Next Article in Journal
Alkali-Activated Metakaolin as a Zeolite-Like Binder for the Production of Adsorbents
Next Article in Special Issue
Mono- and Dinuclear Aluminium Complexes Derived from Biguanide and Carbothiamide Ligands
Previous Article in Journal
Metal Hydride Compressors with Gas-Gap Heat Switches: Concept, Development, Testing, and Space Flight Operation for the Planck Sorption Cryocoolers
Previous Article in Special Issue
Syntheses and Structures of Novel λ33-Phosphanylalumanes Fully Bearing Carbon Substituents and Their Substituent Effects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A PAlP Pincer Ligand Bearing a 2-Diphenylphosphinophenoxy Backbone

Department of Material Chemistry, Graduate School of Engineering, Kyoto University, Nishikyo-ku, Kyoto 615-8510, Japan
*
Author to whom correspondence should be addressed.
Submission received: 15 August 2019 / Revised: 19 November 2019 / Accepted: 20 November 2019 / Published: 28 November 2019
(This article belongs to the Special Issue Organoaluminum Compounds)

Abstract

:
A PAlP pincer ligand derived from 2-diphenylphosphino-6-isopropylphenol was synthesized. The Lewis acidity of the Al center of the ligand was evaluated with coordination of (O)PEt3. A zwitterionic rhodium-aluminum heterobimetallic complex bearing the PAlP ligand was synthesized through its complexation with [RhCl(nbd)]2. Moreover, reduction of the zwitterionic rhodium-aluminum complex with KC8 afforded heterobimetallic complexes bearing an X-type PAlP pincer ligand.

Graphical Abstract

1. Introduction

Multidentate ligands containing an electropositive element have emerged recently to provide their transition-metal complexes with unique reactivity [1,2,3,4]. Among them, pincer ligands containing a group 13 element are of particular interest owing to their properties including: (1) electron-accepting ability as Z-type ligands through their vacant p orbitals and (2) high electron-donating ability as X-type ligands due to their low electronegativity [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. Those bearing Al have gained much attention due to having the lowest electronegativity of Al among the group 13 elements [36], which could give X-type ligands with high Lewis acidity and electron-donicity (Figure 1). In this context, we previously reported on PAlP pincer ligand 1 and its rhodium complexes, which catalyzed C2-selective mono-alkylation of pyridine with unactivated alkenes through C(2)–H activation of pyridine coordinating to the Al center by the Rh center (Figure 1c) [20]. Herein, we report on PAlP pincer ligand 2 derived from 2-diphenylphosphino-6-isopropylphenol, expecting its Al center to be more Lewis acidic than that in 1 due to the higher electronegativity of oxygen compared to that of nitrogen (Figure 1d).

2. Results and Discussion

PAlP pincer ligand 2 was prepared from commercially available 2-isopropylphenol (3) in three steps (Scheme 1). Methoxymethylation of 3 with methoxymethyl chloride afforded methoxymethyl (MOM)-protected 2-isopropylphenol 4. Lithiation of 4 at the position ortho to the methoxymethoxy (OMOM) group with n-BuLi followed by the reaction with Ph2PCl afforded 2-diphenylphosphino-6-isopropylphenol (5) in 61% yield. Finally, 2 was obtained by the reaction of Et2AlCl with two molar equivalents of 5 in 85% yield.
The Lewis acidity of 1 and 2 was evaluated and compared with common Al Lewis acids based on the method reported by Beckett (Scheme 2) [37,38]. The signal observed in 31P NMR spectrum of 2·(O)PEt3 was in a magnetic field slightly lower than that of 1·(O)PEt3, whereas those of the complexes from methylaluminum bis(2,6-di-tert-butyl-4-methylphenoxide) (MAD) [39,40] and AlMe3 appeared in a much higher magnetic field. The diphenylphosphino groups in 2·(O)PEt3 were observed equally, while the diisopropylphosphino groups in 1·(O)PEt3 were observed unequally. This would suggest that one diisopropylphosphino group in 1·(O)PEt3 coordinates to the Al center to afford penta- or hexacoordinated aluminum species. Although we cannot conclude whether 2 or 1 has a higher Lewis acidity due to the structural difference of the Al moiety between 2·(O)PEt3 and 1·(O)PEt3, the Lewis acidity of 2 is expected to be higher than MAD and AlMe3.
Coordination of 2 to transition metals to prepare heterobimetallic complexes was investigated next. As an example, the complexation of 2 with [RhCl(nbd)]2 in toluene afforded a rhodium-aluminum complex 6 as an orange precipitate in high yield (Scheme 3). Single crystals were obtained by slow diffusion of n-hexane into a concentrated benzene solution of 6. The crystal structure of 6 was revealed by X-ray diffraction analysis (Figure 2). In sharp contrast to the rhodium-aluminum complex bearing 1 as a supporting ligand, where the Al center acts as a neutral Z-type ligand [20], the Al moiety in 6 abstracts chlorine atom from the Rh center to generate a zwitterionic complex [5,12,28]. This could be attributed to the structural flexibility and/or the higher Lewis acidity of the Al center in 2 compared to that in 1.
A reduction of 6 with 2.2 equivalents of KC8 in the presence of 4-dimethylaminopyridine (DMAP) afforded heterobimetallic complexes 7 and 8 in 92% and 4% yield, respectively (Scheme 4). Recrystallization by slow diffusion of n-pentane to a saturated benzene solution of a mixture of 7 and 8 generated crystals of 8. An X-ray diffraction analysis of a crystal of 8 revealed that 2 became an X-type PAlP pincer ligand (Figure 3). Al in 8 would be formally served as an anionic Al(I) ligand, which is based on our knowledge that Al in an X-type PAlP pincer rhodium complex generated through a reduction of a rhodium complex bearing 1 by KC8 was suggested to be an anionic Al(I) ligand by DFT calculations [20]. In the crystal structure, a distance between Rh and Al is 2.318 Å, which is shorter than the sum of the Van der Waals radii of Rh and Al [41]. Coordination of DMAP to the Al center indicates a Lewis acidity of the Al center. Double bonds of a norbornadiene were bridged over two rhodium centers to stabilize the electron-rich rhodium centers. A coordination site trans to the Al is vacant, which would result from strong σ-electron-donicity of the X-type Al ligand. Although all attempts to obtain crystals of 7 are failed, high resolution ESI mass spectrometry and the fact that addition of nbd (11 μmol) to a benzene solution of 8 (9.0 μmol) afforded 7 would support the structure of 7 as a heterobimetallic complex in which the Rh is chelated by a nbd ligand [20,42].

3. Materials and Methods

3.1. General

All manipulations of oxygen- and moisture-sensitive materials were conducted with a standard schlenk technique under an argon atmosphere or in a glove box under a nitrogen atmosphere. Medium pressure liquid chromatography (MPLC) was performed using Kanto Chemical silica gel 60 (spherical, 40–50 μm) (KANTO CHEMICAL CO., INC., Tokyo, Japan). Analytical thin layer chromatography (TLC) (Merck KGaA, Darmstadt, Germany) was performed on Merck TLC silica gel 60 F254 (0.25 μm) plates. Visualization was accomplished with UV light (254 nm).

3.2. Apparatus

Proton, carbon, phosphorus, and aluminum nuclear magnetic resonance spectra (1H, 13C, 31P, and 27Al NMR) were recorded on a JEOL ECS-400 (1H NMR 400 MHz; 13C NMR 101 MHz; 31P NMR 162 MHz; and 27Al NMR 104 MHz) spectrometer) (JEOL, Tokyo, Japan) with solvent resonance as the internal standard (1H NMR, CDCl3 at 7.26 ppm, CD2Cl2 at 5.32 ppm, C6D6 at 7.16 ppm; 13C NMR, CDCl3 at 77.0 ppm, CD2Cl2 at 53.8 ppm, and C6D6 at 128.0 ppm). NMR data are reported as follows: chemical shift, multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, quint = quintet, sept = septet, br = broad, m = multiplet, and vt = virtual triplet), coupling constants (Hz), and integration. High resolution mass spectra were obtained with Thermo Scientific Exactive (ESI or APCI). Medium pressure liquid chromatography (MPLC) was performed with a Yamazen EPLC-W-Prep 2XY or SHOKO SCIENTIFIC Purif-espoir2. Elemental analyses were performed on a J-Science Micro corder JM10, JM11, and a YANACO Micro corder MT-6.

3.3. Chemicals

Unless otherwise noted, commercially available chemicals were used without further purification. (Rh(nbd)(μ-Cl))2 (nbd = 2,5-norbornadiene) was synthesized following the reported procedure [43]. Anhydrous n-hexane, toluene, and THF were purchased from Kanto Chemical and purified by passage through activated alumina under positive argon pressure as described by Grubbs et al. [44]. Anhydrous n-pentane and benzene were purchased from FUJIFILM Wako Pure Chemical Corporation, and used after further dehydration with activated MS4Å.

3.4. Synthesis of 1-Isopropyl-2-Methoxymethoxybenzene (4)

To a suspension of sodium hydride (1.3 g, 55 mmol) in THF (20 mL), 2-isoprorylphenol (3, 5.0 g, 37 mmol) in THF (20 mL) was added dropwise at 0 °C and the mixture was stirred for 10 min at the same temperature. To the resulting mixture, chloromethyl methyl ether (3.6 mL, 48 mmol) was added dropwise, and the mixture was stirred for 150 min at room temperature. After being quenched with H2O, the mixture was extracted with Et2O, and the combined organic layers were washed with 3 M NaOH aq. and dried over anhydrous MgSO4. After filtration through a KIRIYAMA filter paper, the filtrate was concentrated in vacuo to give a colorless liquid (6.5 g). The product was used in the next step without further purification. 1H NMR (400 MHz, CDCl3): δ 1.27 (d, J = 6.9 Hz, 6H), 3.40 (sept, J = 6.9 Hz, 1H), 3.53 (s, 3H), 5.24 (s, 2H), 7.02 (dd, J = 7.3, 7.3 Hz, 1H), 7.10 (d, J = 7.8 Hz, 1H), 7.17 (dd, J = 7.8, 7.3 Hz, 1H), and 7.27 (d, J = 7.3 Hz, 1H). 13C{1H} NMR (101 MHz, CDCl3): δ 22.8, 26.8, 56.0, 94.4, 113.9, 121.8, 126.1, 126.6, 137.5, and 154.3. All the resonances of 1H and 13C NMR spectra of the product were consistent with the reported values [45].

3.5. Synthesis of 2-Diphenylphosphino-6-Isopropylphenol (5)

To a solution of 4 (6.6 g, 37 mmol) in Et2O (120 mL) we added tetramethylethylenediamine (5.5 mL, 37 mmol) and a solution of n-BuLi in n-hexane (1.6 M, 23 mL, 37 mmol) at −78 °C under an Ar atmosphere. The reaction mixture was allowed to warm up to room temperature and stirred for 12 h. To the mixture, chlorodiphenylphosphine (6.8 mL, 37 mmol) was added dropwise at 0 °C and the resulting mixture was stirred for 3 h at room temperature. All of the volatiles were removed under reduced pressure, and THF (40 mL) and 5 M HCl (40 mL) were added to the residue. The mixture was heated at 50 °C for 24 h, then neutralized using an aqueous NH3 solution. The aqueous layer was extracted with Et2O, and the combined organic layers were washed with water and dried over anhydrous MgSO4. After filtration through a KIRIYAMA filter paper, the filtrate was concentrated under reduced pressure to give a pale yellow viscous liquid. The residue was purified by MPLC (ethyl acetate:n-hexane = 1:99) on silica gel to give a colorless viscous oil (7.2 g, 22 mmol, 61%). 1H NMR (400 MHz, CDCl3): δ 1.27 (d, J = 6.9 Hz, 6H), 3.33 (sept, J = 6.9 Hz, 1H), 6.54 (d, J = 8.7 Hz, 1H), 6.84–6.89 (m, 2H), 7.25 (t, J = 3.1 Hz, 1H), and 7.33–7.41 (m, 10H). 13C{1H} NMR (101 MHz, CDCl3): δ 22.5, 27.2, 120.1, 120.8 (d, J = 3.5 Hz), 128.3, 128.6 (d, J = 6.9 Hz), 128.9, 132.0 (d, J = 2.3 Hz), 133.4 (d, J = 18.5 Hz), 134.9, 135.1 (d, J = 4.6 Hz), and 156.7 (d, J = 19.7 Hz). 31P{1H} NMR (162 MHz, CDCl3): δ −30.2. HRMS–[APCI(+)] (m/z): [M + H]+ calcd for C21H22OP, 321.1403; found, 321.1407.

3.6. Synthesis of Bis(2-(Diphenylphosphino)-6-Isopropylphenoxy)-Aluminum Chloride (2)

To a solution of 5 (1.5 g, 4.7 mmol) in toluene (10 mL) we added a solution of diethylaluminum chloride in n-hexane (0.87 M, 2.7 mL, 2.3 mmol). The resulting solution was stirred for 12 h at room temperature, and all of the volatiles were removed in vacuo to afford the desired product as a white solid (1.4 g, 2.0 mmol, 85% yield). There would be two kinds of structural isomers (major:minor = 71:29). 1H NMR (400 MHz, C6D6): δ major: 1.08 (d, J = 6.9 Hz, 12H), 3.13 (sept, J = 6.7 Hz, 2H), 6.76 (dd, J = 7.6, 7.6 Hz, 2H), 6.88–7.00 (m, 14H), 7.10 (t, J = 7.1 Hz, 2H), 7.24 (d, J = 7.3 Hz, 2H), 7.31 (br, 2H), 7.72 (br, 4H). minor: 1.12 (d, J = 6.9 Hz, 12H), 3.26–3.34 (m, 2H), 6.72 (dd, J = 7.6, 7.6 Hz, 2H), and 7.21 (d, J = 7.3 Hz, 2H). Other peaks were overlapped with those of the major isomer. 13C{1H} NMR (101 MHz, C6D6): δ 23.0, 23.4, 26.96, 26.98, 117.0, 117.4, 119.5, 120.1 (d, J = 4.6 Hz), 127.9, 128.4 (d, J = 8.1 Hz), 128.5, 128.8 (d, J = 9.3 Hz), 128.9, 129.0, 130.0, 130.8, 131.1, 131.3, 131.4, 134.0 (d, J = 11.6 Hz), 138.5, 138.7 (d, J = 3.5 Hz), 162.6 (d, J = 23.1 Hz), 163.4 (d, J = 8.1 Hz), and 163.6 (d, J = 8.1 Hz). 31P{1H} NMR (162 MHz, C6D6): δ −30.5 (minor), and −37.2 (major). 27Al{1H} NMR (104 MHz, C6D6): δ 65.6 (br) m.p. 119 °C.

3.7. Synthesis of 6

To a solution of (Rh(nbd)(μ-Cl))2 (660 mg, 1.4 mmol) in toluene (5.0 mL), a solution of 2 (2.0 g, 2.9 mmol) in toluene (10 mL) was added at 10 °C. The reaction mixture was stirred for 20 h at 10 °C to give an orange precipitate. The orange precipitate was collected by filtration and washed with toluene (1.0 mL × 3) and followed by n-pentane (2.0 mL × 3). After being dried under reduced pressure, 6 was obtained as an orange solid (1.3 g, 1.4 mmol, 49%). Orange crystals for elemental analysis and X-ray diffraction analysis were obtained by a slow diffusion of n-hexane to a concentrated benzene solution of 6 at room temperature. 1H NMR (400 MHz, CD2Cl2): δ 1.15 (d, J = 6.4 Hz, 6H), 1.16 (d, J = 6.9 Hz, 6H), 1.38 (s, 2H), 3.59 (sept, J = 6.6 Hz, 2H), 3.76 (s, 2H), 4.00 (br s, 4H), 6.29–6.32 (m, 2H), 6.57 (t, J = 7.6 Hz, 2H), 7.01 (t, J = 7.3 Hz, 4H), 7.11 (d, J = 7.3 Hz, 2H), 7.19 (t, J = 7.3 Hz, 2H), 7.31–7.36 (m, 4H), 7.43–7.50 (m, 6H), and 7.59–7.63 (m, 4H). 13C{1H} NMR (101 MHz, CD2Cl2): δ 22.6, 24.4, 26.8, 51.7, 66.1, 74.8, 118.9 (vt, J = 118.9 Hz), 119.2 (vt, J = 4.0 Hz), 128.1 (vt, J = 5.2 Hz), 128.5 (vt, J = 4.6 Hz), 129.4, 129.7, 130.1, 130.5, 131.4 (vt, J = 21.4 Hz), 132.6 (vt, J = 21.4 Hz), 134.7 (vt, J = 6.9 Hz), 135.5 (vt, J = 4.6 Hz), 139.4, and 157.4 (vt, J = 4.0 Hz). 31P{1H} NMR (162 MHz, CD2Cl2): δ 23.3 (JP–Rh = 152.6 Hz). 27Al{1H} NMR (104 MHz, CD2Cl2): δ 67.3 (br) Anal. Calcd C49H48AlCl2O2P2Rh: C, 63.17; H, 5.19. Found: C, 63.00; H, 5.24. HRMS–[ESI(+)] (m/z): [M+Na]+ calcd for C49H48AlCl2O2P2RhNa, 953.1269; found, 953.1273. m.p. 157 °C (decomp).

3.8. Synthesis of 7 and 8

To a THF (4 mL) solution of 6 (50 mg, 50 μmol) and 4-dimethylaminopyridine (7.0 mg, 50 μmol), a suspension of potassium graphite (16 mg, 0.12 mmol) in THF (3 mL) was slowly added at −78 °C. The reaction mixture was stirred for 10 h at room temperature. THF was evaporated and benzene (10 mL) was added to the mixture. After filtration through a KIRIYAMA filter paper, the filtrate was concentrated to afford a brown precipitate. The precipitate was washed with n-pentane (3.0 mL × 2) and dried under reduced pressure to give X-type rhodium-aluminum bimetallic complexes 7 and 8 as a brown solid in 92% and 4% yield (total 48 mg, 48 μmol, 96%). Orange crystals of 8 for X-ray diffraction analysis were obtained by a slow diffusion of n-pentane to the concentrated benzene solution of a mixture of 7 and 8 at room temperature. All our attempts to obtain crystals of 7 failed.
7: 1H NMR (400 MHz, C6D6): δ 1.13 (d, J = 7.3 Hz, 6H), 1.29 (d, J = 7.3 Hz, 1H), 1.32 (d, J = 6.4 Hz, 6H), 1.44 (d, J = 7.3 Hz, 1H), 1.77 (s, 6H), 2.90 (s, 2H), 3.53 (br s, 2H), 3.58 (sept, J = 6.7 Hz, 2H), 4.62 (br s, 2H), 5.54 (d, J = 7.3 Hz, 2H), 6.81 (t, J = 7.8 Hz, 2H), 6.89–6.97 (m, 6H), 7.10 (t, J = 7.3 Hz, 2H), 7.15–7.21 (m, 6H), 7.24 (d, J = 7.3 Hz, 2H), 7.49–7.56 (m, 6H), and 8.07–8.12 (m, 4H). 13C{1H} NMR (101 MHz, C6D6): δ 22.8, 23.5, 27.3, 37.9, 40.8 (vt, J = 8.1 Hz), 40.9 (vt, J = 7.5 Hz), 49.2, 63.4, 98.1, 105.7, 118.0, 123.9 (vt, J = 23.7 Hz), 127.0, 127.2, 127.4, 127.9, 130.7, 133.1 (vt, J = 6.4 Hz), 134.4 (vt, J = 7.5 Hz), 138.4, 139.4 (vt, J = 12.1 Hz), 145.4 (vt, J = 16.8 Hz), 146.1, 154.8, and 161.3 (vt, J = 5.8 Hz), an sp2 carbon signal overlaps with others. 31P{1H} NMR (162 MHz, C6D6): δ 27.0 (JP–Rh = 157.0 Hz). 27Al{1H} NMR (104 MHz, C6D6): it is too broad to identify the corresponding signal. Elemental analysis did not give a satisfactory result due to instability of 7 toward O2 and H2O. HRMS–[ESI(+)] (m/z): [M]+ calcd for C56H58AlN2O2P2Rh, 982.2838; found, 982.2834. m.p. 165 °C (decomp).
8: 1H NMR (400 MHz, C6D6): δ −0.69 (s, 2H), 1.19 (d, J = 6.4 Hz, 12H), 1.29 (d, J = 6.9 Hz, 12H), 1.84 (s, 12H), 3.04 (s, 4H), 3.07 (s, 2H), 3.49 (sept, J = 6.6 Hz, 4H), 5.16 (d, J = 6.0 Hz, 4H), 6.76 (t, J = 7.3 Hz, 4H), 6.93–6.96 (m, 22H), 7.12–7.17 (m, 6H), 7.20 (d, J = 7.3 Hz, 4H), 7.47 (d, J = 6.0 Hz, 4H), 7.50–7.57 (m, 8H), and 7.83–7.91 (m, 8H). 13C{1H} NMR (101 MHz, C6D6): δ 22.9, 23.3, 27.7, 38.1, 47.0, 58.9, 67.1 (br), 105.7, 118.1, 124.7, 125.2, 126.8, 130.4, 133.4 (br), 134.4 (vt, J = 6.9 Hz), 137.8, 140.5 (vt, J = 22.5 Hz), 146.2, 154.8, and 160.8 (vt, J = 6.4 Hz). 31P{1H} NMR (162 MHz, C6D6): δ 24.1 (JP–Rh = 139.5 Hz). 27Al{1H} NMR (104 MHz, C6D6): it is too broad to identify the corresponding signal. Anal. Calcd C105H108Al2N4O4P4Rh2: C, 67.31; H, 5.81; N, 2.99. Found: C, 66.35; H, 5.82; N, 2.50. Elemental analysis did not give a satisfactory result due to instability of 8 toward O2 and H2O. HRMS–[ESI(+)] (m/z): [M]+ calcd for C105H108Al2N4O4P4Rh2, 1872.5056; found, 1872.5005. m.p. 169 °C (decomp).

3.9. Evaluation of Lewis Acidity of Aluminum Compounds

In a glove box, aluminum compound (10 μmol) and the triethylphosphine oxide (10 μmol) were placed together in a J-young NMR tube and dissolved in C6D6 (500 μL). The resulting mixture was analyzed by 31P NMR spectroscopy with a capillary sealing H3PO4 (85 wt % in H2O, diluted with 100-fold D2O) as an internal standard (0.70 ppm). In the case of 2, a set of signals probably derived from a structural isomer (~10%) was also observed in the 31P spectrum of 2·(O)PEt3.

3.10. X-Ray Diffraction Study and X-Ray Crystallographic Analysis

The crystals of 6 and 8 were mounted on the CryoLoop (Hampton Research Corp., California, USA) with a layer of light mineral oil and placed in a nitrogen stream at 143(1) K. The X-ray structural determinations of 6 were performed on a Rigaku/Saturn70 CCD diffractometers using graphite-monochromated Mo Kα radiation (λ = 0.71070 Å) at 153 K, and processed using CrystalClear (Rigaku, Tokyo, Japan) [46,47]. The X-ray structural determinations of 8 were performed on a Rigaku/Saturn724+ CCD diffractometers using graphite-monochromated Mo Kα radiation (λ = 0.71075 Å) at 153 K, and processed using CrystalClear (Rigaku) [46,47]. The structures were solved by a direct method and refined by full-matrix least-square refinement on F2. The structures of 6 and 8 were solved by direct methods (SHELXS-97 and SHELX 2018) [48]. Non-hydrogen atoms were anisotropically refined except for solvent atoms due to their disorder. Hydrogen atoms were included in the refinement on calculated positions riding on their carrier atoms. CCDC 1946453 (for 6) and 1963268 (for 8) contain the supplementary crystallographic data. These data can be obtained from The Cambridge Crystallographic Data Centre.

4. Conclusions

In conclusion, a new Al tethered diphosphine ligand 2 derived from 2-diphenylphosphino-6-isopropylphenol has been developed. The utility of 2 as a supporting ligand was demonstrated through complexation of 2 with [RhCl(nbd)]2 to give zwitterionic complex 6. Moreover, reduction of 6 with KC8 afforded heterobimetallic complexes 7 and 8 bearing an X-type PAlP pincer ligand. Further applications of 2 as a supporting ligand for transition-metal catalysis are under investigation.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/7/12/140/s1. CIF Files and CheckCIF Files, The detailed experimental procedures and NMR spectra of each compounds. CCDC 1946453 (for 6) and 1963268 (for 8) contain the supplementary crystallographic data for this paper. These data can be obtained free of via http://www.ccdc.cam.ac.uk/conts/retrieving.html.

Author Contributions

K.S. and I.F. planned, conducted and analysed experiments. K.S. and Y.N. cowrote the manuscript. Y.N. directed the project.

Funding

This research was supported by the CREST program “Establishment of Molecular Technology towards the Creation of New Functions” (JPMJCR14L3) from the JST.

Acknowledgments

We thank Naofumi Hara for X-ray crystallographic analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fontaine, F.G.; Boudreau, J.; Thibault, M.H. Coordination Chemistry of Neutral (Ln)–Z Amphoteric and Ambiphilic Ligands. Eur. J. Inorg. Chem. 2008, 2008, 5439–5454. [Google Scholar] [CrossRef]
  2. Amgoune, A.; Bourissou, D. σ-Acceptor, Z-type Ligands for Transition Metals. Chem. Commun. 2011, 47, 859–871. [Google Scholar] [CrossRef] [PubMed]
  3. Bouhadir, G.; Bourissou, D. Complexes of Ambiphilic Ligands: Reactivity and Catalytic Applications. Chem. Soc. Rev. 2016, 45, 1065–1079. [Google Scholar] [CrossRef]
  4. Yamashita, M. The Organometallic Chemistry of Boron-Containing Pincer Ligands based on Diazaboroles and Carboranes. Bull. Chem. Soc. Jpn. 2016, 89, 269–281. [Google Scholar] [CrossRef]
  5. Sircoglou, M.; Bouhadir, G.; Saffon, N.; Miqueu, K.; Bourissou, D. A Zwitterionic Gold(I) Complex from an Ambiphilic Diphosphino-Alane Ligand. Organometallics 2008, 27, 1675–1678. [Google Scholar] [CrossRef]
  6. Sircoglou, M.; Mercy, M.; Saffon, N.; Coppel, Y.; Bouhadir, G.; Maron, L.; Bourissou, D. Gold(I) Complexes of Phosphanyl Gallanes: From Interconverting to Separable Coordination Isomers. Angew. Chem. Int. Ed. 2009, 48, 3454–3457. [Google Scholar] [CrossRef]
  7. Segawa, Y.; Yamashita, M.; Nozaki, K. Syntheses of PBP Pincer Iridium Complexes: A Supporting Boryl Ligand. J. Am. Chem. Soc. 2009, 131, 9201–9203. [Google Scholar] [CrossRef]
  8. Spokoyny, A.M.; Reuter, M.G.; Stern, C.L.; Ratner, M.A.; Seideman, T.; Mirkin, C.A. Carborane-Based Pincers: Synthesis and Structure of SeBSe and SBS Pd(II) Complexes. J. Am. Chem. Soc. 2009, 131, 9482–9483. [Google Scholar] [CrossRef]
  9. Harman, W.H.; Peters, J.C. Reversible H2 Addition across a Nickel–Borane Unit as a Promising Strategy for Catalysis. J. Am. Chem. Soc. 2012, 134, 5080–5082. [Google Scholar] [CrossRef] [PubMed]
  10. Hasegawa, M.; Segawa, Y.; Yamashita, M.; Nozaki, K. Isolation of a PBP-Pincer Rhodium Complex Stabilized by an Intermolecular C–H σ Coordination as the Fourth Ligand. Angew. Chem. Int. Ed. 2012, 51, 6956–6960. [Google Scholar] [CrossRef] [PubMed]
  11. Masuda, Y.; Hasegawa, M.; Yamashita, M.; Nozaki, K.; Ishida, N.; Murakami, M. Oxidative Addition of a Strained C–C Bond onto Electron-Rich Rhodium(I) at Room Temperature. J. Am. Chem. Soc. 2013, 135, 7142–7145. [Google Scholar] [CrossRef] [PubMed]
  12. Sircoglou, M.; Saffon, N.; Miqueu, K.; Bouhadir, G.; Bourissou, D. Activation of M–Cl Bonds with Phosphine–Alanes: Preparation and Characterization of Zwitterionic Gold and Copper Complexes. Organometallics 2013, 32, 6780–6784. [Google Scholar] [CrossRef]
  13. Lin, T.P.; Peters, J.C. Boryl–Metal Bonds Facilitate Cobalt/Nickel-Catalyzed Olefin Hydrogenation. J. Am. Chem. Soc. 2014, 136, 13672–13683. [Google Scholar] [CrossRef] [PubMed]
  14. Harman, W.H.; Lin, T.P.; Peters, J.C. A d10 Ni–(H2) Adduct as an Intermediate in H–H Oxidative Addition across a Ni–B Bond. Angew. Chem. Int. Ed. 2014, 53, 1081–1086. [Google Scholar] [CrossRef]
  15. Cowie, B.E.; Tsao, F.A.; Emslie, D.J.H. Synthesis and Platinum Complexes of an Alane-Appended 1,1’-Bis(phosphino)ferrocene Ligand. Angew. Chem. Int. Ed. 2015, 54, 2165–2169. [Google Scholar] [CrossRef]
  16. Shih, W.C.; Gu, W.; MacInnis, M.C.; Timpa, S.D.; Bhuvanesh, N.; Zhou, J.; Ozerov, O.V. Facile Insertion of Rh and Ir into a Boron–Phenyl Bond, Leading to Boryl/Bis(phosphine) PBP Pincer Complexes. J. Am. Chem. Soc. 2016, 138, 2086–2089. [Google Scholar] [CrossRef]
  17. Shih, W.C.; Ozerov, O.V. Selective ortho C–H Activation of Pyridines Directed by Lewis Acidic Boron of PBP Pincer Iridium Complexes. J. Am. Chem. Soc. 2017, 139, 17297–17300. [Google Scholar] [CrossRef]
  18. Takaya, J.; Iwasawa, N. Synthesis, Structure, and Catalysis of Palladium Complexes Bearing a Group 13 Metalloligand: Remarkable Effect of an Aluminum-Metalloligand in Hydrosilylation of CO2. J. Am. Chem. Soc. 2017, 139, 6074–6077. [Google Scholar] [CrossRef]
  19. Saito, T.; Hara, N.; Nakao, Y. Palladium Complexes Bearing Z-type PAlP Pincer Ligands. Chem. Lett. 2017, 46, 1247–1249. [Google Scholar] [CrossRef]
  20. Hara, N.; Saito, T.; Semba, K.; Kuriakose, N.; Zheng, H.; Sakaki, S.; Nakao, Y. Rhodium Complexes Bearing PAlP Pincer Ligands. J. Am. Chem. Soc. 2018, 140, 7070–7073. [Google Scholar] [CrossRef]
  21. Saito, N.; Takaya, J.; Iwasawa, N. Stabilized Gallylene in a Pincer-Type Ligand: Synthesis, Structure, and Reactivity of PGaIP-Ir Complexes. Angew. Chem. Int. Ed. 2019, 58, 9998–10002. [Google Scholar] [CrossRef] [PubMed]
  22. Burlitch, J.M.; Leonowicz, M.E.; Petersen, R.B.; Hughes, R.E. Coordination of Metal Carbonyl Anions to Triphenyaluminum, -gallium, and -indium and the Crystal Structure of Tetraethylammonium Triphenyl((η5-cyclopentadienyl)dicarbonyliron)aluminate (Fe–Al). Inorg. Chem. 1979, 18, 1097–1105. [Google Scholar] [CrossRef]
  23. Golden, J.T.; Peterson, T.H.; Holland, P.L.; Bergman, R.G.; Andersen, R.A. Adduct Formation and Single and Double Deprotonation of Cp*(PMe3)Ir(H)2 with Main Group Metal Alkyls and Aryls: Synthesis and Structure of Three Novel Ir–Al and Ir–Mg Heterobimetallics. J. Am. Chem. Soc. 1998, 120, 223–224. [Google Scholar] [CrossRef]
  24. Braunschweig, H.; Gruss, K.; Radacki, K. Interaction between d- and p-Block Metals: Synthesis and Structure of Platinum–Alane Adducts. Angew. Chem. Int. Ed. 2007, 46, 7782–7784. [Google Scholar] [CrossRef]
  25. Bauer, J.; Braunschweig, H.; Brenner, P.; Kraft, K.; Radacki, K.; Schwab, K. Late-Transition-Metal Complexes as Tunable Lewis Bases. Chem. Eur. J. 2010, 16, 11985–11992. [Google Scholar] [CrossRef]
  26. Derrah, E.J.; Sircoglou, M.; Mercy, M.; Ladeira, S.; Bouhadir, G.; Miqueu, K.; Maron, L.; Bourissou, D. Original Transition Metal→Indium Interactions upon Coordination of a Triphosphine-Indane. Organometallics 2011, 30, 657–660. [Google Scholar] [CrossRef]
  27. Rudd, P.A.; Liu, S.; Gagliardi, L.; Young V.G., Jr.; Lu, C.C. Metal–Alane Adducts with Zero-Valent Nickel, Cobalt, and Iron. J. Am. Chem. Soc. 2011, 133, 20724–20727. [Google Scholar] [CrossRef]
  28. Devillard, M.; Nicolas, E.; Appelt, C.; Backs, J.; Mallet-Ladeira, S.; Bouhadir, G.; Slootweg, J.C.; Uhl, W.; Bourissou, D. Novel Zwitterionic Complexes Arising from the Coordination of an Ambiphilic Phosphorus–Aluminum Ligand to Gold. Chem. Commun. 2014, 50, 14805–14808. [Google Scholar] [CrossRef]
  29. Devillard, M.; Nicolas, E.; Ehlers, A.W.; Backs, J.; Mallet-Ladeira, S.; Bouhadir, G.; Slootweg, J.C.; Uhl, W.; Bourissou, D. Dative Au–Al Interactions: Crystallographic Characterization and Computational Analysis. Chem. Eur. J. 2015, 21, 74–79. [Google Scholar] [CrossRef]
  30. Wang, G.; Xu, L.; Li, P. Double N,B-Type Bidentate Boryl Ligands Enabling a Highly Active Iridium Catalyst for C–H Borylation. J. Am. Chem. Soc. 2015, 137, 8058–8061. [Google Scholar] [CrossRef]
  31. Devillard, M.; Declercq, R.; Nicolas, E.; Ehlers, A.W.; Backs, J.; Saffon-Merceron, N.; Bouhadir, G.; Slootweg, J.C.; Uhl, W.; Bourissou, D. A Significant but Constrained Geometry Pt–Al Interaction: Fixation of CO2 and CS2, Activation of H2 and PhCONH2. J. Am. Chem. Soc. 2016, 138, 4917–4926. [Google Scholar] [CrossRef] [PubMed]
  32. Cammarota, R.C.; Vollmer, M.V.; Xie, J.; Ye, J.; Linehan, J.C.; Burgess, S.A.; Appel, A.M.; Gagliardi, L.; Lu, C.C. A Bimetallic Nickel–Gallium Complex Catalyzes CO2 Hydrogenation via the Intermediacy of an Anionic d10 Nickel Hydride. J. Am. Chem. Soc. 2017, 139, 14244–14250. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, G.; Liu, L.; Wang, H.; Ding, Y.S.; Zhou, J.; Mao, S.; Li, P. N,B-Bidentate Boryl Ligand-Supported Iridium Catalyst for Efficient Functional-Group-Directed C–H Borylation. J. Am. Chem. Soc. 2017, 139, 91–94. [Google Scholar] [CrossRef] [PubMed]
  34. Shi, Y.; Gao, Q.; Xu, S. Chiral Bidentate Boryl Ligand Enabled Iridium-Catalyzed Enantioselective C(sp3)–H Borylation of Cyclopropanes. J. Am. Chem. Soc. 2019, 141, 10599–10604. [Google Scholar] [CrossRef]
  35. Hicks, J.; Mansikkamäki, A.; Vasko, P.; Goicoechea, J.M.; Aldridge, S. A Nucleophilic Gold Complex. Nat. Chem. 2019, 11, 237–241. [Google Scholar]
  36. Allred, A.L.; Rochow, E.G. A Scale of Electronegativity Based on Electrostatic Force. J. Inorg. Nucl. Chem. 1958, 5, 264–268. [Google Scholar] [CrossRef]
  37. Beckett, M.A.; Strickland, G.C.; Holland, J.R.; Varma, K.S. A Convenient n.m.r. Method for the Measurement of Lewis Acidity at Boron Centres: Correlation of Reaction Rates of Lewis Acid Initiated Epoxide Polymerizations with Lewis Acidity. Polymer 1996, 37, 4629–4631. [Google Scholar] [CrossRef]
  38. Beckett, M.A.; Brassington, D.S.; Coles, S.J.; Hursthouse, M.B. Lewis Acidity of Tris(pentafluorophenyl)borane: Crystal and Molecular Structure of B(C6F5)3·OPEt3. Inorg. Chem. Commun. 2000, 3, 530–533. [Google Scholar] [CrossRef]
  39. Maruoka, K.; Itoh, T.; Yamamoto, H. Methylaluminum Bis(2,6-di-tert-butyl-4-alkylphenoxide). A New Reagent for Obtaining Unusual Equatorial and Anti-Cram Selectivity in Carbonyl Alkylation. J. Am. Chem. Soc. 1985, 107, 4573–4576. [Google Scholar] [CrossRef]
  40. Maruoka, K.; Itoh, T.; Sakurai, M.; Nonoshita, K.; Yamamoto, H. Amphiphilic Reactions by Means of Exceptionally Bulky Organoaluminum Reagents. Rational Approach for Obtaining Unusual Equatorial, Anti-Cram, and 1,4 Selectivity in Carbonyl Alkylation. J. Am. Chem. Soc. 1988, 110, 3588–3597. [Google Scholar] [CrossRef]
  41. Batsanov, S.S. Van der Waals Radii of Elements. Inorg. Mater. 2001, 37, 871–885. [Google Scholar] [CrossRef]
  42. Wassenaar, J.; Siegler, M.A.; Spek, A.L.; Bruin, B.; Reek, J.N.H.; Vlugt, J.I. Versatile New C3-Symmetric Tripodal Tetraphosphine Ligands; Structural Flexibility to Stabilize CuI and RhI Species and Tune Their Reactivity. Inorg. Chem. 2010, 49, 6495–6508. [Google Scholar] [CrossRef] [PubMed]
  43. Abel, E.W.; Bennett, M.A.; Wilkinson, G. Norbornadiene–Metal Complexes and Some Related Compounds. J. Chem. Soc. 1959, 3178–3182. [Google Scholar] [CrossRef]
  44. Pangborn, A.B.; Giardello, M.A.; Grubbs, R.H.; Rosen, R.K.; Timmers, F.J. Safe and Convenient Procedure for Solvent Purification. Organometallics 1996, 15, 1518–1520. [Google Scholar] [CrossRef]
  45. Ihori, Y.; Yamashita, Y.; Ishitani, H.; Kobayashi, S. Chiral Zirconium Catalysts Using Multidentate BINOL Derivatives for Catalytic Enantioselective Mannich-Type Reactions; Ligand Optimization and Approaches to Elucidation of the Catalyst Structure. J. Am. Chem. Soc. 2005, 127, 15528–15535. [Google Scholar] [CrossRef]
  46. Pflugrath, J.W. Rigaku Corporation, 1999, and CrystalClear Software User’s Guide, Molecular Structure Corporation, 2000. Acta Crystallogr. D 1999, 55, 1718–1725. [Google Scholar] [CrossRef] [Green Version]
  47. Pflugrath, J.W. The Finer Things in X-ray Diffraction Data Collection. Acta Crystallogr. 1999, 55, 1718–1725. [Google Scholar] [CrossRef] [Green Version]
  48. Sheldrick, G.M. A Short History of SHELX. Acta Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Reported pincer ligands containing an Al group in [15] (a), [18] (b), [20] (c) and this work (d).
Figure 1. Reported pincer ligands containing an Al group in [15] (a), [18] (b), [20] (c) and this work (d).
Inorganics 07 00140 g001
Scheme 1. The preparation routine of PAlP pincer ligand 2.
Scheme 1. The preparation routine of PAlP pincer ligand 2.
Inorganics 07 00140 sch001
Scheme 2. Evaluation of Lewis Acidity.
Scheme 2. Evaluation of Lewis Acidity.
Inorganics 07 00140 sch002
Scheme 3. Preparation of a rhodium-aluminum complex 6.
Scheme 3. Preparation of a rhodium-aluminum complex 6.
Inorganics 07 00140 sch003
Figure 2. The crystal structure of 6; (a) side view, (b) top view; thermal ellipsoids set at 30% probability; H atoms omitted for clarity.
Figure 2. The crystal structure of 6; (a) side view, (b) top view; thermal ellipsoids set at 30% probability; H atoms omitted for clarity.
Inorganics 07 00140 g002
Scheme 4. Reduction of 6 to prepare 7 and 8.
Scheme 4. Reduction of 6 to prepare 7 and 8.
Inorganics 07 00140 sch004
Figure 3. The crystal structure of 8; thermal ellipsoids set at 30% probability; carbon atoms on phosphorus atoms described in wireframes and H atoms omitted for clarity.
Figure 3. The crystal structure of 8; thermal ellipsoids set at 30% probability; carbon atoms on phosphorus atoms described in wireframes and H atoms omitted for clarity.
Inorganics 07 00140 g003

Share and Cite

MDPI and ACS Style

Semba, K.; Fujii, I.; Nakao, Y. A PAlP Pincer Ligand Bearing a 2-Diphenylphosphinophenoxy Backbone. Inorganics 2019, 7, 140. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7120140

AMA Style

Semba K, Fujii I, Nakao Y. A PAlP Pincer Ligand Bearing a 2-Diphenylphosphinophenoxy Backbone. Inorganics. 2019; 7(12):140. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7120140

Chicago/Turabian Style

Semba, Kazuhiko, Ikuya Fujii, and Yoshiaki Nakao. 2019. "A PAlP Pincer Ligand Bearing a 2-Diphenylphosphinophenoxy Backbone" Inorganics 7, no. 12: 140. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7120140

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop