Next Article in Journal
Development of Ring-Expansion RAFT Polymerization of tert-Butyl Acrylate with a Cyclic Trithiocarbonate Derivative toward the Facile Synthesis of Cyclic Polymers
Next Article in Special Issue
Synthesis and Characterization of Novel Cobalt Carbonyl Phosphorus and Arsenic Clusters
Previous Article in Journal
Spin-Steered Photosynthesis of H2O2 in Magnetic Single-Atom Modified Covalent Triazine Frameworks: A Density Functional Theory Study
Previous Article in Special Issue
Ligand Hydrogenation during Hydroformylation Catalysis Detected by In Situ High-Pressure Infra-Red Spectroscopic Analysis of a Rhodium/Phospholene-Phosphite Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Structural Studies of peri-Substituted Acenaphthenes with Tertiary Phosphine and Stibine Groups †

by
Laurence J. Taylor
1,
Emma E. Lawson
2,
David B. Cordes
2,
Kasun S. Athukorala Arachchige
3,
Alexandra M. Z. Slawin
2,
Brian A. Chalmers
2,* and
Petr Kilian
2,*
1
School of Chemistry, University of Nottingham, Nottingham NG7 2RD, UK
2
EaStCHEM School of Chemistry, University of St Andrews, North Haugh, St Andrews, Fife KY16 9ST, UK
3
Centre for Microscopy and Microanalysis, The University of Queensland, St Lucia, QLD 4072, Australia
*
Authors to whom correspondence should be addressed.
This paper is dedicated to Professor J. Derek Woollins on the occasion of his retirement, for his towering contribution and unwavering support of all things related to main group chemistry.
Molecules 2024, 29(8), 1841; https://doi.org/10.3390/molecules29081841
Submission received: 21 March 2024 / Revised: 10 April 2024 / Accepted: 15 April 2024 / Published: 18 April 2024

Abstract

:
Two mixed peri-substituted phosphine-chlorostibines, Acenap(PiPr2)(SbPhCl) and Acenap(PiPr2)(SbCl2) (Acenap = acenaphthene-5,6-diyl) reacted cleanly with Grignard reagents or nBuLi to give the corresponding tertiary phosphine-stibines Acenap(PiPr2)(SbRR’) (R, R’ = Me, iPr, nBu, Ph). In addition, the Pt(II) complex of the tertiary phosphine-stibine Acenap(PiPr2)(SbPh2) as well as the Mo(0) complex of Acenap(PiPr2)(SbMePh) were synthesised and characterised. Two of the phosphine-stibines and the two metal complexes were characterised by single-crystal X-ray diffraction. The peri-substituted species act as bidentate ligands through both P and Sb atoms, forming rather short Sb-metal bonds. The tertiary phosphine-stibines display through-space J(CP) couplings between the phosphorus atom and carbon atoms bonded directly to the Sb atom of up to 40 Hz. The sequestration of the P and Sb lone pairs results in much smaller corresponding J(CP) being observed in the metal complexes. QTAIM (Quantum Theory of Atoms in Molecules) and EDA-NOCV (Energy Decomposition Analysis employing Naturalised Orbitals for Chemical Valence) computational techniques were used to provide additional insight into a weak n(P)→σ*(Sb-C) intramolecular bonding interaction (pnictogen bond) in the phosphine-stibines.

1. Introduction

Tertiary amines and phosphines play a key role as tuneable ligands, with uses in transition metal catalysis and other applications. The heavier tertiary pnictines (ER3, E = As, Sb, Bi, R = alkyl, aryl) also serve as L-type ligands in a number of complexes, although they generally display lower donor strength than corresponding N and P-based ligands [1,2].
Several unusual properties stemming from the close-proximity of two pnictine groups in peri-substituted scaffolds have been noted. The first of these was in the 1960’s, when the remarkably high basicity of proton sponge 1,8-bis(dimethylamino)naphthalene (A, Figure 1) was reported by Alder [3]. Tertiary bis(phosphines) such as the phosphorus analogues of the proton sponge B (Figure 1) were reported shortly thereafter [4,5], as were several of their metal complexes [6,7,8].
Syntheses of peri-substituted tertiary bis(arsines), such as C (Figure 1), and their complexes were also reported as early as the 1960’s [9]. However, no crystal structures of such ligands or complexes have appeared in the literature. Prototypical naphthalene bis(stibines) with dimethylstibino (D1) and diphenylstibino groups (D2) were synthesised by Reid, together with their Mo(0) and Pt(II) complexes [10]. The aryl species D2 (both naphthalene and acenaphthene variants) and D3 (naphthalene variant) were recently structurally characterised by Schulz, together with the two bis(bismuthines) E [11,12]. However, no structural data for any of the bis(stibine) or bis(bismuthine) metal complexes have been published to date.
The related Sb−Sb bonded species F [12], as well as the doubly backboned species G [13,14,15] and H [11], have received significant attention recently, and a few transition metal complexes with these as ligands have also been reported [13].
Species I (Figure 1) with two differing Group 15 peri-atoms display intriguing dative bonding and NMR properties [16,17]. Surprisingly, only two bis(tertiary) phosphine-stibine and phosphine-bismuthine peri-substituted species have been reported to date: ISb [18] and IBi [19]. Both of these display repulsive interactions between the two pnictogen-centred groups, although their geometries indicate a weak pnictogen bond (nP→σ*(Pn-C)) is present, as indicated in Figure 1 by a dashed line. This is in contrast to the related E(III)−E(III) halophosphines, such as J [18,20] and K [18,21], which display strong dative pnictogen-pnictogen bonds.
Apart from the phosphine-stibine ISb [18], the most closely related work to this paper are the geminally substituted bis- and tris(acenaphthene) species L and M [22,23]. Only one metal complex of these has been structurally characterised, the Rh(I) species N [23].
As a continuation of our synthetic, structural and bonding studies of peri-substituted species, we investigated the utility of the halostibines J and K (Figure 1), reported by us earlier [18], as synthons towards primary stibine functionalities. Prompted by the paucity of the literature data, we have also probed the coordination chemistry of the produced tertiary phosphine-stibines.

2. Results and Discussion

2.1. Synthesis and Spectroscopic Properties of the Tertiary Stibines 4, 5, 7, and 8

The three major precursors for the syntheses reported in this paper were bis(aryl) stibine 2, chloro(aryl) stibine 3 and dichlorostibine 6 (Scheme 1). Syntheses and structural information for these compounds, starting from 1, have been reported by us [18].
The reactive Sb−Cl motifs in chlorostibine 3 and dichlorostibine 6 were used to form new Sb−C bonds via reactions with carbon nucleophiles. The chlorostibine 3 was reacted with one equivalent of alkyl-Grignard reagents, MeMgBr and iPrMgCl, to afford alkyl-aryl stibines 4 (86%) and 5 (81%), respectively.
The reaction of dichlorostibine 6 with two equivalents of MeMgBr afforded dimethylstibine 7 as an off-white solid (yield ca. 90%; exact yield determination was not possible as 1H NMR indicated solvation by Et2O). Reacting nBuLi with 6 also resulted in Sb−C bond formation; reaction with two equivalents of nBuLi afforded crude di-n-butylstibine 8 in quantitative yield (obtained as an oil). The crystallisation of 8 from common organic solvents was not successful. However, a small amount of crystalline 8 was obtained through the long standing of the oil at room temperature (see below).
All the Sb−C bond-forming reactions were remarkably clean, as judged by 31P{1H} and 1H NMR spectroscopy. The newly prepared compounds were further characterised by 13C DEPT-Q NMR, HRMS (peaks corresponding to (M + H)+ with correct isotopic patterns were observed in all cases) and (for 7 and 8) also by Raman spectroscopy. The purity of 5 was confirmed by CHN microanalysis. The novel tertiary stibines appear to be hydrolytically stable (in some cases, an aqueous wash was involved in the work-up); however they are oxidised in the presence of air. Both 7 and 8 decomposed in chloroform solutions within several days, indicating instability in halogenated solvents.
The reaction of 6 with one equivalent of nBuLi gave an oil after the workup. This oil was shown by 31P{1H} NMR to be a complex mixture, with a major peak at δP 18.5 ppm, corresponding to the doubly substituted species 8, indicating that selective single substitution using an organolithium as a nucleophile may be difficult to achieve.
The 31P{1H} NMR spectra of the phosphine-stibines 4, 5, 7 and 8 display singlets within a narrow range of δP (−18.5 to −20.8 ppm). Notable through-space couplings (indicated by the TS superscript in the J notation, TSJ) are observed in the 13C{1H} NMR spectra between the phosphorus atom and carbons attached to the antimony atom. In 4, the ipso-C of the Sb-Ph moiety shows a 5TSJCP of 16.2 Hz. An even larger 5TSJCP of 34.1 Hz is observed for the Sb-CH3 of 4. Interestingly, the acenaphthene ipso-carbon atom shows no detectable coupling to the phosphorus atom, despite having a shorter bond path (formally 3JCP).
A similar situation is observed in 5 (5TSJCP = 17.3 Hz (ipso-Ph) and 5TSJCP = 36.8 Hz (Sb-CH), although in this case small magnitude splitting with the ipso-acenaphthene carbon (C1 in the numbering scheme shown in Figure 2) is observable (3JCP = 2.4 Hz). Similar magnitudes of JCP involving carbon atoms bonded directly to Sb atoms are also observed for 7 (5TSJCP = 34.9 Hz (CH3); 3JCP = 5.1 Hz (C1, Acenap)) and 8 (5TSJCP = 30.6 Hz (CH2); 3JCP = 5.0 Hz (C1, Acenap)). Observation of the through-space couplings in 4, 5, 7 and 8 is consistent with the significant overlap of P and Sb lone pairs as confirmed by single crystal X-ray diffraction (vide infra) and is in agreement with observations made in our previous study of P-Sb acenaphthenes [18].

2.2. Synthesis and Spectroscopic Properties of Tertiary Stibine Metal Complexes 2.PtCl2 and 4.Mo(CO)4

Peri-substituted species 2 and 4, bearing tertiary phosphine and tertiary stibine groups, were reacted with platinum(II) and molybdenum(0) motifs to explore their coordination chemistry. It was of interest to see if the phosphine-stibine species would act as bidentate ligands, with the metal coordinating through both phosphorus and antimony atoms.
[PtCl2(cod)] was reacted with 2 in dichloromethane, giving 2.PtCl2 as a yellow powder in a good yield (76%). Similarly, the reaction of [Mo(CO)4(nbd)] with 4 in dichloromethane gave 4.Mo(CO)4 as a brown powder in a near-quantitative yield (Scheme 1). Both complexes were stable to air in the solid and solution in the chlorinated solvents used to acquire their NMR spectra (CD2Cl2 and CDCl3, respectively).
The 31P{1H} NMR spectrum of 2.PtCl2 consists of a singlet with a set of 195Pt satellites (δP 7.8 ppm, 1JPPt = 3357 Hz), with the complementary doublet observed in the 195Pt{1H} NMR spectrum (δPt −4541 ppm). The coordination of platinum centres resulted in a high-frequency shift (c.f. free ligand 2, δP −21.9 ppm) [18] as well as loss of the through-space JCP coupling (c.f. 5TSJCP 40.3 Hz for ipso-Ph carbon in 2).
Coordination of the Mo(CO)4 fragment to 4 resulted in an even more pronounced high-frequency shift for 4.Mo(CO)4 (δP 43.4; c.f. δP −19.6 ppm in free ligand 4). Similar to 2.PtCl2, the JCP couplings between the phosphorus atom and the carbon atoms adjacent to the antimony atom are much smaller magnitudein 4.Mo(CO)4 than 4. This is notable as the through-bond coupling paths are shorter in the complex (formally 3JCP, 2.3 and 2.9 Hz), compared to those in the free ligand 4 (5TSJCP, 16.2 and 34.1 Hz).

2.3. Structural Discussion

Two of the phosphine-stibines (5 and 8), as well as the two complexes 2.PtCl2 and 4.Mo(CO)4, were subjected to the single crystal diffraction study. The structures are shown in Figure 2 and Figure 3 and Table 1.
Crystals of 5 were grown from ethanol. The structure of 5 displays a moderately strained geometry, with a P9∙∙∙Sb1 distance of 3.172(3) Å (129% of ∑rcovalent, 76% of ∑rvdW) [24,25] and a splay angle of 15.1(12)°. These parameters indicate that, while the two functional groups in the peri-positions are forced into close proximity, the P∙∙∙Sb interaction is primarily repulsive. However, a more detailed look at the peri-region geometry indicates the presence of a weak intramolecular pnictogen bond (n(P)→σ*(Sb–CiPr)), which manifests through a quasi-linear arrangement of the P9∙∙∙Sb1−C19 motif (168.5°, see Figure 3).
Crystals of 8 were obtained by prolonged standing of the crude oily product. The molecule of 8 in the structure displays a similar geometry to 5, with a slightly larger P∙∙∙Sb distance of 3.218(2) Å. In contrast to 5, the “homoleptic” substitution pattern of the Sb atom in 8 allows direct comparison of the Sb-C bond lengths for the two n-butyl groups. This reveals that the Sb1−C13 bond length is significantly elongated compared to the Sb1−C17 bond length (2.197(6) vs. 2.100(7) Å). This indicates the donation of electron density (n(P9)) into the (antibonding) σ*(Sb1−C13) orbital, consistent with the formation of a (quasi-linear) pnictogen bond n(P9)→σ*(Sb1–C3), P9···Sb1−C13 angle 167.1°, see Figure 3.
Crystals of 2.PtCl2 were grown from dichloromethane/hexane with a solvated molecule of dichloromethane. The platinum atom adopts a distorted square planar geometry, with the P and Sb atoms of ligand 2 bound in a cis fashion. Coordination of the PtCl2 fragment results in elongation of the P∙∙∙Sb distance to 3.357(12) Å (c.f. 3.191(1) Å in 2) and significantly increased out-of-plane distortions within the acenaphthene ligand (see Table 1) [18]. While the P−Pt distance (2.248(4) Å) is as expected, the Sb−Pt distance in 2.PtCl2 (2.4570(10) Å) is one of the shortest Sb–Pt bonds known, most likely due to the geometric constraints of the ligand. Of the 143 Sb–Pt bonds recorded in the Cambridge Structural Database to date, only 6 are shorter than the bond in compound 2.PtCl2. Those 6 examples are all Sb(V) species with highly electrophilic Sb centres [23,26,27,28], hence 2.PtCl2 is the shortest Pt−Sb bond for a stibine (R3Sb) ligand.
Crystals of 4.Mo(CO)4 were grown from hexane. The molybdenum adopts a (distorted) octahedral geometry as expected, with ligand 4 attached in cis fashion. As above, the P-Mo distance is as expected; however, the Sb–Mo bond length of 2.7007(6) Å, is one of the shortest Sb–Mo bonds known. Of the 518 independent Sb–Mo bonds (in 97 compounds) recorded in the Cambridge Structural Database, only 8 are shorter than the bond in compound 4.Mo(CO)4. The three compounds showing the shortest distances (the shortest being 2.64386(19) Å) are all stibine (or halostibine) complexes, possessing a tridentate scaffold combining stibine and phosphine functionalities, with similar constraints as those seen in 4 [29].

2.4. Computational Analysis

DFT calculations were employed to further investigate the nP→σ*(Sb–C) interaction in 5 and 8. Geometry optimisations were performed on these compounds (PBE0/SARC-ZORA-TZVP for Sb, PBE0/ZORA-def2-TZVP for all other atoms), and the resulting structures were in good agreement with the X-ray geometries. In particular, the Sb−C bond lengths in 8 were well reproduced, with the Sb−C bond opposite the P atom being elongated (Sb1−C13 2.197(6) Å experimental, 2.208 Å calculated; Sb1−C17 2.100(7) Å experimental, 2.175 Å calculated).
A Quantum Theory of Atoms in Molecules (QTAIM) [30,31] analysis was applied to 5 and 8. Bond critical points (BCPs) were located between the Sb1 and P9 atoms for both 5 and 8, indicative of a bonding interaction. Selected QTAIM parameters evaluated at BCPs for these molecules are summarized in Table 2. The Sb1···P9 BCPs all display a relatively low electron density (ρBCP) and a small and positive Laplacian (∇2ρBCP), which are typical of interactions between heavier elements [32,33].
The bond degree parameter [32] (BD = HBCPBCP; HBCP = energy density at BCP) [32,34] and the ratio of |VBCP|/GBCP [32] (VBCP = electronic potential energy at the BCP and GBCP = electronic kinetic energy at the BCP) are two valuable metrics in QTAIM for analysing bonds between heavier elements. The BD indicates the amount of covalency in a bond, with larger negative values denoting a greater covalent interaction [32,34]. The P9∙∙∙Sb1 interactions in 5 and 8 both show small, negative values, suggesting a weakly covalent interaction (Table 2). | BD | is smaller for 8 than 5, suggesting less covalency in the P9∙∙∙Sb1 interaction for 8. This can be rationalised by the Sb(nBu)2 moiety being more electron-rich than Sb(Ph)iPr, and thus a poorer electron acceptor. The interaction energy (Ei = 1 2 V B C P ) [35], which can be used as a rough estimate of bond strength, similarly indicates a weaker P9∙∙∙Sb1 interaction in 8.
The |VBCP|/GBCP ratio differentiates between different bond types: purely closed-shell interactions such as van der Waals or ionic bonds exhibit |VBCP|/GBCP < 1, while fully covalent interactions show |VBCP|/GBCP > 2. Bonds with intermediate ratios (1 < |VBCP|/GBCP < 2) are termed transit closed-shell interactions, such bonds possess partial covalent character [32]. Both 5 and 8 display 1 < |VBCP|/GBCP < 2, with a larger value for 5 than 8. This once again suggests a more covalent P9∙∙∙Sb1 in 5 than 8. Crucially, the BD and |VBCP|/GBCP suggest that the P9∙∙∙Sb1 interaction in both 5 and 8 is not purely closed shell (i.e., Van der Waals), and that there is some degree of electron sharing between the P and Sb atoms, consistent with a nP→σ*(Sb–C) interaction. Also of note is the difference in QTAIM parameters for Sb1−C13 and Sb1−C17 in 8. Sb−C13 shows a slightly reduced BD, |VBCP|/GBCP and Ei compared with Sb−C17 (Table 2), consistent with a weakening of the Sb1–C13 bond due to donation into the Sb–C σ* orbital.
This P9∙∙∙Sb1 interaction was further probed by an Energy Decomposition Analysis employing Naturalised Orbitals for Chemical Valence (EDA-NOCV) [36,37,38,39]. This allows the donor-acceptor interaction between P9 and Sb1 to be visualised and also allows for quantification of the interaction energy. For this analysis, the molecules were divided into two closed-shell fragments: an Acenap(PiPr2) anion and an SbR2+ cation. The total interaction energy between these fragments (ΔEint) was computed and divided into terms for ΔEsteric, ΔEorb and ΔEdisp (Table 3). ΔEorb and ΔEdisp are the orbital and dispersion interaction energies, respectively. ΔEsteric is the combined electrostatic attraction and Pauli-repulsion energy terms [40]. In both 5 and 8, ΔEsteric is negative, indicating a significant electrostatic attraction. This is a result of formally assigning the fragments as cationic and anionic. 8 is observed to have a slightly smaller ΔEint, ΔEsteric, ΔEorb and ΔEdisp than 5, which can again be contributed to more electron rich groups on Sb weakening the donor-acceptor interaction.
The ΔEorb term can be broken down into pairs of natural orbitals for chemical valence (NOCVs), which represent the orbital interactions between the Acenap(PiPr2) and SbR2+ fragments. For each pair of NOCVs, a deformation density plot (Δρk), which represents the flow of electrons between the molecular fragments, and its corresponding energy contribution to ΔEorb, can be determined [36]. The first deformation density plots (Δρ1) for 5 and 8, which have the largest energetic contribution to ΔEorb, are dominated by electron flow from the (anionic) carbon of Acenap(PiPr2) to the Sb atom. However, Δρ2 and Δρ3 for 5 and 8 both appear to show electron donation from the P lone pair to a Sb−C σ* orbital (Figure 4). The energy contributions of these interactions are Δρ2 = −15.8 kcal mol−1, Δρ3 = −12.1 kcal mol−1 for 5; Δρ2 = −14.4 kcal mol−1, Δρ3 = −11.0 kcal mol−1 for 8. These values are likely a significant overestimate of the nP→σ*(Sb–C) interaction energy, as the deformation density plots also show significant contributions from the π-systems of 5 and 8. However, these plots do strongly support the existence of donor-acceptor interactions between P9 and Sb1 in both compounds. Note that blue isosurface (an indicator of accepting electron density) is primarily observed on the Sb–C bond opposite the P-atom and not the other Sb–Ph (5) or Sb–nBu (8) bond (Figure 4).

3. Experimental Section

3.1. General Considerations

Unless otherwise stated, all experimental procedures were carried out under an atmosphere of dry nitrogen using standard Schlenk techniques or under an argon atmosphere in a Saffron glove box. Dry solvents were used unless otherwise stated and were either collected from an MBraun SPS-800 Solvent Purification System, or dried and stored according to literature procedures [41]. The peri-substituted acenapthene precursors 1 [42], 2, 3 and 6 [18] were synthesised according to literature procedures. “In vacuo” refers to a pressure of ca. 2 × 10−2 mbar.

3.2. NMR Spectroscopy

All novel compounds were characterised where possible by 1H, 13C DEPTQ and 31P{1H} NMR spectroscopy, including measurements of 1H{31P}, H-H DQF COSY, H-C HSQC, H-C HMBC and H-P HMBC. 13C{1H} NMR spectra were recorded using the DEPTQ-135 pulse sequence with broadband proton decoupling. Measurements were performed at 20 °C using a Bruker Avance 300, Bruker Avance II 400 or Bruker Avance III 500 (MHz) spectrometer. For both 1H and 13C NMR, chemical shifts are relative to Me4Si, which was used as an external standard. The residual solvent peaks were used for calibration (CHCl3, δH 7.26, δC 77.16 ppm; CD2Cl2, δH 5.32, δC 53.84 ppm). For 31P NMR, 85% H3PO4 in D2O (δP 0 ppm) was used as an external standard. 195Pt NMR was acquired for 2.PtCl2, and 1.2 M Na2[PtCl6] in D2O (δPt 0 ppm) was used as the external standard. The NMR numbering scheme is shown in Figure 5.

3.3. Other Analyses

Elemental analyses (C, H and N) were performed at London Metropolitan University. High resolution mass spectrometry was performed by the EPSRC UK National Mass Spectrometry Facility (NMSF) at Swansea University using either a Waters Xevo G2-S (ASAP) or a Thermofisher LTQ Orbitrap XL (APCI) mass spectrometer. Electrospray ionisation (ES) spectra were acquired at the University of St Andrews Mass Spectrometry Facility using a Thermo Exactive Orbitrap Mass Spectrometer. Both IR and Raman spectra were collected on a Perkin Elmer 2000 NIR FT spectrometer. KBr tablets were used in IR measurements; powders in sealed glass capillaries were used for Raman spectra acquisitions. Melting (or decomposition) points were determined by heating solid samples in glass capillaries using a Stuart SMP30 melting point apparatus.

3.4. [iPr2P-Ace-SbPh2]PtCl2, 2.PtCl2

To a suspension of dichloro(1,5-cyclooctadiene)platinum(II) (72 mg, 170 µmol) in dichloromethane (4 mL), a solution of 2 (100 mg, 170 µmol) in dichloromethane (10 mL) was added dropwise. The solution was left to stir at room temperature overnight. The volatiles were removed in vacuo to give 2.PtCl2 as a pale-yellow powder (104 mg, 76%). Crystals suitable for X-ray diffraction were grown from the vapour diffusion of hexane into a saturated solution of the compound in dichloromethane. M.p. 108 °C with decomposition.
1H NMR: δH (500.1 MHz, CD2Cl2) 8.09 (1H, dd, 3JHP 11.0, 3JHH 7.6 Hz, H-8), 7.71 (1H, d, 3JHH 7.0 Hz, H-2), 7.67–7.63 (4H, m, o-Ph CH), 7.57 (1H, d, 3JHH 7.5 Hz, H-7), 7.54–7.51 (2H, m, p-Ph CH), 7.49–7.44 (5H, m, H-3, m-Ph CH), 3.55–3.45 (6H, m, H-11, 12, 2× iPr CH), 1.36 (6H, dd, 3JHP 18.2, 3JHH 7.0 Hz, 2× iPr CH3), 1.17 (6H, dd, 3JHP 16.1, 3JHH 7.0 Hz, 2× iPr CH3).
13C DEPTQ NMR: δC (125.8 MHz, CD2Cl2) 152.9 (s, qC-6), 152.7 (s, qC-4), 140.3 (d, 3JCP 7.2 Hz, qC-5), 139.8 (d, 2JCP 6.8 Hz, qC-10), 139.1 (s, C-2), 135.7 (d, 2JCP 4.3 Hz, C-8), 135.5 (s, o-Ph CH), 131.1 (s, m-Ph CH), 129.4 (s, p-Ph CH), 126.6 (s, i-Ph qC), 120.4 (s, C-3), 119.4 (d, 3JCP 9.4 Hz, C-7), 114.0 (d, 1JCP 48.4 Hz, qC-9), 111.2 (d, 3JCP 7.8 Hz, qC-1), 30.5 (s, C-11/12), 30.0 (s, C-11/12), 29.9 (d, 1JCP 34.7 Hz, 2× iPr CH), 19.4 (s, 2× iPr CH3), 19.3 (s, 2 × iPr CH3).
31P{1H} NMR: δP (202.5 MHz, CD2Cl2) 7.8 (s with 195Pt satellites, 1JPPt 3357.0 Hz).
195Pt{1H} NMR: δPt (107.0 MHz, CD2Cl2) −4541 (d, 1JPtP 3357.0 Hz).
IR (KBr disc, cm−1) νmax 3047m (νAr–H), 2925s (νC–H), 1601s, 1479m, 1434vs, 1334m, 1254m, 1033m, 998w, 848m, 734vs, 693s, 451m, 270m, 241s.
Raman (glass capillary, cm−1) νmax 3052s (νAr–H), 2926s (νC–H), 1604m, 1578m, 1444m, 1337m, 1000vs, 661s, 581m, 319s, 180vs.
MS (ES+): m/z (%) 775.06 (100) [M − Cl].

3.5. iPr2P-Ace-Sb(Ph)Me, 4

To a stirred suspension of 3 (0.50 g, 0.99 mmol) in tetrahydrofuran (20 mL), cooled to −78 °C, methylmagnesium bromide (0.50 mL, 3.0 M solution in diethyl ether, 1.5 mmol) was added dropwise. The reaction mixture was allowed to warm to room temperature and stirred for 1 h, then cooled to 0 °C, and degassed water (0.5 mL) was added cautiously. Volatiles were removed in vacuo, and the resulting oil was redissolved in hexane (40 mL). The resulting suspension was filtered to remove insoluble impurities, and volatiles were removed in vacuo to afford 4 as a pale-yellow oil (0.41 g, 0.85 mmol, 86%).
1H NMR δH (400 MHz, CDCl3) 7.67–7.59 (4H, m, ArH-2, ArH-8, m-Ph CH), 7.32 (1H, d, 3JHH = 7.1 Hz, ArH-7), 7.29–7.25 (3H, m, o/p-Ph CH), 7.16 (1H, d, 3JHH = 7.1 Hz, ArH-3), 3.38 (4H, s, H-11, H-12, 2 × CH2), 2.25 (1H, septd, 3JHH = 6.9 Hz, 2JHP = 5.1 Hz, iPr CH), 2.09 (1H, septd, 3JHH = 7.0 Hz, 2JHP = 3.2 Hz, iPr CH), 1.22 (3H, dd, 3JHP = 15.0 Hz, 3JHH = 6.9 Hz, iPr CH3), 1.18 (3H, d, 6JHP = 1.8 Hz, Me(Sb)), 1.04 (3H, dd, 3JHP = 14.7 Hz, 3JHH = 6.9 Hz, iPr CH3), 0.97 (3H, dd, 3JHP = 12.5 Hz, 3JHH = 7.0 Hz, iPr CH3), 0.66 (3H, dd, 3JHP = 11.8 Hz, 3JHH = 7.0 Hz, iPr CH3).
13C{1H} NMR δC (75 MHz, CDCl3) 149.1 (s, qC-6), 147.2 (d, 4JCP = 1.6 Hz, qC-4), 146.0 (d, 1JCP = 40.5 Hz, qC-9), 142.1 (s, qC-1), 140.0 (d, 3JCP = 7.8 Hz, qC-5), 138.2 (s, C-2), 136.6 (s, m-Ph CH), 134.0 (s, qC-10), 133.9 (d, 2JCP = 2.4 Hz, C-8), 130.1 (d, 5JCP = 16.2 Hz, i-Ph qC), 128.4 (s, o-Ph CH), 127.6 (s, p-Ph CH), 120.0 (s, C-3), 119.0 (s, C-7), 30.3 (s, C-11/C-12), 30.0 (s, C-11/C-12), 26.1 (d, 1JCP = 12.5 Hz, iPr CH), 25.9 (d, 1JCP = 13.7 Hz, iPr CH), 20.6 (d, 2JCP = 17.7 Hz, iPr CH3), 20.1 (d, 2JCP = 13.3 Hz, iPr CH3), 20.0 (d, 2JCP = 7.3 Hz, iPr CH3), 19.2 (d, 2JCP = 7.2 Hz, iPr CH3), 4.74 (d, 5JCP = 34.1 Hz, Me(Sb)).
31P NMR δP (109 MHz, CDCl3) −19.7 (m).
31P{1H} NMR δP (109 MHz, CDCl3) −19.6 (s).
MS (APCI+) m/z 390.05 (85%, M − Ph − Me), 405.07 (100, M − Ph), 467.09 (73, M − Me), 483.12 (11, M + H), 499.11 (4, M + O + H).
HRMS (APCI+) C25H31PSb (M + H)+; calculated: 483.1196; found: 483.1195.

3.6. [iPr2P-Ace-Sb(Ph)Me]Mo(CO)4, 4.Mo(CO)4

A solution of compound 4 (240 mg, 0.497 mmol) in dichloromethane (40 mL) was added to a stirred suspension of cis-tetracarbonyl(norbornadiene)molybdenum(0) (0.164 g, 0.546 mmol) in dichloromethane (10 mL), and the resulting suspension was stirred at room temperature for 3 days. Insoluble material was removed by filtration through celite, and volatiles were removed in vacuo to afford 4.Mo(CO)4 as a pale brown solid (329 mg, 0.476 mmol, 96%). Crystals suitable for single crystal X-ray diffraction were grown from hexane at −25 °C.
1H NMR δH (400 MHz, CDCl3) 7.79 (1H, d, 3JHH = 6.9 Hz, ArH-2), 7.70 (1H, ≈ t, 3JHP = 7.8 Hz, 3JHH = 7.8 Hz, ArH-8), 7.56–7.50 (2H, m, m-Ph CH), 7.39–7.33 (4H, m, ArH-7, o/p-Ph CH), 7.30 (1H, d, 3JHH = 6.9 Hz, ArH-3), 3.40 (4H, s, H-11/H-12), 2.50–2.38 (1H, m, iPr CH), 2.35–2.25 (1H, m, iPr CH), 1.58 (s, 3H, Me(Sb)), 1.23 (3H, dd, 3JPH = 15.4 Hz, 3JHH = 6.9 Hz, iPr CH3), 1.10 (3H, dd, 3JPH = 15.8 Hz, 3JHH = 7.1 Hz, iPr CH3), 1.04 (3H, dd, 3JPH = 15.6 Hz, 3JHH = 7.1 Hz, iPr CH3), 0.99 (3H, dd, 3JPH = 15.0 Hz, 3JHH = 6.9 Hz, iPr CH3).
13C{1H} NMR δC (101 MHz, CDCl3) 218.4 (d, 2JCP = 8.1 Hz, CO), 216.4 (d, 2JCP = 20.3 Hz, CO), 211.1 (d, 2JCP = 9.3 Hz, CO), 210.8 (d, 2JCP = 9.1 Hz, CO), 150.6 (s, qC-6), 150.4 (d, 4JCP = 1.4 Hz, qC-4), 140.9 (d, 3JCP = 6.5 Hz, qC-5), 139.5 (d, 2JCP = 10.3 Hz, qC-10), 138.3 (s, C-2), 135.6 (d, 3JCP = 2.3 Hz, i-Ph qC), 133.9 (s, m-Ph CH), 133.3 (s, C-8), 129.4 (s, p-Ph CH), 129.0 (s, o-Ph CH), 124.6 (d, 1JCP = 20.4, qC-9), 123.4 (d, 3JCP = 2.4 Hz, qC-1), 119.8 (s, C-3), 118.9 (d, 3JCP = 5.7 Hz, C-7), 30.2 (s, C-11/C-12), 29.5 (s, C-11/C-12), 28.7 (d, 1JCP = 15.9 Hz, iPr CH), 28.1 (d, 1JCP = 14.9 Hz, iPr CH), 19.1 (d, 2JCP = 4.3 Hz, iPr CH3), 19.0 (d, 2JCP = 4.1 Hz, iPr CH3), 18.24 (d, 2JCP = 5.9 Hz, iPr CH3), 18.15 (d, 2JCP = 4.4 Hz, iPr CH3), 4.4 (d, 3JCP = 2.9 Hz, Me(Sb)).
31P NMR δP (109 MHz, CDCl3) 43.3 (m).
31P{1H} NMR δP (109 MHz, CDCl3) 43.4 (s).
IR (KBr disk, cm−1) νmax 3007 (νAr–H, w), 2963 (νC–H, m), 2014 (νC≡O, vs), 1899 (νC≡O, vs), 1603 (m), 1261 (m), 1084 (s), 1023 (s), 802 (m), 733 (m), 694 (m), 613 (m), 585 (m), 453 (m).
MS (APCI+) m/z 271.16 (100%, M − Mo(CO)4 − Sb(Me)Ph + H), 287.16 (M − Mo(CO)4 − Sb(Me)Ph + O + H), 637.02 (1, M − 2CO + H)
HRMS (APCI+) C27H31MoO2PSb (M − 2CO + H)+; calculated: 637.0149; found: 637.0148.

3.7. iPr2P-Ace-Sb(Ph)iPr, 5

To a stirred suspension of 3 (1.00 g, 1.98 mmol) in tetrahydrofuran (20 mL), cooled to −78 °C, a solution of isopropylmagnesium chloride (1.5 mL, 1.70 M solution in THF, 2.55 mmol) was added dropwise. The reaction mixture was allowed to warm to room temperature with stirring overnight, then cooled to 0 °C, and degassed water (0.5 mL) was added cautiously. Volatiles were removed in vacuo to give an oil. Hexane (50 mL) was added, and the resultant suspension was filtered to remove the insoluble salts. Volatiles were removed in vacuo to afford 5 as a yellow oil, which crystallised to a yellow solid on standing at room temperature for several days (0.824 g, 1.61 mmol, 81%). Crystals suitable for single crystal X-ray diffraction were grown from ethanol at −25 °C. M. p. 73–76 °C.
Elemental Analysis: C27H34PSb; calculated (%) C 63.43, H 6.70; found (%) C 63.31, H 6.73.
1H NMR δH (500 MHz; CDCl3) 7.85 (1H, d, 3JHH = 7.0 Hz, ArH-2), 7.61 (1H, dd, 3JHH = 7.1 Hz, 3JHP = 3.6 Hz, ArH-8), 7.58–7.54 (2H, m, m-Ph CH), 7.30 (1H, d, 3JHH = 7.1 Hz, ArH-3), 7.26 (1H, d, 3JHH = 7.1 Hz, ArH-7), 7.23–7.19 (3H, m, o/p-Ph CH), 3.39 (4H, s, H-11, H-12), 2.32 (1H, sept, 3JHH = 7.2 Hz, iPr(Sb) CH), 2.22 (1H, septd, 3JHH = 7.1 Hz, 2JHP = 3.4 Hz, iPr(P) CH), 2.04–1.93 (1H, m, iPr(P) CH), 1.36 (3H, d, 3JHH = 7.2 Hz, iPr(Sb) CH3), 1.25–1.19 (6H, m, iPr(Sb) CH3, iPr(P) CH3), 0.99 (3H, dd, 3JHP = 11.9 Hz, 3JHH = 7.0 Hz, iPr(P) CH3), 0.95 (3H, dd, 3JHP = 14.6 Hz, 3JHH = 6.9 Hz, iPr(P) CH3), 0.42 (3H, dd, 3JHP = 12.4 Hz, 3JHH = 7.0 Hz, iPr(P) CH3).
13C{1H} NMR δC (101 MHz; CDCl3) 148.9 (s, qC-6), 147.3 (d, 4JCP = 1.7 Hz, qC-4), 144.2 (d, 1JCP = 28.2 Hz, qC-9), 142.2 (d, 2JCP = 27.2 Hz, qC-10), 140.1 (d, 3JCP = 7.8 Hz, qC-5), 138.1 (s, C-2), 136.7 (s, m-Ph CH), 134.0 (d, 2JCP = 2.4 Hz, C-8), 133.9 (d, 3JCP = 5.6 Hz, qC-1), 130.7 (d, 5JCP = 17.3 Hz, i-Ph qC), 128.2 (s, o-Ph CH), 127.4 (s, p-Ph CH), 120.1 (s, C-7), 119.0 (s, C-3), 30.2 (s, C-11/C-12), 30.0 (s, C-11/C-12), 26.4 (d, 1JCP = 14.8 Hz, iPr(P) CH), 26.2 (d, 1JCP = 14.0 Hz, iPr(P) CH), 25.1 (d, 5JCP = 36.8 Hz, iPr(Sb) CH), 22.6 (s, iPr(Sb) CH3), 21.7 (s, iPr(Sb) CH3), 20.7 (d, 2JCP = 16.8 Hz, iPr(P) CH3), 20.1 (d, 2JCP = 8.1 Hz, iPr CH3), 19.9 (d, 2JCP = 16.6 Hz, iPr(P) CH3), 19.2 (d, 2JCP = 9.3 Hz, iPr(P) CH3).
31P NMR δP (109 MHz; CDCl3) −20.8 (m).
31P{1H} NMR δP (109 MHz; CDCl3) −20.8 (s).
Raman: (glass capillary, cm−1) νmax 3038s (νAr–H), 2921vs (νC–H), 1601s, 1562s, 1442s, 1325vs, 1001vs, 657m, 582s, 490s.
MS (APCI+) m/z 390.05 (100%, M − iPr − Ph), 433.10 (60, M − Ph), 467.09 (50, M − iPr), 511.15 (16, M + H), 527.15 (4, M + O + H).
HRMS (APCI+): C27H35PSb (M + H)+; calculated: 511.1509; found: 511.1510.

3.8. iPr2P-Ace-SbMe2, 7

A solution of 6 (0.49 g, 1.06 mmol) in diethyl ether (40 mL) was cooled to −78 °C. A solution of methylmagnesium bromide in tetrahydrofuran (0.7 mL, 3.0 M solution, 2.10 mmol, diluted with diethyl ether, 4 mL) was added dropwise with stirring over one hour. The resulting suspension was stirred at -78 °C for a further 90 min before warming to ambient temperature overnight. The suspension was filtered, and volatiles were removed from the filtrate in vacuo, affording 7 as an off-white solid (0.42 g, 94%). The yield is approximate, as 1H NMR spectra indicate the presence of solvated diethylether. M.p. 239 °C with decomposition.
1H NMR: δH (400.1 MHz, CDCl3) 7.77 (1H, d, 3JHH 7.0 Hz, H-2), 7.66 (1H, dd, 3JHH 7.1, 3JHP 3.6 Hz, H-8), 7.32 (1H, d, 3JHH 7.1 Hz, H-7), 7.26 (1H, d, 3JHH 7.0 Hz, H-3), 3.38 (4H, s, H-11, 12), 2.19 (2H, br sept, 3JHH 7.0 Hz, iPr CH), 1.17 (6H, dd, 3JHP 14.6, 3JHH 6.9 Hz, iPr CH3), 0.96 (6H, d, 6tsJHP 1.2 Hz, Sb-CH3), 0.94 (6H, dd, 3JHP 12.3, 3JHH 7.0 Hz, iPr CH3).
13C DEPTQ NMR: δC (100.6 MHz, CDCl3) 149.0 (s, qC-6), 147.0 (s, qC-4), 141.5 (d, 2JCP 4.3 Hz, qC-10), 139.8 (d, 3JCP 1.7 Hz, qC-5), 136.0 (s, C-2), 133.8 (d, 2JCP 2.5 Hz, C-8), 133.3 (d, 3JCP 5.1 Hz, qC-1), 130.3 (d, 1JCP 20.0 Hz, qC-9), 119.9 (s, C-3), 118.9 (s, C-7), 30.2 (s, C-11/12), 29.9 (s, C-11/12), 26.0 (d, 1JCP 13.5 Hz, iPr CH), 20.4 (d, 2JCP 16.7 Hz, iPr CH3), 19.9 (d, 2JCP 9.0 Hz, iPr CH3), 3.4 (d, 5tsJCP 34.9 Hz, Sb-CH3).
31P{1H} NMR: δP (162.0 MHz, CDCl3) −18.5 (s).
Raman: (glass capillary, cm−1) νmax 2933vs (νC-H), 2124w, 1601m, 1562m, 1447m, 1325vs, 577s, 509s, 482vs.
HRMS (ASAP+): m/z Calcd. for C20H29PSb 421.1045, found 421.1042 [M + H]; Calcd. for C19H25PSb 405.0732, found 405.0724 [M − Me].

3.9. iPr2P-Ace-Sb(nBu)2, 8

A solution of 6 (1.00 g, 2.16 mmol) in diethyl ether (40 mL) was cooled to −78 °C. To this, a solution of n-butyllithium in hexane (1.7 mL, 2.5 M solution, 4.25 mmol) was added dropwise with stirring over one hour. The resulting suspension was stirred at this temperature for a further hour before warming to ambient temperature overnight. The suspension was filtered, and the volatiles were removed from the filtrate in vacuo, affording 8 as a yellow oil (yield quantitative). A few crystals suitable for single-crystal X-ray diffraction formed spontaneously from the oil after long standing at room temperature. M. p. 53 °C.
1H NMR: δH (400.1 MHz, CDCl3) 7.73 (1H, d, 3JHH 7.0 Hz, H-2), 7.67 (1H, dd, 3JHH 7.1, 3JHP 3.6 Hz, C-8), 7.31 (1H, d, 3JHH 7.1 Hz, H-7), 7.25 (1H, d, 3JHH 7.0 Hz, H-3), 3.38 (4H, s, H-11, 12), 2.22 (2H, dsept, 3JHH 6.9, 2JHP 3.9 Hz, iPr CH), 1.71−1.48 (8H, m, SbCH2 and SbCH2CH2), 1.44−1.36 (4H, m, Sb CH2CH2CH2), 1.23 (6H, dd, 3JHP 14.4, 3JHH 6.9 Hz, 2× iPr CH3), 0.95 (6H, dd, 3JHP 12.4, 3JHH 7.2 Hz, 2× iPr CH3), 0.90 (6H, t, 3JHH 7.2 Hz, 2× n-Bu CH3).
13C DEPTQ NMR: δC (100.6 MHz, CDCl3) 148.9 (s, qC-6), 146.8 (d, 4JCP 1.7 Hz, qC-4), 142.1 (d, 2JCP 26.9 Hz, qC-10), 139.9 (d, 3JCP 7.8 Hz, qC-5), 136.5 (s, C-2), 133.8 (d, 2JCP 2.4 Hz, C-8), 131.9 (d, 3JCP 5.0 Hz, qC-1), 130.9 (d, 1JCP 18.0 Hz, qC-9), 119.9 (s, C-3), 118.9 (s, C-7), 30.2 (s, SbCH2CH2), 29.9 (s, C-11, 12), 27.0 (s, SbCH2CH2CH2), 26.4 (d, 1JCP 14.3 Hz, iPr CH), 20.7 (d, 5tsJCP 30.6 Hz, SbCH2), 20.5 (d, 2JCP 16.9 Hz, iPr CH3), 20.2 (d, 2JCP 9.5 Hz, iPr CH3), 14.0 (s, n-Bu CH3).
31P{1H} NMR: δP (162.0 MHz, CDCl3) −18.5 (s)
Raman: (glass capillary, cm−1) νmax 2919vs (νC−H), 2868vs, 1556m, 1435m, 1325s, 1157w, 583m.
HRMS (ASAP+): m/z Calcd. for C26H41PSb [M + H]: 505.1984, found 505.1986.

3.10. X-ray Diffraction

X-ray diffraction data for compound 2.PtCl2 were collected at 125 K using the St Andrews Automated Robotic Diffractometer (STANDARD) [43], consisting of a Rigaku sealed-tube X-ray generator equipped with a SHINE monochromator [Mo Kα radiation (λ = 0.71075 Å)], and a Saturn 724 CCD area detector, coupled with a Microglide goniometer head and an ACTOR SM robotic sample changer. Diffraction data for compounds 4.Mo(CO)4, 5 and 8 were collected at 173 K using a Rigaku FR-X Ultrahigh Brilliance Microfocus RA generator/confocal optics [Mo Kα radiation (λ = 0.71075 Å)] with an XtaLAB P200 diffractometer. Intensity data for all compounds were collected using ω steps, accumulating area detector images spanning at least a hemisphere of reciprocal space. Data for all compounds analysed were collected using CrystalClear [44] and processed (including correction for Lorentz, polarization and absorption) using either CrystalClear or CrysAlisPro [45]. Structures were solved by direct (SHELXS [46]), Pattterson (PATTY [47]) or charge-flipping (Superflip [48]) methods and refined by full-matrix least-squares against F2 (SHELXL-2019/3 [49]). Non-hydrogen atoms were refined anisotropically, and hydrogen atoms were refined using a riding model. In 5, both isopropyl groups bound to phosphorus were disordered over two positions. Atoms were split and refined with partial occupancies, and restraints to bond distances were required. Crystals of 8 were affected by pseudo-merohedral twinning, showing a twin law of [−0.9999 0.0198 0.0009 −0.0032 0.9992 −0.0203 −0.0007 −0.1315 −0.9981] and a refined twin fraction of 0.489. All calculations were performed using the CrystalStructure interface [50]. Selected crystallographic data are presented in Table 4.

3.11. Computational Methodology

3.11.1. Geometry Optimisations and QTAIM Analysis

Geometry optimisations were performed for models 5 and 8 using coordinates derived from their X-ray crystal structures. These models were geometry optimised without restraints using the ORCA 5.0.4 software package [51] utilising the PBE0 density functional [52] and all-electron ZORA corrected [53] def2-TZVP basis sets [54,55,56,57] for all atoms (except Sb), SARC-ZORA-TZVP [58] basis sets for the Sb atoms, along with SARC/J auxiliary basis sets decontracted def2/J up to Kr [59] and SARC auxiliary basis sets beyond Kr. [58,60,61,62]. Gradient corrections were performed with Grimme’s 3rd generation dispersion correction [63,64]. TightSCF and TightOpt convergence criteria were employed, and the location of true minima in these optimisations was confirmed by frequency analysis, which demonstrated that no imaginary vibrations were present.
Extended wavefunction (.wfx) files were generated from these optimisations using Orca 5.0.4 [51]. AIM analysis was performed using MultiWFN 3.8 [65].

3.11.2. EDA-NOCV Analysis

EDA-NOCV calculations were carried out on models of 5 and 8 using coordinates derived from the geometry-optimised structures. The compounds were divided into an Acenap(PiPr2) and SbR2+ fragments (5 = +Sb(iPr)Ph; 8 = +Sb(nBu)2). Calculations were carried out using the ORCA 5.0.4 software package [51] utilising the PBE0 density functional [52,66] and all-electron ZORA corrected [53] def2-TZVP basis sets [54,55,56,57] for all atoms (except Sb), SARC-ZORA-TZVP basis sets [58] for the Sb atoms, along with SARC/J auxiliary basis sets decontracted def2/J up to Kr [59] and SARC auxiliary basis sets beyond Kr [58,60,61,62]. Gradient corrections were performed with Grimme’s 3rd generation dispersion correction [63,64]. VeryTightSCF convergence settings and an integration accuracy value of 6.0 were employed. Deformation density plots were calculated from .cube files of the relevant NOCVs using the MultiWFN 3.8 software package [65] and visualised using Avogadro v1.2.0 [67].

4. Conclusions

The synthetic utility of peri-substituted phosphine-chlorostibines 3 and 6 in reactions with carbon nucleophiles has been demonstrated. Reactions with Grignard reagents or nBuLi proceeded rather cleanly and gave alkyl/aryl and alkyl tertiary stibines 4, 5, 7 and 8 with very good yields.
The coordination chemistry of the selected tertiary phosphine-stibines has also been probed. Two complexes, 2.PtCl2 and 4.Mo(CO)4, have been synthesised. Single-crystal X-ray diffraction confirmed that both the phosphine and the stibine groups are attached to the platinum(II) and Mo(0) centres.
In the phosphine-phosphine peri-substituted species, such as iPr2P-Ace-PPh2, large magnitude 4TSJPP (180 Hz for the above species) were observed due to the forced overlap of the two lone pairs on the phosphorus atoms [68]. As antimony has no spin ½ isotopes, the direct observation of P−Sb couplings (formally 4TSJPSb) was not possible in the phosphine-stibines reported here. However, long-range 5TSJCP couplings of up to 36.8 Hz were observed for carbon atoms attached to Sb atoms in all the phosphine stibines. This indicates the presence of a strong through-space coupling pathway through the phosphorus and antimony atoms (both possessing a lone pair), and the through-bond pathway contribution (5J) is expected to be negligible [69].
The QTAIM analysis supports the existence of a P∙∙∙Sb interaction in 5 and 8, which is not purely closed-shell (i.e., Van der Waals), and the visualisation of the deformation densities in an EDA-NOCV analysis supports the view that electron density from the P atom flows towards an apparent Sb–C σ*orbital.

Author Contributions

L.J.T., E.E.L. and B.A.C. carried out the required synthetic steps, collected all data (except X-ray), and analysed the data. A.M.Z.S., D.B.C. and K.S.A.A. collected the X-ray data and solved the structures. L.J.T. performed all computational analysis. P.K. and B.A.C. designed the study. P.K., L.J.T. and D.B.C. wrote the manuscript. P.K. provided the supervision. All authors have contributed to the proof-reading and editing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. We are grateful to the University of St Andrews School of Chemistry Undergraduate Project grants.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

CCDC 2337824-2337827 contains 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/structures. The research data underpinning this publication can be accessed at https://doi.org/10.17630/8f5cd85b-0e47-44ab-9d9a-e27d44a123f6 [70].

Conflicts of Interest

The authors declare no conflicts of interest.

References and Note

  1. Champness, N.R.; Levason, W. Coordination chemistry of stibine and bismuthine ligands. Coord. Chem. Rev. 1994, 133, 115–217. [Google Scholar] [CrossRef]
  2. Levason, W.; Reid, G. Developments in the coordination chemistry of stibine ligands. Coord. Chem. Rev. 2006, 250, 2565–2594. [Google Scholar] [CrossRef]
  3. Alder, R.W.; Bowman, P.S.; Steele, W.R.S.; Winterman, D.R. The remarkable basicity of 1,8-bis(dimethylamino)naphthalene. Chem. Commun. 1968, 723–724. [Google Scholar] [CrossRef]
  4. Costa, T.; Schmidbaur, H. 1,8-Naphthalindiylbis(dimethylphosphan): Konsequenzen sterischer Hinderung für Methylierung und Borylierung. Chem. Berichte 1982, 115, 1374–1378. [Google Scholar] [CrossRef]
  5. Karaçar, A.; Thönnessen, H.; Jones, P.G.; Bartsch, R.; Schmutzler, R. 1,8-Bis(phosphino)naphthalenes: Synthesis and molecular structures. Heteroat. Chem. 1997, 8, 539–550. [Google Scholar] [CrossRef]
  6. Yam, V.W.W.; Chan, C.L.; Choi, S.W.K.; Wong, K.M.C.; Cheng, E.C.C.; Yu, S.C.; Ng, P.-K.; Chan, W.-K.; Cheung, K.-K. Synthesis, photoluminescent and electroluminescent behaviour of four-coordinate tetrahedral gold(i) complexes. X-ray crystal structure of [Au(dppn)2]Cl. Chem. Commun. 2000, 53–54. [Google Scholar] [CrossRef]
  7. Karaçar, A.; Freytag, M.; Jones, P.G.; Bartsch, R.; Schmutzler, R. Platinum(II) Complexes of P-Chiral 1, 8-Bis(diorganophosphino)naphthalenes; Crystal Structures of dmfppn, rac-[PtCl2(dmfppn)], and rac-[PtCl2(dtbppn)] (dmfppn = 1, 8-di(methyl-pentafluorophenylphosphino)naphthalene and dtbppn = 1, 8-di(tert-butylphenylphosphino)naphthalene). Z. Anorg. Allg. Chem. 2002, 628, 533–544. [Google Scholar] [CrossRef]
  8. Kilian, P.; Knight, F.R.; Woollins, J.D. Synthesis of ligands based on naphthalene peri-substituted by Group 15 and 16 elements and their coordination chemistry. Coord. Chem. Rev. 2011, 255, 1387–1413. [Google Scholar] [CrossRef]
  9. Di Sipio, L.; Sindellari, L.; Tondello, E.; De Michelis, G.; Oleari, L. NiII complexes with 1,8-naphthalene bis(dimethylarsine). Coord. Chem. Rev. 1967, 2, 117–128. [Google Scholar] [CrossRef]
  10. Jura, M.; Levason, W.; Reid, G.; Webster, M. Preparation and properties of sterically demanding and chiral distibine ligands. Dalton Trans. 2008, 5774–5782. [Google Scholar] [CrossRef]
  11. Gehlhaar, A.; Wölper, C.; van der Vight, F.; Jansen, G.; Schulz, S. Noncovalent Intra- and Intermolecular Interactions in Peri-Substituted Pnicta Naphthalene and Acenaphthalene Complexes. Eur. J. Inorg. Chem. 2022, 2022, e202100883. [Google Scholar] [CrossRef]
  12. Gehlhaar, A.; Weinert, H.M.; Wölper, C.; Semleit, N.; Haberhauer, G.; Schulz, S. Bisstibane–distibane conversion via consecutive single-electron oxidation and reduction reaction. Chem. Commun. 2022, 58, 6682–6685. [Google Scholar] [CrossRef]
  13. Dzialkowski, K.; Gehlhaar, A.; Wolper, C.; Auer, A.A.; Schulz, S. Structure and Reactivity of 1,8-Bis(naphthalenediyl)dipnictanes. Organometallics 2019, 38, 2927–2942. [Google Scholar] [CrossRef]
  14. Ganesamoorthy, C.; Heimann, S.; Hölscher, S.; Haack, R.; Wölper, C.; Jansen, G.; Schulz, S. Synthesis, structure and dispersion interactions in bis(1,8-naphthalendiyl)distibine. Dalton Trans. 2017, 46, 9227–9234. [Google Scholar] [CrossRef] [PubMed]
  15. Gehlhaar, A.; Schiavo, E.; Wölper, C.; Schulte, Y.; Auer, A.A.; Schulz, S. Comparing London dispersion pnictogen–π interactions in naphthyl-substituted dipnictanes. Dalton Trans. 2022, 51, 5016–5023. [Google Scholar] [CrossRef] [PubMed]
  16. Chuit, C.; Corriu, R.J.P.; Monforte, P.; Reyé, C.; Declercq, J.-P.; Dubourg, A. Evidences for intramolecular N → P coordination in (8-dimethylamino-1-naphthyl) diphenylphosphane and derivatives. J. Organomet. Chem. 1996, 511, 171–175. [Google Scholar] [CrossRef]
  17. Biskup, D.; Bergmann, T.; Schnakenburg, G.; Gomila, R.M.; Frontera, A.; Streubel, R. Synthesis of a 1-aza-2-phospha-acenaphthene complex profiting from coordination enabled chloromethane elimination. RSC Adv. 2023, 13, 21313–21317. [Google Scholar] [CrossRef] [PubMed]
  18. Chalmers, B.A.; Bühl, M.; Athukorala Arachchige, K.S.; Slawin, A.M.Z.; Kilian, P. Structural, Spectroscopic and Computational Examination of the Dative Interaction in Constrained Phosphine–Stibines and Phosphine–Stiboranes. Chem.-A Eur. J. 2015, 21, 7520–7531. [Google Scholar] [CrossRef]
  19. Nejman, P.S.; Curzon, T.E.; Bühl, M.; McKay, D.; Woollins, J.D.; Ashbrook, S.E.; Cordes, D.B.; Slawin, A.M.Z.; Kilian, P. Phosphorus–Bismuth Peri-Substituted Acenaphthenes: A Synthetic, Structural, and Computational Study. Inorg. Chem. 2020, 59, 5616–5625. [Google Scholar] [CrossRef]
  20. Bergsch, J.U.; Slawin, A.M.Z.; Kilian, P.; Chalmers, B.A. Phosphine–Stibine and Phosphine–Stiborane peri-Substituted Donor–Acceptor Complexes. Molbank 2023, 2023, M1653. [Google Scholar] [CrossRef]
  21. Hupf, E.; Lork, E.; Mebs, S.; Chęcińska, L.; Beckmann, J. Probing Donor–Acceptor Interactions in peri-Substituted Diphenylphosphinoacenaphthyl–Element Dichlorides of Group 13 and 15 Elements. Organometallics 2014, 33, 7247–7259. [Google Scholar] [CrossRef]
  22. Chalmers, B.A.; Meigh, C.B.E.; Nejman, P.S.; Bühl, M.; Lébl, T.; Woollins, J.D.; Slawin, A.M.Z.; Kilian, P. Geminally Substituted Tris(acenaphthyl) and Bis(acenaphthyl) Arsines, Stibines, and Bismuthine: A Structural and Nuclear Magnetic Resonance Investigation. Inorg. Chem. 2016, 55, 7117–7125. [Google Scholar] [CrossRef] [PubMed]
  23. Furan, S.; Hupf, E.; Boidol, J.; Brunig, J.; Lork, E.; Mebs, S.; Beckmann, J. Transition metal complexes of antimony centered ligands based upon acenaphthyl scaffolds. Coordination non-innocent or not? Dalton Trans. 2019, 48, 4504–4513. [Google Scholar] [CrossRef] [PubMed]
  24. Batsanov, S.S. Van der Waals Radii of Elements. Inorg. Mater. 2001, 37, 871–885. [Google Scholar] [CrossRef]
  25. Cordero, B.; Gomez, V.; Platero-Prats, A.E.; Reves, M.; Echeverria, J.; Cremades, E.; Barragan, F.; Alvarez, S. Covalent radii revisited. Dalton Trans. 2008, 2832–2838. [Google Scholar] [CrossRef] [PubMed]
  26. You, D.; Smith, J.E.; Sen, S.; Gabbaï, F.P. A Stiboranyl Platinum Triflate Complex as an Electrophilic Catalyst. Organometallics 2020, 39, 4169–4173. [Google Scholar] [CrossRef]
  27. You, D.; Gabbaï, F.P. Unmasking the Catalytic Activity of a Platinum Complex with a Lewis Acidic, Non-innocent Antimony Ligand. J. Am. Chem. Soc. 2017, 139, 6843–6846. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, H.; Gabbaï, F.P. Solution and Solid-State Photoreductive Elimination of Chlorine by Irradiation of a [PtSb]VII Complex. J. Am. Chem. Soc. 2014, 136, 10866–10869. [Google Scholar] [CrossRef] [PubMed]
  29. Piesch, M.; Gabbaï, F.P.; Scheer, M. Phosphino-Stibine Ligands for the Synthesis of Heterometallic Complexes. Z. Anorg. Allg. Chem. 2021, 647, 266–278. [Google Scholar] [CrossRef]
  30. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Clarendon Press: Oxford, UK, 1994. [Google Scholar]
  31. Bader, R.F. A quantum theory of molecular structure and its applications. Chem. Rev. 1991, 91, 893–928. [Google Scholar] [CrossRef]
  32. Lepetit, C.; Fau, P.; Fajerwerg, K.; Kahn, M.L.; Silvi, B. Topological analysis of the metal-metal bond: A tutorial review. Coord. Chem. Rev. 2017, 345, 150–181. [Google Scholar] [CrossRef]
  33. Gervasio, G.; Bianchi, R.; Marabello, D. About the topological classification of the metal–metal bond. Chem. Phys. Lett. 2004, 387, 481–484. [Google Scholar] [CrossRef]
  34. Espinosa, E.; Alkorta, I.; Elguero, J.; Molins, E. From weak to strong interactions: A comprehensive analysis of the topological and energetic properties of the electron density distribution involving X–H⋯F–Y systems. J. Chem. Phys. 2002, 117, 5529–5542. [Google Scholar] [CrossRef]
  35. Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen bond strengths revealed by topological analyses of experimentally observed electron densities. Chem. Phys. Lett. 1998, 285, 170–173. [Google Scholar] [CrossRef]
  36. Mitoraj, M.; Michalak, A. Natural orbitals for chemical valence as descriptors of chemical bonding in transition metal complexes. J. Mol. Model. 2007, 13, 347–355. [Google Scholar] [CrossRef] [PubMed]
  37. Michalak, A.; Mitoraj, M.; Ziegler, T. Bond Orbitals from Chemical Valence Theory. J. Phys. Chem. A 2008, 112, 1933–1939. [Google Scholar] [CrossRef] [PubMed]
  38. Mitoraj, M.P.; Michalak, A.; Ziegler, T. A Combined Charge and Energy Decomposition Scheme for Bond Analysis. J. Chem. Theory Comput. 2009, 5, 962–975. [Google Scholar] [CrossRef] [PubMed]
  39. Mitoraj, M.P.; Michalak, A. σ-Donor and π-Acceptor Properties of Phosphorus Ligands: An Insight from the Natural Orbitals for Chemical Valence. Inorg. Chem. 2010, 49, 578–582. [Google Scholar] [CrossRef] [PubMed]
  40. In the Orca 5.0.4 implementation of EDA-NOCV, the electrostatic and Pauli terms cannot be separated.
  41. Armarego, W.L.F.; Chai, C.L.L. Purification of Laboratory Chemicals, 6th ed.; Elsevier: Burlington, MA, USA, 2009. [Google Scholar]
  42. Chalmers, B.A.; Arachchige, K.S.A.; Prentis, J.K.D.; Knight, F.R.; Kilian, P.; Slawin, A.M.Z.; Woollins, J.D. Sterically Encumbered Tin and Phosphorus peri-Substituted Acenaphthenes. Inorg. Chem. 2014, 53, 8795–8808. [Google Scholar] [CrossRef]
  43. Fuller, A.L.; Scott-Hayward, L.A.S.; Li, Y.; Bühl, M.; Slawin, A.M.Z.; Woollins, J.D. Automated Chemical Crystallography. J. Am. Chem. Soc. 2010, 132, 5799–5802. [Google Scholar] [CrossRef]
  44. CrystalClear-SM Expert; v2.0 and 2.1; Rigaku Americas: The Woodlands, TX, USA, 2010; Rigaku Corporation: Tokyo, Japan, 2015.
  45. CrysAlisPro; v1.171.42.49; Rigaku Oxford Diffraction, Rigaku Corporation: Oxford, UK, 2022.
  46. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef]
  47. Beurskens, P.T.; Beurskens, G.; de Gelder, R.; Garcia-Granda, S.; Gould, R.O.; Israel, R.; Smits, J.M.M. DIRDIF-99; Crystallography Laboratory, University of Nijmegen: Nijmegen, The Netherlands, 1999. [Google Scholar]
  48. Palatinus, L.; Chapuis, G. SUPERFLIP—A computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. J. Appl. Crystallogr. 2007, 40, 786–790. [Google Scholar] [CrossRef]
  49. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  50. CrystalStructure; v4.3.0; Rigaku Americas: The Woodlands, TX, USA; Rigaku Corporation: Tokyo, Japan, 2018.
  51. Neese, F. Software update: The ORCA program system—Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
  52. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  53. van Lenthe, E.; Snijders, J.G.; Baerends, E.J. The zero-order regular approximation for relativistic effects: The effect of spin–orbit coupling in closed shell molecules. J. Chem. Phys. 1996, 105, 6505–6516. [Google Scholar] [CrossRef]
  54. Weigend, F.; Furche, F.; Ahlrichs, R. Gaussian basis sets of quadruple zeta valence quality for atoms H–Kr. J. Chem. Phys. 2003, 119, 12753–12762. [Google Scholar] [CrossRef]
  55. Schäfer, A.; Horn, H.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets for atoms Li to Kr. J. Chem. Phys. 1992, 97, 2571–2577. [Google Scholar] [CrossRef]
  56. Schäfer, A.; Huber, C.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829–5835. [Google Scholar] [CrossRef]
  57. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  58. Pantazis, D.A.; Chen, X.-Y.; Landis, C.R.; Neese, F. All-Electron Scalar Relativistic Basis Sets for Third-Row Transition Metal Atoms. J. Chem. Theory Comput. 2008, 4, 908–919. [Google Scholar] [CrossRef] [PubMed]
  59. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  60. Pantazis, D.A.; Neese, F. All-Electron Scalar Relativistic Basis Sets for the Lanthanides. J. Chem. Theory Comput. 2009, 5, 2229–2238. [Google Scholar] [CrossRef] [PubMed]
  61. Pantazis, D.A.; Neese, F. All-Electron Scalar Relativistic Basis Sets for the Actinides. J. Chem. Theory Comput. 2011, 7, 677–684. [Google Scholar] [CrossRef]
  62. Pantazis, D.A.; Neese, F. All-electron scalar relativistic basis sets for the 6p elements. Theor. Chem. Acc. 2012, 131, 1292. [Google Scholar] [CrossRef]
  63. 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. [Google Scholar] [CrossRef]
  64. Goerigk, L.; Grimme, S. A thorough benchmark of density functional methods for general main group thermochemistry, kinetics, and noncovalent interactions. Phys. Chem. Chem. Phys. 2011, 13, 6670–6688. [Google Scholar] [CrossRef]
  65. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  66. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  67. Hanwell, M.D.; Curtis, D.E.; Lonie, D.C.; Vandermeersch, T.; Zurek, E.; Hutchison, G.R. Avogadro: An advanced semantic chemical editor, visualization, and analysis platform. J. Cheminform. 2012, 4, 17. [Google Scholar] [CrossRef]
  68. Chalmers, B.A.; Nejman, P.S.; Llewellyn, A.V.; Felaar, A.M.; Griffiths, B.L.; Portman, E.I.; Gordon, E.L.; Fan, K.J.H.; Woollins, J.D.; Buhl, M.; et al. A Study of Through-Space and Through-Bond J(PP) Coupling in a Rigid Nonsymmetrical Bis(phosphine) and Its Metal Complexes. Inorg. Chem. 2018, 57, 3387–3398. [Google Scholar] [CrossRef] [PubMed]
  69. Malkina, O.L.; Hierso, J.-C.; Malkin, V.G. Distinguishing “Through-Space” from “Through-Bonds” Contribution in Indirect Nuclear Spin–Spin Coupling: General Approaches Applied to Complex JPP and JPSe Scalar Couplings. J. Am. Chem. Soc. 2022, 144, 10768–10784. [Google Scholar] [CrossRef] [PubMed]
  70. Taylor, L.J.; Lawson, E.E.; Cordes, D.B.; Athukorala Arachchige, K.S.; Slawin, A.M.Z.; Chalmers, B.A.; Kilian, P. Synthesis and Structural Studies of Peri-Substituted Acenaphthenes with Tertiary Phosphine and Stibine Groups Dataset; University of St Andrews Research Portal: St Andrews, UK, 2024. [Google Scholar] [CrossRef]
Figure 1. Literature Group 15 peri-substituted species mentioned in the introduction.
Figure 1. Literature Group 15 peri-substituted species mentioned in the introduction.
Molecules 29 01841 g001
Scheme 1. Syntheses of the peri-substituted phosphine-stibines reported in this paper. Compounds highlighted in frames are newly synthesised here. Note: nbd = norbornadiene, cod = 1,5-cyclooctadiene.
Scheme 1. Syntheses of the peri-substituted phosphine-stibines reported in this paper. Compounds highlighted in frames are newly synthesised here. Note: nbd = norbornadiene, cod = 1,5-cyclooctadiene.
Molecules 29 01841 sch001
Figure 2. Molecular structures of 5, 8, 2.PtCl2, and 4.Mo(CO)4. Anisotropic displacement ellipsoids are plotted at the 50% probability level. Hydrogen atoms, solvent molecules and minor components of disorder are omitted for clarity.
Figure 2. Molecular structures of 5, 8, 2.PtCl2, and 4.Mo(CO)4. Anisotropic displacement ellipsoids are plotted at the 50% probability level. Hydrogen atoms, solvent molecules and minor components of disorder are omitted for clarity.
Molecules 29 01841 g002
Figure 3. Molecular structures of 5 and 8 with molecules aligned approximately along the acenaphthene plane to show the quasi-linear P∙∙∙Sb-C motifs. Hydrogen atoms and minor components of disorder are omitted for clarity.
Figure 3. Molecular structures of 5 and 8 with molecules aligned approximately along the acenaphthene plane to show the quasi-linear P∙∙∙Sb-C motifs. Hydrogen atoms and minor components of disorder are omitted for clarity.
Molecules 29 01841 g003
Figure 4. Deformation densities for 5 and 8 associated with nP→σ*(Sb-C) interaction. Charge flow is from the negative isosurface (red) to the positive isosurface (blue). All isosurfaces are plotted at 0.001 au.
Figure 4. Deformation densities for 5 and 8 associated with nP→σ*(Sb-C) interaction. Charge flow is from the negative isosurface (red) to the positive isosurface (blue). All isosurfaces are plotted at 0.001 au.
Molecules 29 01841 g004
Figure 5. NMR numbering scheme.
Figure 5. NMR numbering scheme.
Molecules 29 01841 g005
Table 1. Selected bond distances, displacements, angles and torsion angles for the phosphine-stibines and their metal complexes.
Table 1. Selected bond distances, displacements, angles and torsion angles for the phosphine-stibines and their metal complexes.
Compound582.PtCl2∙CH2Cl24.Mo(CO)4
peri-region distances (Å)
P9···Sb13.172(3)3.218(2)3.357(4)3.3762(16)
P9–M1--2.248(4)2.5432(16)
Sb1–M1--2.4570(10)2.7007(6)
peri-region bond angles (°)
P9···Sb1–C (quasi-linear)168.48(19) [a]167.14(16) [b]--
P9–M1–Sb1--90.93(9)80.10(4)
Splay [c]15.1(12)16.3(9)16(2)17.2(8)
Out-of-plane displacements (Å)
P90.256(6)0.202(5)0.509(13)0.571(6)
Sb10.064(6)0.213(5)0.788(13)0.406(6)
M1--0.428(17)1.007(7)
peri-region torsion angle (°)
P9–C9···C1–Sb16.7(3)11.2(3)30.7(7)24.0(3)
[a] for Sb-iPr carbon C19; [b] for carbon C13; [c] Splay angle = sum of the bay region angles—360.
Table 2. Selected properties of the electron density at bond critical points according to QTAIM analysis. ρ(r) and ∇2ρ(r) are given in standard atomic units.
Table 2. Selected properties of the electron density at bond critical points according to QTAIM analysis. ρ(r) and ∇2ρ(r) are given in standard atomic units.
CompoundBondρ(r)2ρ(r)Ei (kcal mol−1)BD|V(r)|/G(r)ε
5P9∙∙∙Sb10.02720.0349−4.6−0.1081.250.0446
C1–Sb10.10530.0620−34.9−0.4551.760.0901
C9–P90.1600−0.0631−87.1−0.9162.120.1394
Sb1–C130.11100.0652−38.0−0.4721.760.0275
Sb1–C190.10200.0278−30.9−0.4501.870.0330
8P9∙∙∙Sb10.02460.0332−4.0−0.0881.210.0536
C1–Sb10.10380.0623−34.2−0.4501.750.0933
C9–P90.1603−0.0637−87.3−0.9172.120.1128
Sb1–C130.10180.0401−31.7−0.4471.820.0412
Sb1–C170.10740.0347−34.1−0.4661.850.0212
Table 3. Energy decomposition analysis for compounds 5 and 8. All values are in kcal mol−1. ΔEint = ΔEsteric + ΔEorb + ΔEdisp.
Table 3. Energy decomposition analysis for compounds 5 and 8. All values are in kcal mol−1. ΔEint = ΔEsteric + ΔEorb + ΔEdisp.
CompoundΔEintΔEstericΔEorbΔEdisp
5−226.67−40.55−173.78−12.34
8−221.92−39.61−171.04−11.27
Table 4. Selected crystallographic data.
Table 4. Selected crystallographic data.
2.PtCl24.Mo(CO)458
formula C31H34Cl4PPtSbC29H30MoO4PSbC27H34PSbC6H40PSb
fw 896.24691.22511.29505.33
crystal descriptionColourless blockColourless prismYellow chipColourless chip
crystal size [mm3]0.09 × 0.06 × 0.030.17 × 0.15 × 0.040.10 × 0.08 × 0.060.12 × 0.10 × 0.03
temperature [K]125173173173
space group Pna21P 1 ¯ P21/nC2/c
a [Å]14.797(2)10.1706(3)9.751(3)27.3960(13)
b [Å]18.166(3)12.6048(5)13.591(2)8.6563(4)
c [Å]11.8407(19)12.8276(7)19.091(5)22.5385(14)
α [°] 76.520(14)
β [°] 67.547(11)102.261(8)109.617(6)
γ [°] 67.048(12)
vol [Å]33182.8(9)1392.63(19)2472.3(11)5034.7(5)
Z4248
ρ (calc) [g/cm3]1.8701.6481.3741.333
μ [mm−1]5.6261.5071.1891.167
F(000)172868810482096
reflections collected24,78024,15029,60828,282
independent reflections (Rint)6286 (0.0886)4970 (0.1133)4528 (0.0604)22,368 (0.0755)
parameters, restraints347, 1330, 0297, 45260, 0
GoF on F21.1061.0811.1090.720
R1 [I > 2σ(I)]0.05050.05700.06200.0523
wR2 (all data)0.09780.14940.13080.1151
largest diff. peak/hole [e/Å3]0.88, −1.302.41, −0.720.67, −0.582.03, −1.39
Flack parameter0.012(6)---
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

Taylor, L.J.; Lawson, E.E.; Cordes, D.B.; Athukorala Arachchige, K.S.; Slawin, A.M.Z.; Chalmers, B.A.; Kilian, P. Synthesis and Structural Studies of peri-Substituted Acenaphthenes with Tertiary Phosphine and Stibine Groups. Molecules 2024, 29, 1841. https://doi.org/10.3390/molecules29081841

AMA Style

Taylor LJ, Lawson EE, Cordes DB, Athukorala Arachchige KS, Slawin AMZ, Chalmers BA, Kilian P. Synthesis and Structural Studies of peri-Substituted Acenaphthenes with Tertiary Phosphine and Stibine Groups. Molecules. 2024; 29(8):1841. https://doi.org/10.3390/molecules29081841

Chicago/Turabian Style

Taylor, Laurence J., Emma E. Lawson, David B. Cordes, Kasun S. Athukorala Arachchige, Alexandra M. Z. Slawin, Brian A. Chalmers, and Petr Kilian. 2024. "Synthesis and Structural Studies of peri-Substituted Acenaphthenes with Tertiary Phosphine and Stibine Groups" Molecules 29, no. 8: 1841. https://doi.org/10.3390/molecules29081841

Article Metrics

Back to TopTop