1.687/−0.2831.444/−1.8080.391/−0.266
a R1 = Σ||Fo| − |Fc||/Σ|Fo|, ωR2 = [Σ(Fo2Fc2)2/ΣωFo2]1/2. b GooF = [Σω(Fo2Fc2)2/(n − p)]1/2.
Table 2. Crystallographic data, data collection and refinement.
Table 2. Crystallographic data, data collection and refinement.
78910
empirical formulaC69H52Au2Cl2P2C69H52Cl4P2Pd2C71H52Cl2O2P2Rh2C70H52ClOP2Rh
formula weight1407.881297.641275.781109.41
crystal size [mm]0.216 × 0.095 × 0.0790.199 × 0.101 × 0.0440.272 × 0.151 × 0.1370.233 × 0.121 × 0.059
T [K]150(2)150(2)150(2)150(2)
λ [Å]1.541841.541841.541841.54184
crystal systemtriclinicmonoclinictriclinicmonoclinic
space group P 1 ¯ P2/n P 1 ¯ P21/n
a [Å]13.1508(3)21.0446(7)13.2483(3)13.3016(3)
b [Å]13.5694(4)13.2823(4)14.1374(3)21.0013(4)
c [Å]22.2637(5)23.4004(9)20.4213(4)20.0864(4)
α [°]76.934(2)9089.220(2)90
β [°]87.525(2)107.894(4)73.556(2)109.243(2)
γ [°]64.608(3)9075.287(2)90
V [Å3]3489.67(17)6224.5(4)3541.10(14)5297.7(2)
Z2424
ρ calcd. [g·cm−3]1.3401.3851.1971.391
μ [mm−1]9.1867.0285.1893.997
Θ-range [°]3.705–62.6983.327–62.7483.238–62.7243.140–62.754
refl. coll.27,28724,02249,43921,589
indep. refl.11,102
[Rint = 0.0308]
9919
[Rint = 0.0550]
11,257
[Rint = 0.0564]
8461
[Rint = 0.0294]
data/restr./param.11102/0/6769919/0/69511257/0/7138461/0/676
final R indices
[I > 2σ(I)] a
0.0296, 0.07210.0510, 0.13880.0310, 0.07900.0321, 0.0852
R indices (all data)0.0349, 0.07390.0623, 0.14660.0335, 0.08070.0362, 0.0886
GooF b1.0371.0821.0231.032
Δρmax/min (e·Å−3)1.310/−1.2621.253/−0.7180.734/−0.8400.713/−0.518
a R1 = Σ||Fo| − |Fc||/Σ|Fo|, ωR2 = [Σ(Fo2Fc2)2/ΣωFo2]1/2. b GooF = [Σω(Fo2Fc2)2/(n − p)]1/2.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nährig, F.; Sun, Y.; Thiel, W.R. Phosphine Functionalized CpC Ligands and Their Metal Complexes. Chemistry 2023, 5, 912-933. https://doi.org/10.3390/chemistry5020062

AMA Style

Nährig F, Sun Y, Thiel WR. Phosphine Functionalized CpC Ligands and Their Metal Complexes. Chemistry. 2023; 5(2):912-933. https://doi.org/10.3390/chemistry5020062

Chicago/Turabian Style

Nährig, Florian, Yu Sun, and Werner R. Thiel. 2023. "Phosphine Functionalized CpC Ligands and Their Metal Complexes" Chemistry 5, no. 2: 912-933. https://doi.org/10.3390/chemistry5020062

Article Metrics

Back to TopTop