Next Article in Journal
Cysteine Derivatized 99mTc-Labelled Fatty Acids as β-Oxidation Markers
Next Article in Special Issue
A PAlP Pincer Ligand Bearing a 2-Diphenylphosphinophenoxy Backbone
Previous Article in Journal
Bioinorganic Chemistry of Nickel
Previous Article in Special Issue
Reaction of Dialumane Incorporating Bulky Eind Groups with Pyridines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Syntheses and Structures of Novel λ33-Phosphanylalumanes Fully Bearing Carbon Substituents and Their Substituent Effects

by
Tatsuya Yanagisawa
1,
Yoshiyuki Mizuhata
1,2,* and
Norihiro Tokitoh
1,2,*
1
Institute for Chemical Research, Kyoto University, Gokasho, Uji, Kyoto 611-0011, Japan
2
Integrated Research Consortium on Chemical Sciences, Uji, Kyoto 611-0011, Japan
*
Authors to whom correspondence should be addressed.
Submission received: 16 October 2019 / Revised: 5 November 2019 / Accepted: 5 November 2019 / Published: 7 November 2019
(This article belongs to the Special Issue Organoaluminum Compounds)

Abstract

:
The novel phosphanylalumanes, Al–P single-bond species, fully bearing carbon protecting groups on aluminum and phosphorus atoms, are synthesized by the reactions of aluminum monohalides [(t-Bu)2AlBr and (C6F5)2AlCl·0.5(toluene)] with Mes2PLi. Regarding the t-Bu system, λ33-phosphanylalumane is obtained. Concerning the C6F5 system, on the other hand, the corresponding LiCl complex, λ44-phosphanylalumane, is obtained. The Al–P bond lengths of C6F5-substituted λ34-, and λ44-derivatives are much shorter than those of the reported λ34-phosphanylalumane derivatives and comparable to that observed for the previously reported λ33-phosphanylalumanes. Theoretical calculations reveal that the binding of the C6F5 groups to Al results in a large contribution of Al and a large s-character in the Al–P bond of phosphanylalumanes. Considering t-Bu-substituted phosphanylalumanes, the Al–P bond lengths reflect the coordination number of Al, showing a longer Al–P bond length in the case of λ4-Al as compared with that of λ3-Al. Combining the structural, spectroscopic, and theoretical results, the t-Bu-substituted λ33-phosphanylalumane has well separated vacant p orbital and lone pairs, which is suitable for reactivity studies.

Graphical Abstract

1. Introduction

The bonding between group 13 (E) and group 15 (Pn) elements formulated as R2E–PnR2 have attracted much attention due to their relationship, including the vacant p orbital on E and the lone-pair electrons on Pn. Aminoboranes (R2B–NR2), for example, have an enormous number of researches describing their B=N+ polar double-bond character [1]. To contrast, the heavier analogues, λ33-phosphanylalumanes (R2Al–PR2), decrease the π-type interaction between the E and Pn atoms compared to those of aminoboranes [2,3,4] and phosphanylboranes [5,6] due to the longer E–Pn σ-bond. These aspects can afford a characteristic reactivity reflecting the adjacent but separated Lewis acids and bases, similar to the synergetic interactions of the Lewis acid and Lewis base of frustrated Lewis pairs (FLPs) toward small-molecules [7,8,9,10].
The physical properties and reactivity of λ33-phosphanylalumanes, however, hardly have been clarified so far. The main reason is that there are few synthetic examples due to the difficulty of protecting the vacant 3p(Al) orbital [11,12,13]. Most of λ33-phosphanylalumanes have been synthesized by the salt elimination reactions reported by Power [14], Nöth and Paine [15,16], and our group [17] (Figure 1a). Multi-nuclear phosphanylalumane derivatives are also known. When there are reactive substituents (H or a halogen atom) on P or Al moiety, another aluminum or phosphorus reagents react to give a λ33-diphosphanyl-λ3-alumane [16] and λ3-phosphanyl-λ33-dialumane [17], respectively (Figure 1b). A cyclic λ33-phosphanylalumane (Mes*Al–PPh)3 is synthesized by a dehydrogenation reaction, which is thought to be the trimerization of an Al–P double-bond compound (Figure 1c) [18]. Regarding each case, at least one substituent on an Al or P atom is a heteroatom (non-carbon atom) substituent, which will affect the nature of the Al–P bond. The Lewis acidity of aluminum, for example, is greatly impaired by mesomeric effect between the vacant 3p(Al) orbital and the one pair on aluminum-bound substituents [19,20]. Electropositive substituents (SiPh3, SiMe3, and SnMe3) attached to the P atom also may alter the nature of the Al–P bond. Therefore, the synthesis and elucidation of the reactivity of unperturbed, all-carbon-substituted λ33-phosphanylalumanes might be a new challenge.
Here, we report the synthesis of λ33-phosphanylalumanes which are substituted fully by carbon protecting groups on the aluminum and phosphorus atoms. To understand the substituent effect on the aluminum, we chose t-Bu and C6F5 groups (Figure 1d).

2. Results and Discussion

The reaction of (t-Bu)2AlBr [21] with Mes2PLi [22] afforded a λ33-phosphanylalumane 1 quantitatively, judged by 31P NMR spectroscopy (Scheme 1a). Conversely, the use of (C6F5)2AlCl·0.5(toluene) [23] as an aluminum source in a hexane solution afforded an LiCl complex, λ44-phosphanylalumane 2·LiCl, in a 54% yield (Scheme 1b). We expected that the use of (C6F5)2AlBr [23] would render a more effective salt elimination but the corresponding LiBr complex (2·LiBr) was detected by NMR spectroscopy. No salt elimination in the case of the C6F5 substituent suggested the higher Lewis acidity of the aluminum atom on (C6F5)2Al-moiety as compared with that on the t-Bu2Al moiety. The attempts of LiCl elimination from 2·LiCl by heating or the addition of silver tetrafluoroborate were not successful. Although the reaction of (C6F5)2AlCl·0.5(toluene) in an Et2O solution, rather than hexane afforded a complicated mixture, recrystallization from which gave an etherate 2·Et2O in a 38% yield. The addition of Et2O to 2·LiCl led to the formation of compound 2·Et2O, which was evidenced in the mixture of (C6F5)2AlCl·0.5(toluene) with Mes2PLi in Et2O. These results suggest that 2·LiCl was formed at the initial stage in both conditions and the partial decomposition occurred via Et2O in addition to the formation of 2·Et2O.
Subsequently, the coordination of Lewis bases was examined to explore the Lewis acidity of λ33-phosphanylalumane 1. Although 1 did not react with carbon monoxide upon heating to 70 °C, the reaction of 1 with t-butyl isocyanide at room temperature gave the Lewis base-coordinated λ34-complex 1·(t-BuNC) quantitatively (Scheme 2). Heating of a C6D6 solution of 1·(t-BuNC) with the aim of further isomerization and coupling of isocyanides [24,25,26] produced a mixture containing a 1, 1·(t-BuNC), Mes2PH, and Mes2P–PMes2 as judged by the 31P NMR spectrum. The recrystallization from the mixture afforded a trace amount of single crystals of bis(t-butyl)aluminium cyanide tetramer 3 [27], as determined by NMR measurements and X-ray crystallography. The elimination of the t-Bu group from t-BuNC to form diorganylaluminum cyanides (R2AlCN)n also has been observed in the reaction of the Al–Al bond compound with t-BuNC [24]. Heating at elevated temperatures promotes desorption of the (t-Bu)2(t-BuNC)Al- and Mes2P-moieties to give 3, Mes2PH, and Mes2P–PMes2.
Structures of all new phosphanylalumane derivatives 1, 1·(t-BuNC), 2·LiCl, and 2·Et2O were determined finally by X-ray crystallography, and the results showed good agreement with the optimized structures calculated at the B3LYP-D3/6-31G(d) level. These structures are depicted in Figure 2 and the structural features of these compounds are summarized in Table 1. Particularly, 2·LiCl had a one-dimensional infinite chain structure with continuous [–Cl–Al–P–Li–] parts. One Li atom of 2·LiCl was coordinated by P and two F atoms of the C6F5 groups. Concerning complex 2·LiCl, its anion part (2Cl) without Li+, also was used for the calculations. There was almost no structural difference between the optimized structures, 2·LiCl and 2Cl. Lewis base-free λ33-phosphanylalumane 2 could not be obtained experimentally, but the optimized structural parameters of 2 also are described for the comparison.
The sum of the bond angles around the P atom (ΣP) of these compounds are 325–328°, indicating the pyramidalized structures of the phosphorus moieties despite the different coordination. Conversely, the sum of the bond angles around the Al atom (ΣAl) reflected the environment on the coordination of Al atoms. λ33-Phosphanylalumane 1 retained the planar structure with the Al center (ΣAl = 357°). The λ4-Al moieties of 1·(t-BuNC) and 2·Et2O were pyramidalized slightly (ΣAl = 354° and 347°, respectively), reflecting the weak coordination of isocyanide and diethyl ether to the λ3-Al atoms. To contrast, the λ4-Al center of 2·LiCl was pyramidalized extremely (ΣAl = 319°), suggesting its strong Al–Cl covalent bond.
The viewpoint of the Al–P bond lengths brought us the effect of electronic situations on the Al atom. The Al–P bond lengths of 1 (2.343(1) and 2.347(1) Å for two independent molecules) were comparable to that of the reported λ33-phosphanylalumanes (Tip2Al–P(SiPh3)Mes: 2.342(2) Å) [14] or the Al3P3 six-membered-ring compound ((Mes*Al–PPh)3: 2.323(3)–2.336(3) Å) [18]. The coordination of t-BuNC to the λ3-Al center resulted in the large elongation of the Al–P bond (1·(t-BuNC): 2.4120(6) Å). Alternately, the Al–P bond lengths of C6F5-substituted λ4-Al-derivatives, 2·LiCl (2.348(1) Å) and 2·Et2O (2.359(2) Å), were close to those of λ33-phosphanylalumane 1 and much shorter than those of 1·(t-BuNC) and the reported phosphanyl-λ4-alumanes, ((t-Bu)2Al–P(SiPh3)Tip·Et2O: 2.416(3) Å) [12] and (Bbp(Br)Al–P(H)Mes·LiBr(Et2O)2: 2.4055(7) Å) [13]. Considering the observations above, the C6F5-substituted λ3-Al-derivative 2 is expected to have a quite short Al–P bond. However, the calculated Al–P bond length of 2 almost was similar to or slightly shorter than that of 1 (1: 2.348 Å vs. 2: 2.322 Å). These results indicated that the C6F5 groups greatly affected the Al–P bond regardless of the coordination number to the Al center, the reasons of which will be discussed later.
The observed structural features reflect the NMR spectroscopic results. The 31P NMR spectra for λ33-phosphanylalumane 1 displayed a resonance at δ = −111.7 ppm that is the most upfield among those of the synthesized phosphanylalumane derivatives (1·(t-BuNC): −96.7 ppm, 2·Et2O: −104.8 ppm, and 2·LiCl: −89.7 ppm). Additionally, these 31P NMR signals are in much lower magnetic fields than those of tmp2Al–P(SiMe3)2 and tmp2Al–P(SnMe3)2 (δ = −238 and −282 ppm, respectively) [16] having an electron-donating substituent on λ3-P atoms, while being more upfield than that of tmp2Al–PPh2 (δ = −42.9 ppm) [15]. These values indicate a sufficiently electron-rich environment for the λ3-P atom with carbon protecting groups in 1. Conversely, the 27Al NMR spectra for 1 showed a single broad resonance at δ = 261 ppm. This value is upfield as compared with those of the reported λ33-phosphanylalumanes, i.e., tmp2Al–PPh2 (δ = 110 ppm) [15] and tmp2Al–P(SiMe3)2 (δ = 78 ppm) [16].
The photophysical behavior of λ33-phosphanylalumane 1 has been investigated for the first time to obtain further insight for the character of a λ3-Al moiety. UV–vis absorption spectrum of 1 is shown in Figure 3 with the simulated spectrum of 1 derived from time-dependent density functional theory (TD-DFT) calculations at the B3LYP-D3/6-311+G(d,p)//B3LYP-D3/6-31G(d) level of theory, which reproduced the spectrum observed experimentally. The shoulder absorption at ~350 nm should be assigned to the n-p transition from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO), which corresponds to the electron transition of a lone pair of P to a vacant p orbital of Al (Figure 3a; (A)). The absorption maximum at 314 nm (ε 5.2 × 103) was assigned to the n-π* transition from the HOMO to the LUMO+1, which corresponds to the electron transition of a lone pair of P to π* orbitals of the Mes2P moiety (Figure 3a; (B)). Combining the experimental and theoretical results, the vacant p orbital of λ33-phosphanylalumane 1 greatly contributes to the absorption behavior. Particularly, the λ33-phosphanylalumanes with the unperturbed vacant p orbital of Al exhibit a yellow color (1 and Tip2Al–P(SiPh3)Mes (no UV–vis absorption spectrum) [14]), whereas the λ33-phosphanylalumanes with the filled p orbital of Al by the Al-bound substituents are colorless (tmp2Al–P(SiMe3)2 [16] and Mes*(Cl)Al–P(H)Mes* [17]).
The consideration of the Al–P bonds for the newly synthesized phosphanylalumane derivatives was further deepened by the natural bond orbital (NBO) analysis, as summarized in Table 2. The π-type NBOs involving the p(Al) orbitals were not found in 1 and 2. The natural population analysis (NPA) charge on Al of 1 (1.760) is the most positive among these phosphanylalumane derivatives, indicating the substantial retention of the vacant 3p(Al) orbital. The NBO corresponding to the σ(Al–P) bond in 1 consisted mainly of 3s(Al) and 3p(P) orbitals and was polarized toward an λ3-P moiety. The coordination of t-BuNC to λ3-Al increases the contribution of the hybridization of Al in σ(Al–P) bonds (1: 18% Al vs. 1·(t-BuNC): 21% Al) and decreases the s-character of Al (1: Al sp1.7 vs. 1·(t-BuNC): Al sp2.6), as well as the experimentally observed elongation of an Al–P bond. Substitution of t-Bu groups with C6F5 groups resulted in the higher s-character of the Al atom in 2 than that of 1. These are due to the use of more p-character in the bonding to the aryl group than that to the alkyl group and of the large inductive effect of the F atoms. This result was supported by the large Al contribution of 2, 2Cl, and 2·LiCl compared to 1 in σ(Al–P) bonds (1: 18% Al vs. 2, 2Cl, and 2·LiCl: 23–26% Al) and the large s-character of Al (1: Al sp1.7 vs. 2, 2Cl, and 2·LiCl: Al sp0.6−0.9), as well as the short Al–P bond lengths independent of the coordination-number change in the Al moieties of C6F5-derivatives. Additionally, compound 2 showed a WBI value greater than that of 1 (1: 0.6595 vs. 2: 0.7781) corresponding to the short Al–P bonds of 2·LiCl and 2·Et2O, but the NPA charge of the λ3-Al moiety of 2 was less than that of 1 (1: 1.760 vs. 2: 1.537). These results of the theoretical calculations suggest that compound 1 is a suitable λ33-phosphanylalumane for reactivity studies.
The singularity of the Al–P single-bond in 1 among the bonds between the group 13 and group 15 elements also was experimentally investigated. Treatment of 1 with benzophenone at room temperature rapidly gave Mes2P–PMes2 quantitatively, as judged by the 31P NMR spectrum (Scheme 3a), but the corresponding Al moiety was not identified. This result implied that the coordination of benzophenone to Al promoted homolysis to produce an Mes2P radical and the corresponding radical containing an Al moiety, respectively. It was difficult to clear the radical mechanism by using (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO) because compound 1 reacted with TEMPO to give a mixture containing Mes2P–PMes2 and Mes2PH. Conversely, a B–P bond compound with λ3-B and P centers, phosphaboradibenzofulvene, gave a [2+2]-cycloaddition product when treated with benzophenone (Scheme 3b) [28]. Contrary to the large contribution of the B–P π-bond character in a phosphaboradibenzofulvene, these results suggested the Al–P single-bond character and the clear separation of the Lewis acid and Lewis base moieties due to the slight interaction between Al and P in 1.

3. Materials and Methods

3.1. General

All the manipulations were performed under a dry argon atmosphere using standard Schlenk techniques or gloveboxes. Solvents were purified by the Ultimate Solvent System, Glass Contour Company (Nikko Hansen and Co., Ltd., Osaka, Japan) (THF, toluene, and n-hexane) [29] or by trap-to-trap distillation from a potassium mirror prior to use (C6D6 and n-hexane for UV–vis spectrum measurements). NMR spectra were measured on a JMM-ECA600 (JEOL Ltd., Tokyo, Japan) (1H: 600 MHz, 7Li: 233 Hz, 13C: 151 MHz, 19F: 565 MHz, 27Al: 156 MHz, 31P: 243 MHz) in the Joint Usage/Research Center (JURC, Institute for Chemical Research, Kyoto University) or on a AL-300 spectrometer (JEOL Ltd., Tokyo, Japan) (1H: 300 MHz, 13C: 75 MHz, 19F: 282 MHz, 31P: 121 MHz). Regarding the 1H NMR spectra, signals arising from residual partially hydrogenated C6D5H (7.15 ppm for 1H) were used as references. C6D6 (128.0 ppm for 13C) and Al(NO3)3 in D2O (0 ppm for 27Al) were used as references. 1H and 13C NMR signals were assigned with the aid of the 1H–1H COSY, 1H–13C HSQC, and 1H–13C HMBC spectra. Melting points were determined on a Yanaco micro melting point apparatus and uncorrected. Elemental analyses were carried out at the Microanalytical Laboratory, Institute for Chemical Research, Kyoto University.
AlCl3 and AlBr3 (Sigma-Aldrich Co., LLC., Tokyo, Japan), to prepare aluminum reagents, were purified by sublimation prior to use. The aluminum monohalides ((t-Bu)2AlBr [21], and (C6F5)2AlX (X = Cl, Br) or (C6F5)2AlCl·0.5(toluene) [23]) were prepared in accordance with the reported procedures. Mes2PLi was prepared by the reaction of Mes2PH with n-butyllithium in n-hexane, and the yellow solid was washed with n-hexane carefully to remove the residual Mes2PH [22]. Details of theoretical calculations and XRD data are given in the Supplementary Materials.

3.2. Synthesis of 1

To a yellow suspension of Mes2PLi (277.2 mg, 1.00 mmol) in n-hexane (3 mL), (t-Bu)2AlBr (219.5 mg, 0.998 mmol, 1.0 eq.) was slowly added at room temperature. The reaction mixture was stirred for 12 h. The almost quantitative formation of λ33-phosphanylalumane was observed, as judged by 31P NMR spectroscopy. The insoluble materials were removed by filtration through a Celite® pad using n-hexane as an eluent. The filtrate was concentrated and stored at −35 °C, presenting λ33-phosphanylalumane (1) as yellow crystals (isolated yield: 214.0 mg, 0.521 mmol, 52%). 1: Yellow crystals, mp. 54–55 °C. 1H NMR (600 MHz, C6D6, 298 K): δ = 1.09 (d, 18H, 4JHP = 1.2 Hz, C(CH3)3), 2.09 (s, 6H, Mes p-CH3), 2.36 (s, 12H, Mes o-CH3), 6.74 (m, 4H, Mes m-CH) ppm. 13C{1H} NMR (151 MHz, C6D6, 298 K): δ = 20.9 (Mes p-CH3), 24.5 (d, 3JCP = 13.6 Hz, Mes o-CH3), 29.7 (d, 3JCP = 1.7 Hz, Al-C(CH3)3), 30.53 (s, AlCMe3), 129.62 (d, 3JCP = 3.0 Hz, Mes m-C), 133.79 (d, 2JCP = 16.6 Hz, Mes o-C), 136.13 (s, Mes p-C), 141.19 (d, 1JCP = 12.1 Hz, Mes ipso-C) ppm. 27Al NMR (156 MHz, C6D6, 298 K): δ = 261.3 (br.) ppm. 31P{1H} NMR (243 MHz, C6D6, 298 K): δ = −111.7 ppm. UV–vis (hexane): λ/nm = 314 (ε 5.2 × 103), 349 (ε 1.5 × 103). Anal. Calcd. for C26H40AlP: C, 76.06; H, 9.82. Found: C, 76.07; H, 10.10.

3.3. Synthesis of 1·(t-BuNC)

To a yellow solution of 1 (40.3 mg, 0.0982 mmol) in C6D6 (0.6 mL), a small excess amount of t-BuNC (9.6 mg, 0.115 mmol, 1.2 eq.) was slowly added at room temperature. The quantitative formation of 1·(t-BuNC) was observed, as judged by 31P NMR spectroscopy. The mixture was concentrated at room temperature presenting the orange crystals. The crystals were washed with hexane to afford 1·(t-BuNC) as light orange crystals (isolated yield: 16.4 mg, 0.0332 mmol, 34%). 1·(t-BuNC): Light orange crystals, mp. 97–99 °C. 1H NMR (600 MHz, C6D6, 298 K): δ = 0.64 (s, 9H, C≡N–C(CH3)3), 1.35 (d, 4JHP = 0.6 Hz, 18H, AlC(CH3)3), 2.14 (s, 6H, Mes p-CH3), 2.55 (s, 12H, Mes o-CH3), 6.80 (m, 4H, Mes m-H) ppm. 13C{1H} NMR (151 MHz, C6D6, 298 K): δ = 18.01 (d, 2JCP = 12.1 Hz, AlCMe3), 20.95 (s, Mes p-CH3), 25.47 (d, 3JCP = 10.6 Hz, Mes o-CH3), 28.40 (s, C≡N–C(CH3)3), 32.56 (d, 3JCP = 2.4 Hz, AlC(CH3)3), 57.72 (s, C≡NCMe3), 129.04 (d, 3JCP = 4.5 Hz, Mes m-C), 132.56 (d, 2JCP = 33.2 Hz, Al←C≡N-t-Bu), 134.66 (s, Mes p-C), 137.19 (d, 2JCP = 18.1 Hz, Mes o-C), 142.19 (d, 1JCP = 10.6 Hz, Mes ipso-C) ppm. 27Al NMR (156 MHz, C6D6, 298 K): δ = +154.7 (br.) ppm. 31P{1H} NMR (121 MHz, C6D6, 298 K): δ = −154 ppm. Anal. Calcd for C31H49AlNP: C, 75.42; H, 10.00; N, 2.84. Found: C, 75.02; H, 10.10; N, 2.83.

3.4. Synthesis of 2·LiCl

To a yellow suspension of Mes2PLi (25.4 mg, 0.0939 mmol) in n-hexane (2 mL), (C6F5)2AlCl·0.5(toluene) (47.0 mg, 0.0942 mmol) was added at room temperature. The reaction mixture was stirred until the yellow color diminished (for 12 h). Following removal of the solvent, the residue was washed with the n-hexane. The insoluble materials were filtered through a Celite® pad using toluene as an eluent. n-Hexane was added to the filtrate and the mixture was scratched to give the colorless precipitate. The precipitate was washed with n-hexane carefully to afford (C6F5)2Al–PMes2·LiCl complex (2·LiCl) as a colorless solid (34.3 mg, 0.0510 mmol, 54%). 2·LiCl: Colorless solid, mp. 94 °C (dec.). 1H NMR (600 MHz, C6D6, 298 K): δ = 1.96 (s, 6H, Mes p-CH3), 2.23 (s, 12H, Mes o-CH3), 6.58 (d, 4JHP = 2.4 Hz, 4H, Mes m-H) ppm. 7Li{1H} NMR (233 MHz, C6D6, 298K): δ = −4.73 ppm. 13C{1H} NMR (151 MHz, C6D6, 298K): δ = 20.70 (s, Mes p-CH3), 24.75 (d, 3JCP = 9.1 Hz, Mes o-CH3), 117.05 (br. t, 2JCF = ~49 Hz, C6F5 ipso-C), 133.21 (d, 2JCF = 9.1 Hz, Mes o-C), 137.09 (s, Mes p-C), 137.26 (dqd, 1JCF = 252 Hz, 2JCF = ~12 Hz, 3JCF = ~4.5 Hz, C6F5 m-C), 141.53 (dm, 1JCF = 252 Hz, C6F5 p-C), 143.13 (d, 1JCP = 10.6 Hz, Mes ipso-C), 149.94 (ddm, 1JFC = 224 Hz, 2JFC = 24 Hz, C6F5 o-C) ppm. 19F NMR (282 MHz, C6D6, 298 K): δ = −123.3 (m, 4F, C6F5 o-F), –155.4 (t, J = 19.7 Hz, 2F, C6F5 p-F), −161.2 (m, 4F, C6F5 m-F) ppm. 31P{1H} NMR (121 MHz, C6D6, 298 K): δ = −89.7 (br. s) ppm. No 27Al NMR signal was observed, even after long-time measurement. Due to the flame retardancy of fluorocarbons and the extremely high air-/moisture-sensitivity, satisfactory data of the elemental analysis could not be obtained. The purity of 2·LiCl was confirmed accordingly by the 19F and 31P{1H} NMR spectra.

3.5. Synthesis of 2·LiBr

To a yellow suspension of Mes2PLi (11.9 mg, 0.044 mmol) in n-hexane (2 mL), (C6F5)2AlBr (23.3 mg, 0.0441 mmol) was added at room temperature. The solution was stirred until the yellow color diminished (for 5 h). The precipitate was washed with the n-hexane, and the residue was filtered through a Celite® pad using toluene as an eluent. Following removal of the solvent from the filtrate, the resulting materials (14.5 mg) was checked by NMR spectroscopy, suggesting the formation of (C6F5)2Al–PMes2·LiBr complex (2·LiBr). Further purification and isolation did not succeed. 2·LiBr: A colorless solid of crude materials. 1H NMR (300 MHz, C6D6, 298 K): δ = 1.96 (s, 6H, Mes p-CH3), 2.22 (s, 12H, Mes o-CH3), 6.59 (m, 4H, Mes m-H) ppm. 7Li{1H} NMR (117 MHz, C6D6, 298 K): δ = −4.38 ppm. 19F NMR (282 MHz, C6D6, 298 K): δ = −122.9 (m, 4F, C6F5 o-F), −152.9 (t, J = 19.7 Hz, 2F, C6F5 p-F), −160.8 (m, 4F, C6F5 m-F) ppm. 31P{1H} NMR (121 MHz, C6D6, r.t.): δ = −89.3 (s) ppm.

3.6. Synthesis of 2·Et2O

To a yellow suspension of Mes2PLi (24.2 mg, 0.0485 mmol) in Et2O (2 mL), (C6F5)2AlCl·0.5(toluene) (13.3 mg, 0.0492 mmol) was added at room temperature. Following removal of the solvent immediately, the residue was washed with n-hexane. The insoluble materials were separated by filtration through a Celite® pad using toluene as an eluent. Subsequent to removal of the solvent from the filtrate, the residue was washed several times with n-hexane to afford (C6F5)2Al–PMes2·Et2O (2·Et2O) as a colorless solid (14.3 mg, 0.0203 mmol, 38%). Colorless crystals, mp. 129 °C (dec.). 1H NMR (300 MHz, C6D6, 298 K): δ = 0.35 (t, 6H, OCH2CH3), 2.09 (s, 6H, Mes p-CH3), 2.35 (s, 12H, Mes o-CH3), 3.76 (q, 4H, OCH2CH3), 6.74 (m, 4H, Mes m-H). 19F NMR (282 MHz, C6D6, 298 K): δ = −119.4 (m, 4F, C6F5 o-F), −152.5 (t, J = 19.7 Hz, 2F, C6F5 p-F), −161.1 (m, 4F, C6F5 m-F) ppm. 31P{1H} NMR (121 MHz, C6D6, r.t.): δ = −104.8 ppm. No 27Al NMR signal was observed, even after long-time measurement. Satisfactory data of the 13C NMR spectrum could not be obtained due to the impurities and decomposition. Due to the flame retardancy of fluorocarbons and the extremely high air-/moisture-sensitivity, satisfactory data of the elemental analysis could not be obtained.

3.7. Reaction of 2·LiCl with Et2O

The solid of 2·LiCl (1.2 mg, 0.0018 mmol) was slowly dissolved Et2O (1 mL) at room temperature. The solvent was removed immediately and the mixture was analyzed by 1H and 19F NMR spectroscopy, presenting a mixture the same as the observation of the reaction for the (C6F5)2AlCl with Mes2PLi in Et2O solution. This result indicated that 2·LiCl is formed at the initial stage of the reaction and Et2O is coordinated toward 2·LiCl or promotes decomposition of 2·LiCl.

3.8. Thermal Isomerization of 1·(t-BuNC)

Using a J.Young NMR tube, a solution of 1·(t-BuNC) (5.6 mg, 0.011 mmol) in C6D6 (0.6 mL) was degassed by freeze-pump-thaw cycles and heated at 100 °C over 48 h. The signals of the corresponding 1, 1·(t-BuNC), Mes2PH, Mes2P–PMes2 were observed by 31P{1H} NMR spectroscopy. The recrystallization from a mixture afforded a trace amount of 3 as red crystals.

3.9. Reaction of 1 with Benzophenone

Using a J.Young NMR tube, a solution of 1 (12.7 mg, 0.0309 mmol) in C6D6 (0.6 mL) was treated with an excess amount of benzophenone (9.2 mg, 0.0505 mmol, 1.6 eq.) at room temperature. The quantitative formation of Mes2P–PMes2 was observed, as judged by 31P{1H} NMR spectroscopy.

3.10. X-Ray Crystallographic Analysis

The intensity data were collected on a Saturn 70 CCD diffractometer (Rigaku Corp., Tokyo, Japan) with a VariMax Mo optic system using Mo Kα radiation (λ = 0.71075 Å) [for 1, 1·(t-BuNC), and 2·LiCl], or a Mercury CCD diffractometer (Rigaku Corp., Tokyo, Japan) with graphite monochromated Mo Kα radiation (λ = 0.71069 Å) (for 2·Et2O). An empirical absorption correction was applied to the diffraction data using ABSPACK [30] for 1, 1·(t-BuNC), 2·LiCl, and 2·Et2O. The structure was solved by a direct method (SHELXT [31]) and refined by a full-matrix least-squares method on F2 for all reflections (SHELXL-2016/4 [32]). All hydrogen atoms were placed using AFIX instructions, while all other atoms were refined anisotropically. CCDC-1959322 (1), CCDC-1959323 [1·(t-BuNC)], CCDC-1959324 (2·Et2O), and CCDC-1959325 (2·LiCl) 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.

3.11. Theoretical Calculations

DFT calculations were performed using the Gaussian 16 (Rev. B. 01) [33] program package (Gaussian, Inc., Wallingford, CT, USA) with B3LYP functional [34,35,36] including Grimme dispersion correction (D3) [37,38] along with a combined basis set: 6-31G(d) level. All the geometry optimizations were performed until the residual mean force was smaller than 1.0 × 10−5 a.u. (tight threshold in Gaussian). The frequency calculations were carried out for each optimized structure to confirm the absence of any imaginary frequencies. The vertical excitation energy and electronic absorption spectra were simulated using time-dependent density functional theory (TD-DFT) [39] at the B3LYP-D3/6-311G+(d,p)//B3LYP-D3/6-31G(d) level. Natural bond orbital (NBO) analyses were conducted with the NBO 6.0 program package [40], linked to single-point calculations using Gaussian 16 (Rev. B. 01)

4. Conclusions

To summarize, we reported here the synthesis and structures of novel phosphanylalumane derivatives whose protecting groups on the Al and P atoms are all carbon substituents. Substituent effects on an Al atom were investigated, and the introduction of C6F5 groups on the Al atom substantially increased the Lewis acidity of aluminum. Additionally, they presented extremely short Al–P bond lengths compared to those of related compounds. Based on the results of X-ray crystallographic analysis, theoretical calculations, and the reaction with benzophenone, λ33-phosphanylalumane 1 was found to have a well-separated vacant p orbital on an Al atom and lone pairs on a P atom. Its spectroscopic properties were investigated for the first time, presenting a yellow color due to the charge transfer character between Al and P atoms in λ33-phosphanylalumane. Further studies on the reactivity of phosphanylalumanes are now under way.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/7/11/132/s1. NMR spectra. Crystallographic data (CCDC 1959322–1959325) and theoretical calculations details. Supplementary material includes the CIF and CheckCIF files for 1, 1·(t-BuNC), 2·LiCl, and 2·Et2O.

Author Contributions

T.Y. designed the project and performed the experiments and measurements. T.Y. carried out the X-ray crystallographic analysis and theoretical calculations. T.Y. and Y.M. designed the experiments. T.Y. and Y.M. co-wrote the paper. T.Y., Y.M., and N.T. reviewed and approved the final manuscript. All authors contributed to the discussions.

Funding

This research was partially funded by JSPS KAKENHI (JP24109013 and JP26620028), JSPS Research Fellowship for Young Scientists (JP19J14359), and Integrated Research Consortium on Chemical Science (IRCCS).

Acknowledgments

This study was supported by the Joint Usage/Research Center [JURC, Institute for Chemical Research (ICR), Kyoto University] by providing access to a JEOL JNM-ECA600 NMR spectrometer. The authors are furthermore grateful for computation time, which was provided by the Super Computer System (ICR, Kyoto University). Elemental analyses were carried out at the Microanalytical Laboratory of the ICR (Kyoto University).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Niedenzu, K.; Dawson, J.W. Aminoboranes. In Boron-Nitrogen Compounds; Niedenzu, K., Dawson, J.W., Eds.; Springer: Berlin/Heidelberg, Germany, 1965; pp. 48–84. [Google Scholar]
  2. Bullen, G.J.; Clark, N.H. Crystallographic studies of the boron–nitrogen bond in aminoboranes. Part II. Crystal structure of (dimethylamino) dimethylborane at −95°C. J. Chem. Soc. A 1970, 992–996. [Google Scholar] [CrossRef]
  3. Brauer, D.J.; Bier, H.; Diirrenbach, F.; Pawelke, G.; Weuter, W. Synthesis of diethyl- and diisopropylamino-trifluoromethylboranes and their reactions with HF, HCl, HBr and H2O. The crystal structure of (CF3)2BN(i-Pr)2, revealing a very short B–N bond. J. Organomet. Chem. 1989, 378, 125–137. [Google Scholar] [CrossRef]
  4. Faderl, J.; Deobald, B.; Guilard, R.; Pritzkow, H.; Siebert, W. Syntheses, structures, and reactivity of 2,5-diboryl-1-alkylpyrroles and di(1-alkyl-2-pyrrolyl)boranes. Eur. J. Inorg. Chem. 1999, 1999, 399–404. [Google Scholar] [CrossRef]
  5. Paine, R.T.; Nöth, H. Recent advances in phosphinoborane chemistry. Chem. Rev. 1995, 95, 343–349. [Google Scholar] [CrossRef]
  6. Bailey, J.A.; Pringle, P.G. Monomeric phosphinoboranes. Coord. Chem. Rev. 2015, 297–298, 77–90. [Google Scholar] [CrossRef]
  7. Stephan, D.W. Frustrated Lewis Pairs. J. Am. Chem. Soc. 2015, 137, 10018–10032. [Google Scholar] [CrossRef]
  8. Stephan, D.W.; Erker, G. Frustrated Lewis Pair Chemistry: Development and Perspectives. Angew. Chem. Int. Ed. 2015, 54, 6400–6441. [Google Scholar]
  9. Stephan, D.W. The broadening reach of frustrated Lewis pair chemistry. Science 2016, 354, aaf7229. [Google Scholar] [CrossRef]
  10. Jupp, A.R.; Stephan, D.W. New Directions for Frustrated Lewis Pair Chemistry. Trends Chem. 2019, 1, 35–48. [Google Scholar] [CrossRef]
  11. Atwood, D.A.; Contreras, L.; Cowley, A.H.; Jones, R.A.; Mardones, M.A. Monomeric base-stabilized phosphino- and arsinoalanes. Organometallics 1993, 12, 17–18. [Google Scholar] [CrossRef]
  12. Petrie, M.A.; Power, P.P. Synthesis and characterization of the monomeric phosphinogallanes But2GaPR’R’’ (R’, R’’ = bulky aryl or silyl groups) and related compounds. J. Chem. Soc. Dalton Trans. 1993, 1737–1745. [Google Scholar] [CrossRef]
  13. Agou, T.; Ikeda, S.; Sasamori, T.; Tokitoh, N. Synthesis and structure of Lewis base-coordinated phosphanylalumanes bearing P–H and Al–Br moieties. Eur. J. Inorg. Chem. 2018, 2018, 1984–1987. [Google Scholar] [CrossRef]
  14. Wehmschulte, R.J.; Ruhlandt-Senge, K.; Power, P.P. Synthesis and characterization of unassociated aluminum monophosphides. Inorg. Chem. 1994, 33, 3205–3207. [Google Scholar] [CrossRef]
  15. Knabel, K.; Krossing, I.; Nöth, H.; Schwenk-Kircher, H.; Schmidt-Amelunxen, M.; Seifert, T. The aluminum–nitrogen bond in monomeric bis(amino)alanes: A systematic experimental study of bis(tetramethylpiperidino)alanes and quantum mechanical calculations on the model system (H2N)2AlY. Eur. J. Inorg. Chem. 1998, 1998, 1095–1114. [Google Scholar] [CrossRef]
  16. Habereder, T.; Nöth, H.; Paine, R.T. Synthesis and reactivity of new bis(tetramethylpiperidino)(phosphanyl)alumanes. Eur. J. Inorg. Chem. 2007, 2007, 4298–4305. [Google Scholar] [CrossRef]
  17. Agou, T.; Ikeda, S.; Sasamori, T.; Tokitoh, N. Synthesis and structure of Lewis-base-free phosphinoalumane derivatives. Eur. J. Inorg. Chem. 2016, 2016, 623–627. [Google Scholar] [CrossRef]
  18. Wehmschulte, R.J.; Power, P.P. Reactions of (H2AlMes*)2 (Mes* = 2,4,6-(t-Bu)3C6H2) with H2EAr (E = N, P, or As; Ar = aryl):  Characterization of the ring compounds (Mes*AlNPh)2 and (Mes*AlEPh)3 (E = P or As). J. Am. Chem. Soc. 1996, 118, 791–797. [Google Scholar] [CrossRef]
  19. Agou, T.; Yanagisawa, T.; Sasamori, T.; Tokitoh, N. Synthesis and structure of an iron-bromoalumanyl complex with a tri-coordinated aluminum center. Bull. Chem. Soc. Jpn. 2016, 89, 1184–1186. [Google Scholar] [CrossRef]
  20. Yanagisawa, T.; Mizuhata, Y.; Tokitoh, N. Dibromometallyl-iron complexes generated by the recombination of an alumanyl-iron complex with EBr3 (E = Al, Ga). Heteroatom Chem. 2018, 29, e21465. [Google Scholar] [CrossRef]
  21. Uhl, W.; Schnepf, J.E.O.Z. Synthesis and molecular structure of (N,N’-dimethylpiperazine)lithium-µ-hydrido)(tert-butyl)bis[bis(trimethylsilyl)methyl]alanate with an intramolecular interaction between lithium and C–H-σ-bonds. Z. Anorg. Allg. Chem. 1991, 595, 225–238. [Google Scholar] [CrossRef]
  22. Geier, S.J.; Gilbert, T.M.; Stephan, D.W. Synthesis and reactivity of the phosphinoboranes R2PB(C6F5)2. Inorg. Chem. 2011, 50, 336–344. [Google Scholar] [CrossRef]
  23. Styra, S.; Radius, M.; Moos, E.; Bihlmeier, A.; Breher, F. Aluminium diphosphamethanides: Hidden frustrated Lewis pairs. Chem. Eur. J. 2016, 22, 9508–9512. [Google Scholar] [CrossRef]
  24. Uhl, W.; Schütz, U.; Hillerb, W.; Heckell, M. Reaction of isothiocyanates and isonitriles with R2Al–AlR2 [R = CH(SiMe3)2]; single and double insertion of isonitrile into the Al–Al bond. Chem. Ber. 1994, 127, 1587–1592. [Google Scholar] [CrossRef]
  25. Nagata, K.; Agou, T.; Sasamori, T.; Tokitoh, N. Formation of a diaminoalkyne derivative by dialumane-mediated homocoupling of t-butyl isocyanide. Chem. Lett. 2015, 44, 1610–1612. [Google Scholar] [CrossRef]
  26. Chen, W.; Zhao, Y.; Xu, W.; Su, J.-H.; Shen, L.; Liu, L.; Wu, B.; Yang, X.-J. Reductive linear- and cyclo-trimerization of isocyanides using an Al–Al-bonded compound. Chem. Commun. 2019, 55, 9452–9455. [Google Scholar] [CrossRef]
  27. Knabel, K.; Nöth, H. Synthesis and structures of some aluminum pseudohalides. Z. Naturforsch. 2005, 60b, 155–163. [Google Scholar] [CrossRef]
  28. Breunig, J.M.; Hübner, A.; Bolte, M.; Wagner, M.; Lerner, H.-W. Reactivity of phosphaboradibenzofulvene toward hydrogen, acetonitrile, benzophenone, and 2,3-dimethylbutadiene. Organometallics 2013, 32, 6792–6799. [Google Scholar] [CrossRef]
  29. Pangborn, A.B.; Giardello, M.A.; Grubbs, R.H.; Rosen, R.K. Silicon carbon unsaturated compounds. 21. Isomerization of a 1-silapropadiene in the presence of tetrakis(triethylphosphine)nickel(0). Organometallics 1996, 5, 1518–1520. [Google Scholar] [CrossRef]
  30. CrysAlis, C.C.D. CrysAlis RED including ABSPACK, Versions 1.171.39.46; Oxford Diffraction Ltd.: Abingdon, UK, 2006. [Google Scholar]
  31. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, 71, 3–8. [Google Scholar] [CrossRef]
  32. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  33. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Li, X.; et al. Gaussian 16, Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  34. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5662. [Google Scholar] [CrossRef]
  35. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef] [PubMed]
  36. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104–154119. [Google Scholar] [CrossRef]
  38. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  39. Runge, E.; Gross, E.K.U. Density-functional theory for time-dependent systems. Phys. Rev. Lett. 1984, 52, 997–1000. [Google Scholar] [CrossRef]
  40. Glendening, E.D.; Badenhoop, J.K.; Reed, A.E.; Carpenter, J.E.; Bohmann, J.A.; Morales, C.M.; Landis, C.R.; Weinhold, F. NBO 6.0; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, USA, 2013. [Google Scholar]
Figure 1. Known phosphanylalumanes. (a) Synthetic strategy of λ33-phosphanylalumanes. (b) λ3-Phosphanyl-λ33-dialumane and λ33-diphosphanyl-λ3-alumane. (c) Cyclic λ33-phosphanylalumane. (d) Phosphanylalumane derivatives with carbon protecting groups on Al and P atoms described in this study.
Figure 1. Known phosphanylalumanes. (a) Synthetic strategy of λ33-phosphanylalumanes. (b) λ3-Phosphanyl-λ33-dialumane and λ33-diphosphanyl-λ3-alumane. (c) Cyclic λ33-phosphanylalumane. (d) Phosphanylalumane derivatives with carbon protecting groups on Al and P atoms described in this study.
Inorganics 07 00132 g001
Scheme 1. Synthesis of phosphanylalumane derivatives bearing (a) t-Bu and (b) C6F5 groups on Al atoms.
Scheme 1. Synthesis of phosphanylalumane derivatives bearing (a) t-Bu and (b) C6F5 groups on Al atoms.
Inorganics 07 00132 sch001
Scheme 2. Lewis base (t-BuNC) coordination to 1 and thermal reaction of 1·(t-BuNC).
Scheme 2. Lewis base (t-BuNC) coordination to 1 and thermal reaction of 1·(t-BuNC).
Inorganics 07 00132 sch002
Figure 2. Structures of (a) 1, (b) 1·(t-BuNC), (c) 2·Et2O, and (d) 2·LiCl (thermal ellipsoids at the 50% probability level). One of the two crystallographically independent molecules of 1 is shown here. Hydrogen atoms and solvent molecules are omitted for clarity. Color code: Al, magenta; C, gray; Cl, green; F, yellow; Li, purple; N, blue; O, red; P, orange.
Figure 2. Structures of (a) 1, (b) 1·(t-BuNC), (c) 2·Et2O, and (d) 2·LiCl (thermal ellipsoids at the 50% probability level). One of the two crystallographically independent molecules of 1 is shown here. Hydrogen atoms and solvent molecules are omitted for clarity. Color code: Al, magenta; C, gray; Cl, green; F, yellow; Li, purple; N, blue; O, red; P, orange.
Inorganics 07 00132 g002
Figure 3. (a) Observed (in hexane at 298 K) and simulated (derived from TD-DFT calculations at the B3LYP-D3/6-311+G(d,p)//B3LYP-D3/6-31G(d) level of theory) UV–Vis spectrum for 1, whereby the calculated oscillator strengths are shown as blue vertical lines. (b) Electronic transitions corresponding to the absorptions (A) and (B). Hydrogen atoms are omitted for clarity. Color code: Al, pink; C, light gray; P, orange.
Figure 3. (a) Observed (in hexane at 298 K) and simulated (derived from TD-DFT calculations at the B3LYP-D3/6-311+G(d,p)//B3LYP-D3/6-31G(d) level of theory) UV–Vis spectrum for 1, whereby the calculated oscillator strengths are shown as blue vertical lines. (b) Electronic transitions corresponding to the absorptions (A) and (B). Hydrogen atoms are omitted for clarity. Color code: Al, pink; C, light gray; P, orange.
Inorganics 07 00132 g003
Scheme 3. (a) Reaction of the λ33-phosphanylalumane 1 with benzophenone. (b) Reaction of the phosphaboradibenzofulvene with benzophenone, reported by Lerner et al.
Scheme 3. (a) Reaction of the λ33-phosphanylalumane 1 with benzophenone. (b) Reaction of the phosphaboradibenzofulvene with benzophenone, reported by Lerner et al.
Inorganics 07 00132 sch003
Table 1. Structural parameters of 1, 1·(t-BuNC), 2, 2·LiCl, 2Cl, and 2·Et2O.
Table 1. Structural parameters of 1, 1·(t-BuNC), 2, 2·LiCl, 2Cl, and 2·Et2O.
CompoundAl–P/ÅΣAl/degΣP/degδ(31P)/ppm
1 (obsd.) 12.343(1)/2.347(1)357/356325/323−111.7
1 (calcd.) 22.348357325-
1·(t-BuNC) (obsd.)2.4120(6)354328−96.7
1·(t-BuNC) (calcd.) 22.435355330-
2 (calcd.) 22.322357312-
2·LiCl (obsd.)2.348(1)319327−89.7
2·LiCl (calcd.) 22.434312327-
2Cl (calcd.) 22.411321321-
2·Et2O (obsd.)2.359(2)347326−104.8
2·Et2O (calcd.) 22.364352320-
1 Two independent molecules. 2 Calculated at the B3LYP-D3/6-31G(d) level.
Table 2. Summary of NBO analysis.
Table 2. Summary of NBO analysis.
CompoundsWBI 1QAl2QP2NBO 3 σ(Al–P)
10.65951.7600.0709718.2% Al sp1.71 + 81.8% P sp4.19 (1.87e)
1·(t-BuNC)0.68591.5370.144521.0% Al sp2.57 + 79.0% P sp3.92 (1.88e)
20.77811.5371.64226.4% Al sp1.16 + 73.6% P sp4.93 (1.93e)
2Cl0.68431.4460.161224.4% Al sp0.83 + 75.6% P sp4.18 (1.92e)
2·LiCl0.63461.446−0.0042522.5% Al sp0.92 + 77.5% P sp3.27 (1.93e)
2·Et2O0.71961.6420.114825.0% Al sp0.62 + 75.0% P sp3.74 (1.93e)
1 Wiberg bond indices. 2 QE: Natural population analysis (NPA) charge on E. 3 Natural bond orbital.

Share and Cite

MDPI and ACS Style

Yanagisawa, T.; Mizuhata, Y.; Tokitoh, N. Syntheses and Structures of Novel λ33-Phosphanylalumanes Fully Bearing Carbon Substituents and Their Substituent Effects. Inorganics 2019, 7, 132. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7110132

AMA Style

Yanagisawa T, Mizuhata Y, Tokitoh N. Syntheses and Structures of Novel λ33-Phosphanylalumanes Fully Bearing Carbon Substituents and Their Substituent Effects. Inorganics. 2019; 7(11):132. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7110132

Chicago/Turabian Style

Yanagisawa, Tatsuya, Yoshiyuki Mizuhata, and Norihiro Tokitoh. 2019. "Syntheses and Structures of Novel λ33-Phosphanylalumanes Fully Bearing Carbon Substituents and Their Substituent Effects" Inorganics 7, no. 11: 132. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7110132

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