Next Article in Journal
Room Temperature Ferromagnetic Properties of Ga14N16−nGd2Cn Monolayers: A First Principle Study
Previous Article in Journal
Impact of Erbium Doping in the Structural and Magnetic Properties of the Anisotropic and Frustrated SrYb2O4 Antiferromagnet
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Structural Studies of Complexes of Bis(pentafluorophenyl)mercury with Di(phosphane oxide) Ligands

1
IITB-Monash Academy, Indian Institute of Technology, Bombay 400076, India
2
School of Chemistry, Monash University, Melbourne 3800, Australia
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(3), 530; https://doi.org/10.3390/cryst13030530
Submission received: 21 February 2023 / Revised: 9 March 2023 / Accepted: 16 March 2023 / Published: 20 March 2023
(This article belongs to the Section Crystalline Metals and Alloys)

Abstract

:
The reaction of bis(pentafluorophenyl)mercury with the ligands bis(diphenylphosphano) methane P,P’-dioxide ({Ph2P(O)}2CH2) (1), bis{2-(N,N,N’N’-tetraethyldiaminophosphano) imidazol-1-yl} methane P,P’-dioxide ({2-PO(NEt2)2C3N2H2}2CH2) (2) and bis (2-diphenylphosphanophenyl) ether P,P’-dioxide ({2-PPh2(O)C6H4}2O) (3) afforded crystalline σ-donor complexes [{Hg(C6F5)2}2{Ph2P(O)}2CH2] (1Hg), [Hg(C6F5)2{2-PO(NEt2)2C3N2H2}2CH2]n (2Hg) and [Hg(C6F5)2{2-PPh2(O)C6H4}2O] (3Hg), respectively. The molecular structures of 1Hg, 2Hg and 3Hg show considerable differences. In complex 1Hg, a single bridging bidentate ligand connects two three-coordinate T-shape mercury atoms with a near linear C-Hg-C atom array. Complex 2Hg is a one-dimensional coordination polymer in which adjacent four-coordinate mercury atoms with a linear C-Hg-C atom array are linked by bridging bidentate O,O’- ligands, whilst in complex 3Hg a T-shape three-coordinate mercury atom is ligated by (3) in a monodentate fashion. The Hg-O bond lengths of complexes 1Hg, 2Hg and 3Hg differ substantially (range 2.5373(14)-2.966(3) Å) owing to structural and bonding differences.

1. Introduction

Di(aryl)mercury compounds, HgR2, have a stable C-Hg-C arrangement in their structures and show virtually no Lewis acidity and it is a major challenge to enhance the acceptor properties. However, the early success in the synthesis of complexes between bis(pentafluorophenyl) mercury and ligands such as 2,2-bipyridine (bpy) and 1,2-bis(diphenylphosphano)ethane (dppe) [1] showed that the lack of Lewis acidic character shown by diphenylmercury [1,2,3,4] can be overcome by fluorine substitution. However, neither methylpentafluorophenyl- nor pentafluorophenyl(phenyl)-mercury would form similar complexes [1]. Both [Hg(C6F5)2L] (L = bpy or dppe) and [Hg(C6F5)2phen] (phen = 1,10-phenanthroline, Chart 1, I) were later prepared by thermal decarboxylation of the corresponding pentafluorobenzoatomercury precursors, [Hg(O2CC6F5)2L] [5]. A wider range of [Hg(C6F5)2L] complexes, where for example L = methyl-substituted 2,2′-bipyridine and 1,10-phenanthroline, 2,2′-biquinoline, ethane-1,2-diamine (en), PPh3, PPh3O, together with [{Hg(C6F5)2}2L] L = (EPh2)2CH2 (E = P or As) were prepared (Chart 1, II), but a study of their molecular weights in benzene and chloroform by osmometry showed most were substantially dissociated in solution into Hg(C6F5)2 + L, with only phen and en complexes showing significant stability [6]. Determination of the molecular structure of {Hg(C6F5)2}2(AsPh2(CH2)AsPh2)} [7] showed the diarsane ligand bridging two three-coordinate (T-shape) mercury atoms, where the Hg-As distance is only 0.3 Å less than the sum of the van der Waals radii of arsenic and mercury with a conservative value of 1.73 Å for mercury (Chart 1, III) [8]. Naumann et al. showed that halide ions can coordinate to give species of the type [Hg(C6F5)2X] (X = Cl, Br and I) which were crystallized using bulky cations such as PPh4+ and [K(18-crown-6)]+ [9] (Chart 1, IV). These compounds are three- coordinate systems adopting a T-shape geometry. Later, the same group isolated trinuclear complexes [{Hg(C6F5)2}3X] (X = Cl, Br, I), also utilizing bulky cations (Chart 1, V) [10]. Besides these complexes with σ-donors, a few π-donor complexes with arenes [11,12] (e.g., Chart 1, VI, VII) and transition metal Schiff base derivatives [13,14] have been obtained. Complexes with metallophilic interactions, e.g., AuI-HgII [15,16] are also known. It is striking that despite a considerable amount of investigation, structural information on complexes of [Hg(C6F5)2] containing neutral σ donors is limited to the one complex [{Hg(C6F5)2}2(AsPh2CH2AsPh2)] (Chart 1, III).
In order to obtain crystalline coordination derivatives of bis(pentafluorophenyl)mercury, we have examined complexation with three bulky, potentially chelating, bis(phosphane oxide) ligands, namely bis (diphenylphosphano)methane P,P′-dioxide (1), bis(2-(N,N,N’N’-tetrtaethyldiaminophosphano)imidazol-1-yl}methane P,P’-dioxide (2) and bis(2-diphenylphosphanophenyl) ether P,P′-dioxide (3) (Chart 2).
We were encouraged by recent studies of phosphane-oxide coordination to several mercury, especially organomercury, acceptors (Chart 3) [17,18,19,20,21]. The choice of ligands was also influenced by the spear-like P=O donor, which reduces steric repulsion close to the metal compared to tetraaryldiphosphane or diarsane ligands (Chart 1, III), whilst still retaining more distant bulk to aid crystallization. The favorable polarity, P+—O of the phosphoryl group, and the possibility of a chelate effect was also considered. Representative complexes have now been synthesized and structurally characterized resulting in different types of ligation and considerable variation in the Hg-O bond lengths.

2. Results and Discussion

2.1. Synthesis of [{Hg(C6F5)2}2{Ph2P(O)}2CH2] (1Hg)

The bisphosphane oxide {Ph2P(O)}2CH2 (1) was synthesized by the oxidation of bis(diphenylphosphino)methane with H2O2 [22]. The reaction of 1 with Hg(C6F5)2 in a 1:2 mole ratio in acetonitrile resulted in the formation of the dimercury complex [{Hg(C6F5)2}2{Ph2P(O)}2CH2] 1Hg in which 1 acts as a bridge (Scheme 1).
Slow evaporation of an acetonitrile solution of complex 1Hg yielded a crop of colorless crystal blocks. The same reaction carried out on a 1:1 mole ratio also gave rise to 1Hg, indicating a preference for the bridging coordination of 1. The bulk composition of 1Hg was established by elemental analysis. The 19F{1H} and 31P{1H} NMR spectra are very similar to those of the reactants [22,23] (see ESI, Figures S1 and S2). The IR spectrum of 1Hg revealed a strong ν(P=O) band at 1181 cm−1 (see ESI, Figure S3), somewhat displaced to longer wavelengths from the absorptions of the free ligand (1212, 1197, 1176 cm−1) [24] as expected due to the coordination of a P=O group [25]. Of further interest is that the ‘X-sensitive’ mode involving C-Hg stretching of the HgC6F5 group is shifted from 812 cm−1 in Hg(C6F5)2 [5] to 794 cm−1 in the spectrum of complex 1Hg in a direction consistent with weakening the Hg-C bond on coordination. Carbon-fluorine stretching is observed at 1055 and 957 cm−1.

2.2. Molecular Structure of [{Hg(C6F5)2}2{Ph2P(O)}2CH2] (1Hg)

The compound 1Hg crystallizes in the monoclinic space group C2/c (Figure 1). The coordination geometry of the mercury atoms is a distorted T-shape, where the O-Hg-C angles are 86.56(7)° and 97.38(6)° (Figure 1). The C-Hg-C angle (175.06(9)°) is reduced slightly compared to that in Hg(C6F5)2 (179.1(5)o original rotamer; 177.10(11)° new rotamer) [26], as a result of O–Hg bonding. However, the Hg-C bond lengths ((2.063(2), 2.071(2) Å) are not significantly affected when compared with Hg(C6F5)2 (2.076(11), 2.066(11) Å original rotamer; 2.064(3), 2.065(3) Å new rotamer) [26]. The Hg-O bond length (2.5373(14) Å) is much shorter than the sum (3.23 Å) of the van der Waals radii of mercury and oxygen [8,27]. In addition, the Hg-O bond length of 1Hg is close to values observed for Hg–O in II (2.545 Å), III (2.624 (3) Å) and IV (2.510(2) Å) (Chart 3), where Hg–O bonding is assisted by chelation, while II and IV (Chart 3) have arylmercuric chlorides as the acceptors, which are normally stronger Lewis acids than diarylmercurials. The most convincing comparison is the similarity with Hg-O of II (Chart 3) which has the same coordination number as 1Hg. The stereochemistry of the mercury atom is explicable in terms of linear sp hybridization to provide the acceptor orbitals for the pentafluorophenyl groups leaving a weaker acceptor p orbital normal to the sp axis for the oxygen donor atom. The P = O and P-C bond lengths (Figure 1) are close to those (1.486(2) Å-and 1.815(2) Å, respectively) of the ligand 1 [28], hence coordination does not greatly affect the ligand structure. Examination of the supramolecular interactions in 1Hg show C-H····F-C bonds and displaced π ···π interactions along with intermolecular C···F contacts (see ESI, Figures S4–S6). The CH····FC interactions (2.471–2.496 Å) (Table 1) are less than 2.55 Å considered to be the limit of effectiveness for such interactions [29], and the offset π ···π interactions (3.384 Å), are shorter than the sum of the van der Waals radii of two aromatic rings (3.5 Å) [30]. The C···F contacts (3.007–3.64 Å) (see ESI, Table S1) lie near the sum of the van der Wall radii of C and F [27] and perhaps represent non-bonding/steric limits.

2.3. Synthesis of {2-PO(NEt2)2C3N2H2}2CH2 (2)

Bis{2-(N,N,N’N’-tetraethyldiaminophosphano)imidazol-1-yl}methane (2a) was prepared by a slight modification of the reported procedure [31,32]. Bis(imidazol-1-yl)methane was treated with two equivalents of nBuLi in tetrahydrofuran at −78 °C followed by the addition of two equivalents of P(NEt2)2Cl. The reaction mixture containing compound 2a was observed in the 31P{1H} NMR spectrum with a peak at 67.3 ppm (see ESI, Figure S7), and was oxidized to bis{2-(N,N,N’N’-tetraethyldiamino)phosphano)imidazol-1-yl}methane P,P’-dioxide 2 using H2O2 (Scheme 2), identified by 1H and 31P{1H} NMR spectroscopy and X-ray crystallography. The methylene protons of 2 appear at 8.05 ppm in the 1H NMR spectrum and are in the range similar to analogous reported di(phosphane) oxides (7.12–8.06 ppm, mainly overlapping with Ph-group resonances, which are not present in 2) [33]. The 31P{1H} NMR spectrum of 2 shows a single resonance at 16.0 ppm (see ESI, Figures S8 and S9) and the IR contains strong peaks at (1220–1202) cm−1 corresponding to the ν(P = O) band (see ESI, Figure S11).

2.4. Molecular Structure of {2-PO(NEt2)2C3N2H2}2CH2 (2)

Single crystals of 2 suitable for single crystal X-ray analysis were obtained from a hexane solution of 2 kept at −20 °C for 24 h. The phosphane oxide 2 crystallizes in the monoclinic space group P21/n (Figure 2). The phosphorus atoms have a distorted tetrahedral geometry with the bond angles ranging from 110.07 (10)°–119.81 (13)° (Figure 2). This spread of angles is much greater than observed (111.43(6)-113.10°) for an analogous ligand in which phenyl groups replace the NEt2 groups of 2, namely CH2(1,2-C3H2N2PPh2O)2 [33], presumably owing to the greater steric effect of the NEt2 moiety. However, the variation of angles is increased for the phenyl-substituted ligand if the imidazole moiety is replaced by a triazole unit [33]. The P = O bond distances are in the range 1.4773(11)–1.4783(11) Å, and are marginally shorter than in the phenyl-substituted analogue above (1.4854(10)–1.4881(10) Å [33]. This may seem counter-intuitive given that phenyl groups are electron withdrawing and NEt2 electron donating. The molecular structure of 2 further shows the presence of the following intermolecular interactions in ligand 2: C-H···C-H, C-H···O interactions and H···H contacts (Table 2; see also ESI, Figure S12).

2.5. Synthesis of [Hg(C6F5)2{2-PO(NEt2)2C3N2H2}2CH2]n (2Hg)

Treatment of phosphane oxide 2 with Hg(C6F5)2 in a 1:2 mole ratio (Scheme 3) followed by a slow evaporation of the solution yielded the 1D-coordination polymer 2Hg as colorless blocks.

2.6. Structure of [Hg(C6F5)2{2-PO(NEt2)2C3N2H2}2CH2]n (2Hg)

Complex 2Hg crystallizes in the triclinic space group P-1 and has a polymeric form (Figure 3), in which 2 acts as an O, O’ bridging bidentate ligand between adjacent mercury atoms. Each mercury is four-coordinate with linear C-Hg-C units adding two oxygen donor atoms approximately normal to the C-Hg-C spine. The donor atoms can be envisaged as occupying the axial and two adjacent equatorial positions of an octahedron, i.e., with two adjacent equatorial positions empty. The stereochemistry is essentially as expected for two sp hybrid orbitals on mercury accommodating the C6F5—lone pairs and two p orbitals used for the P = O lone pairs. The Hg-C bond lengths of Hg(C6F5)2 are essentially unchanged on coordination whilst the C-Hg-C angle has opened slightly to exactly linear from that of the free mercurial (179.1(5)° original rotamer; 177.10(11)° new rotamer) [26]. The Hg-O bond lengths (2.966(3), 2.979(4) Å) are ca. 0.025 Å within the sum of the van der Waals radii of Hg and O [8,27], but are much longer than in 1Hg, partly owing to the higher coordination number of mercury in 2Hg than 1Hg. These values are slightly longer than Hg-O (2.824(4)-2.895(4) Å) of the complex of (Me2N)3P = O with the cyclic trimeric mercurial complex [Hg(o-C6F4)]3 [18]. The P-O bond lengths are ~1.475(4)–1.479(4) Å with not much change with respect to the ligand 2.
Further examination of the molecular structure of 2Hg revealed C-H···N-C, C-H···F-C, C-H···C and C-H···N interactions along with intermolecular F-C····C and H···H contacts (Table 3; see also ESI, Figures S13–S17). The C-H···F-C interactions are between 2.373–2.438 Å, less than 2.55 Å which is considered to be the limit for such interactions to be significant [29] and the C-H···N-C interaction is 2.693 Å which is in the reported range [34] (see ESI, Table S3).

2.7. Synthesis of [Hg(C6F5)2{2-PPh2(O)C6H4}2O] (3Hg)

The ligand {2-PPh2(O)C6H5}2O (3) was synthesized as reported [35]. A reaction between 3 and Hg(C6F5)2 was carried out in a 1:1 mole ratio (Scheme 4). Slow evaporation of the solution yielded colorless blocks of complex [Hg(C6F5)2{2-PPh2(O)C6H4}2O] 3Hg. This outcome contrasts with the behavior of ligand 1, which even on a 1:1 stoichiometry, yielded complex 1Hg in a 2Hg(C6F5)2:1 ratio. It also contrasts with the behavior of ligand 2 which yields the polymeric complex 2Hg on this stoichiometry. The 19F{1H} and 31P{1H} NMR spectra of 3Hg are very similar to those of the reactants [23,36] (see ESI, Figures S18 and S19). The IR spectrum of 3Hg shows strong ν(P=O) bands at 1225 and 1165 cm−1, attributable to the free and coordinated P=O groups, respectively, where the latter absorption is slightly shifted to longer wavelength from that of the free ligand 3 at 1172 cm−1 (owing to the coordination of P=O group to mercury) and the former to higher frequencies from the free ligand band at 1216 cm−1 (see ESI, Figures S20 and S21). The ‘X-sensitive mode incorporating C-Hg stretching is located at 806 cm−1 and is only slightly shifted from 812 cm−1 of free Hg(C6F5)2 [5]. This shift is less than observed for 1Hg and may be the result of the longer (weaker) Hg-O bond in 3Hg compared to 1Hg. The carbon-fluorine stretching frequencies at 1069 and 964 cm−1 are in a similar region to those of 1Hg.

2.8. Molecular Structure of [Hg(C6F5)2{2-PPh2(O)C6H4}2O] (3Hg)

Compound 3Hg crystallizes in the triclinic space group P-1 (Figure 4). A T-shaped three coordination is observed at the mercury atom. Unlike ligands 1 and 2, 3 behaves as a monodentate ligand and shows no bridging of the Hg atoms. Only one of the phosphoryl groups is bonded to mercury. The C-Hg-O bond angles around the mercury atom 88.11(15)°–95.62(14)° indicate that the T-shape is distorted (Figure 4). The Hg-O bond length (2.711(3) Å) is well within the sum of the Hg and O van der Waals radii [8,27] and is longer than in the bridged bidentate complex 1Hg but considerably shorter than in the coordination polymer 2Hg. The P=O bond lengths range between 1.486(3) and 1.493(3) Å. The value for the coordinated P1-O1 is marginally longer than free P2-O3 (Figure 4), but the difference does not meet the three ESD criteria. Investigating the supramolecular interactions in 3Hg shows C-H···F-H, C-H···C-H and C-H···N interactions along with intermolecular F-C····C- and H···H contacts (Table 4; see also ESI Figures S22 and S23). The C-H···F-C interactions are in the range ~2.407–2.593 Å with some slightly longer than 2.55 Å [29] (see ESI, Table S4) while the P=O···C contact is 3.191 Å. For all complexes 1Hg, 2Hg and 3Hg the presence of Hg-C bonds leads to Hg···o-F, and Hg···o-C contacts (Table S5) which lie close to the sum of the van der Waals radii [8,27,30].

3. Conclusions

Complexes of bis(pentafluorophenyl)mercury were prepared with three different bis(phosphane) oxides. Each bisphosphane oxide has a different mode of coordination. Ligand {Ph2P(O)}2CH2 1 forms complex [{Hg(C6F5)2}2{Ph2P(O)}2CH2] 1Hg with a single bridging bidentate ligand, whereas ligand {2-PO(NEt2)2C3N2H2}2CH2 2 affords the 1D-coordination polymer [Hg(C6F5)2{2-PO(NEt2)2C3N2H2}2CH2] 2Hg where each mercury is bound by two bridging bidentate ligands. Further, the bisphosphane oxide (2-PPh2(O)C6H5}2O 3 coordinates to Hg(C6F5)2 in a monodentate fashion leading to the formation of [Hg(C6F5)2{2-PPh2(O)C6H4}2O)] 3Hg, a T-shaped monomeric coordination complex. This study on the coordination chemistry of Hg(C6F5)2 reveals there is more to be discovered in the binding of neutral σ-donors to organomercurials, if they are suitably substituted to enhance their acceptor properties.

4. Materials and Methods

4.1. General Considerations

Bis(diphenylphosphano)methane P,P’-dioxide (1) [22], bis(2-diphenylphosphanophenyl)ether P,P’-dioxide (3) [35], bis(imidazol-1yl)methane [37], P(NEt2)2Cl [38] and [Hg(C6F5)2] [23] were synthesized by literature procedures. Room temperature (25 °C) 1H, 31P{1H} and 13C NMR spectra were recorded with a Bruker DPX 300 instrument using CDCl3 or CD3COCD3 as solvents and resonances were referenced to residual hydrogen-atom or carbon-atom of the deuterated solvent. 31P{1H} NMR spectra are measured with 85% H3PO4 as an external standard. Chemical shifts for 19F{1H} were referenced externally to trifluorotoluene. Infrared spectra were recorded using a PerkinElmer Spectrum One FT-IR spectrometer (Model no. 73465) in a KBr disk and ATR-Infrared spectra with a PerkinElmer 1600 FT-IR spectrometer from 4000 to 450 cm-1. Elemental analyses (C, H, N) were performed by the School of Chemistry, Monash University.

4.2. Single Crystal X-ray Structure Determination

Crystals for X-ray structure analysis were grown using saturated solutions in hexane (2), acetonitrile (1Hg), acetonitrile: dichloromethane (2Hg) or chloroform (3Hg). Crystals 2, 1Hg, 2Hg and 3Hg were immersed in Paratone, and were measured on a Rigaku Saturn724 diffractometer (2), a Rigaku SynergyS diffractometer (1Hg) and the MX1beamlines at the Australian Synchrotron (2Hg and 3Hg). The Saturn724 was operated using microsource Mo Kα radiation (λα = 0.71073 Å) at 150 K, the SynergyS was operated using microsource Mo Kα radiation (λα = 0.71073 Å) at 123 K, and the MX1 beamline was operated using a single wavelength(λ = 0.71073 Å) at 100K. Data processing was conducted using the CrysAlisPro.55 software suite [39]. Structural solutions were obtained by ShelXT [40] and refined using full-matrix least-squares methods against F2 using SHELXL [41], in conjunction with the Olex2 [42] graphical user interface. All hydrogen atoms were placed in calculated positions using the riding model. Crystal and refinement data are given in Table 5.

4.3. Experimental Section

4.3.1. Synthesis of {PO(NEt2)2C3N2H2}2CH2 (2)

To a solution of bis(imidazol-1-yl)methane (0.507 g, 3.42 mmol) dissolved in dry THF (50 mL) was added n BuLi (1.6 M solution in hexane, 7.5 mmol, 4.6 mL) dropwise at −78 °C under nitrogen using Schlenk-line techniques and the reaction mixture was slowly warmed to room temperature followed by stirring for 2 h. P(NEt2)2Cl (1.570 g, 7.5 mmol) in THF (20 mL) was added dropwise at −78 °C and the reaction mixture was slowly allowed to reach room temperature and was further stirred for 12 h. The solvent was removed under vacuum and the resulting crude diphosphane was given an aqueous work-up and extracted with dichloromethane. The crude diphosphane was dissolved in THF and 30% H2O2 (5.9 mmol, 0.70 mL) was added to the solution of diphosphane at 0 °C and stirred for 1 h. The reaction mixture was evaporated under reduced pressure giving a yellow-colored viscous liquid. The viscous liquid obtained was dissolved in a minimum amount of hexane and kept at −20 °C for 24 h to yield colorless blocks of 2. Yield = 362 mg (20%) FT-IR (KBr disc cm−1) 3104 (s, br), 2975 (vs), 2942 (m), 2879 (s), 1743 (s), 1669 (s, vbr), 1514 (w), 1471 (s), 1388 (s), 1363 (s), 1285 (s, br), 1262 (s, br), 1220 (m, br), 1212 (s, br), 1202 (m, br), 1110 (s), 1062 (s), 1025 (vs), 951 (vs), 936 (m), 792 (vs), 769 (w), 711 (s), 691 (s), 666 (s), 562 (vs, br), 552 (m, br), 531(m, br). 1H NMR (500 MHz, CDCl3) δ 8.05 (s, 2H), 7.16 (d, J = 2.4 Hz, 2H), 7.10 (s, 2H), 3.27–3.03 (m, 16H), 1.04 (s, 24H). 31P {1H} NMR (202 MHz, CDCl3) δ 15.95. 13C NMR (126 MHz, CDCl3) δ 141.32, 139.68, 130.06, 129.92, 124.00, 123.97, 54.86, 38.60, 38.56.

4.3.2. Synthesis of [{Hg(C6F5)2}2(Ph2P(O))2CH2] (1Hg)

Hg(C6F5)2 (49.2 mg, 0.092 mmol) and 1 (19.2 mg, 0.046 mmol) were stirred together in acetonitrile (10 mL). Slow evaporation of acetonitrile solution yielded colorless blocks of 1Hg. Yield 65 mg (95%), 31P{1H} NMR (162 MHz, d6-Acetone) δ 23.13. 19F{1H} NMR (377 MHz, d6-Acetone) δ −119.9 (3JF,Hg = 449 Hz, 4F), −155.7 (2F), −162.2 (4F, 4JF,Hg = 136 Hz). FT-IR (ATR cm−1):2923 (w, br), 2362 (w, br), 1638 (s,w), 1506 (s), 1473 (vs), 1435 (s), 1368 (s), 1270 (w), 1181 (vs), 1119 (s), 1100 (w), 1055 (s), 999 (w, br), 957 (vs), 794 (s), 744 (m),730 (s), 691 (vs). Elemental analysis Calcd (%) for C49H22F20Hg O2P2: C 39.61, H 1.49; found C 39.35, H 1.32.

4.3.3. Synthesis of [Hg(C6F5)2{2-PO(NEt2)2C3N2H2}2CH2]n (2Hg)

Hg(C6F5)2 (49.2 mg, 0.092 mmol) and 2 (25 mg, 0.046 mmol) were stirred together in CHCl3 (5 mL). The slow evaporation of the chloroform solution yielded colorless blocks of 2Hg in an amount sufficient for structure determination. Half of the Hg(C6F5))2 remained unreacted.

4.3.4. Synthesis of [Hg(C6F5)2{PPh2(O)C6H4}2O] (3Hg)

Hg(C6F5)2 (24.6mg, 0.046 mmol) and 3 (26.25 mg, 0.046 mmol) were stirred together in CHCl3 (5 mL). Slow evaporation of chloroform solution yielded colorless blocks of 3Hg. Yield = 45 mg (88%). 31P{1H} NMR (162 MHz, CDCl3) δ 25.93. 19F{1H} NMR (377 MHz, CDCl3) δ −119.43 (4F, 3JF,Hg = 417 Hz,), −150.34 (2F), −158.79 (4F, 4JF,Hg = 122 Hz).FT-IR (KBr disc cm−1):Overall broad feature at 3058 (s), 2930(m), and 2860(w),1642 (vs), 1596 (vs), 1567 (vs), 1512 (s), 1478 (vs), 1439 (vs, br), 1310 (m), 1272 (m), 1271 (m), 1225 (vs), 1203 (m), 1186 (w), 1165 (vs), 1133 (w), 1122 (vs), 1081 (vs), 1069 (vs), 997 (m), 964 (vs), 878 (s), 806 (vs), 759 (s), 731 (s), 697 (s), 610 (m), 584 (m), 539 (vs), 518 (vs), 490 (w).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst13030530/s1, including NMR spectra, additional experimental, X-ray and refinement data, bond lengths and angle data.

Author Contributions

Conceptualization, G.B.D. and V.L.B.; synthesis, spectroscopy, characterization, original draft, S.R.; X-ray crystallography, O.A.B. and Z.G.; supervision and editing, M.S.B.; supervision, rewriting, editing, G.B.D. and V.L.B. All authors have read and agreed to the published version of the manuscript.

Funding

GBD and VLB gratefully acknowledges the Australian Research Council (ARC) for funding, DP190100798 and FT200100918, respectively.

Data Availability Statement

Crystal data can be obtained free of charge from The Cambridge Crystallographic Data Centre: CCDC 2243139–2243142 for compounds 2, 1Hg, 2Hg and 3Hg, respectively.

Acknowledgments

We are grateful to the Faculty of Science, Monash University and the Department of Chemistry, the Indian Institute of Technology, Bombay, for providing the required facility for carrying out the work. SR gratefully acknowledges the Indian Institute of Technology Bombay-Monash Academy for the financial support. We thank Craig Forsyth and Dipanjan Mondal for X-ray crystallography assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chambers, R.D.; Coates, G.E.; Livingstone, J.G.; Musgrave, W.K.R. 847. The preparation and study of some pentafluorophenylmercury compounds. J. Chem. Soc. 1962, 4367–4371. [Google Scholar] [CrossRef]
  2. Lin, T.-P.; Nelson, R.C.; Wu, T.; Miller, J.T.; Gabbaï, F.P. Lewis acid enhancement by juxtaposition with an onium ion: The case of a mercury stibonium complex. Chem. Sci. 2012, 3, 1128–1136. [Google Scholar] [CrossRef]
  3. Canty, A.; Gatehouse, B. The crystal structures of disordered crystals of adducts of diphenylmercury with bidentate ligands. Acta Crystallogr. B 1972, 28, 1872–1888. [Google Scholar] [CrossRef]
  4. Korpar-Cǒlig, B.; Popović, Z.; Bruvo, M.; Vicković, I. Addition compounds of bis(trifluoromethyl)mercury with tetraphenylphosphonium and tetraphenylarsonium halides and thiocyanates. Inorg. Chim. Acta 1988, 150, 113–118. [Google Scholar] [CrossRef]
  5. Connett, J.; Davies, A.G.; Deacon, G.; Green, J. Organomercury compounds. Part I. Mercury pentafluorobenzoates and their decarboxylation to pentafluorophenylmercury compounds. J. Chem. Soc. C 1966, 106–111. [Google Scholar] [CrossRef]
  6. Canty, A.; Deacon, G. Organomercury compounds. XI. Some coordination derivatives of bispentafluorophenylmercury. Aust. J. Chem. 1971, 24, 489–499. [Google Scholar] [CrossRef]
  7. Canty, A.J.; Gatehouse, B.M. Crystal structure of tetrakispentafluorophenyl-µ-bisdiphenylarsinomethane-dimercury(II), a three-co-ordinate mercury complex. J. Chem. Soc. D 1971, 443–444. [Google Scholar] [CrossRef]
  8. Canty, A.J.; Deacon, G.B. The van der Waals radius of mercury. Inorg. Chim. Acta 1980, 45, L225–L227. [Google Scholar] [CrossRef]
  9. Schulz, F.; Pantenburg, I.; Naumann, D. Synthesen und Strukturen von Perfluororganohalogenomercuraten [Hg(Rf)2X]—(Rf = CF3, C6F5; X = Br, I). Z. Anorg. Allg. Chem. 2003, 629, 2312–2316. [Google Scholar] [CrossRef]
  10. Naumann, D.; Schulz, F. Strukturen von neuen Bis(pentafluorophenyl)halogenomercuraten [{Hg(C6F5)2}3(μ-X)]—(X = Cl, Br, I). Z. Anorg. Allg. Chem. 2005, 631, 715–718. [Google Scholar] [CrossRef]
  11. Burress, C.N.; Bodine, M.I.; Elbjeirami, O.; Reibenspies, J.H.; Omary, M.A.; Gabbaï, F.P. Enhancement of external spin−orbit coupling effects caused by metal−metal cooperativity. Inorg. Chem. 2007, 46, 1388–1395. [Google Scholar] [CrossRef] [PubMed]
  12. Deacon, G.B.; Forsyth, C.M.; Junk, P.C.; Ness, T.J.; Izgorodina, E.; Baldamus, J.; Meyer, G.; Pantenburg, I.; Hitzbleck, J.; Ruhlandt-Senge, K. The Supramolecular Architecture of Arene Complexes of Bis(polyfluorophenyl)mercurials. Eur. J. Inorg. Chem. 2008, 2008, 4770–4780. [Google Scholar] [CrossRef]
  13. Kim, M.; Taylor, T.J.; Gabbaï, F.P. Hg(II)···Pd(II) metallophilic interactions. J. Am. Chem. Soc. 2008, 130, 6332–6333. [Google Scholar] [CrossRef]
  14. Tsunoda, M.; Fleischmann, M.; Jones, J.S.; Bhuvanesh, N.; Scheer, M.; Gabbaï, F.P. Supramolecular aggregation of Ni(salen) with (C6F5)2Hg and [o-C6F4Hg]3. Dalton Trans. 2016, 45, 5045–5051. [Google Scholar] [CrossRef]
  15. López-de-Luzuriaga, J.M.; Monge, M.; Olmos, M.E.; Pascual, D.; Lasanta, T. Amalgamating at the molecular level. A study of the strong closed-shell Au(i)⋯Hg(ii) interaction. Chem. Commun. 2011, 47, 6795–6797. [Google Scholar] [CrossRef] [PubMed]
  16. Lasanta, T.; López-de-Luzuriaga, J.M.; Monge, M.; Olmos, M.E.; Pascual, D. Experimental and Theoretical Evidence of the Existence of Gold(I)⋅⋅⋅Mercury(II) Interactions in Solution through Fluorescence-Quenching Measurements. Chem. Eur. J. 2013, 19, 4754–4766. [Google Scholar] [CrossRef]
  17. Rosario-Amorin, D.; Dehaudt, J.P.; Caudle, L.J.; Dickie, D.A.; Paine, R.T. Synthesis and molecular structures of mercury(II) complexes of carbamoylmethylphosphoryl ligands. Phosphorus Sulfur Silicon Relat. Elem. 2016, 191, 520–526. [Google Scholar] [CrossRef]
  18. Tikhonova, I.A.; Dolgushin, F.M.; Tugashov, K.I.; Petrovskii, P.V.; Furin, G.G.; Shur, V.B. Coordination chemistry of polymercuramacrocycles. Complexation of cyclic trimeric perfluoro-o-phenylenemercury with neutral oxygeneous Lewis bases. J. Organomet. Chem. 2002, 654, 123–131. [Google Scholar] [CrossRef]
  19. Olaru, M.; Krupke, S.; Lork, E.; Mebs, S.; Beckmann, J. Transmetallation of bis(6-diphenylphosphinoxy-acenapth-5-yl)mercury with tin tetrachloride, antimony trichloride and bismuth trichloride. Dalton Trans. 2019, 48, 5585–5594. [Google Scholar] [CrossRef]
  20. Cole, M.L.; Deacon, G.B.; Junk, P.C.; Konstas, K.; Roesky, P.W. Novel Mercury Phosphanylamides. Eur. J. Inorg. Chem. 2005, 2005, 1090–1098. [Google Scholar] [CrossRef]
  21. Padron, D.A.; Klausmeyer, K.K. Syntheses and coordination studies of 2-(diphenylphosphinomethyl)pyridine and its oxide towards mercury(II). Polyhedron 2012, 34, 215–220. [Google Scholar] [CrossRef]
  22. Aslandukov, A.N.; Utochnikova, V.V.; Goriachiy, D.O.; Vashchenko, A.A.; Tsymbarenko, D.M.; Hoffmann, M.; Pietraszkiewicz, M.; Kuzmina, N.P. The development of a new approach toward lanthanide-based OLED fabrication: New host materials for Tb-based emitters. Dalton Trans. 2018, 47, 16350–16357. [Google Scholar] [CrossRef]
  23. Deacon, G.B.; Cosgriff, J.E.; Lawrenz, E.T.; Forsyth, C.M.; Wilkinson, D.L. Section 2.3: Organolanthanide(II) Complexes in Lanthanides and Actinides. In Synthetic Methods of Organometallic and Inorganic Chemistry; Edelmann, F.T., Herrmann, W.A., Eds.; Georg Thieme Verlag: Stuttgart, Germany; New York, NY, USA, 1997; Volume 6, p. 48. [Google Scholar]
  24. Rangarajan, S.; Beaumont, O.A.; Balakrishna, M.S.; Deacon, G.B.; Blair, V.L. Synthesis and Characterisation of Novel Bis(diphenylphosphane oxide)methanidoytterbium(III) Complexes. Molecules 2022, 27, 7704. [Google Scholar] [CrossRef] [PubMed]
  25. Deacon, G.B.; Green, J.H.S. Vibrational spectra of ligands and complexes—VI: Infrared spectra of triphenylarsine, triphenylarsine oxide, and triphenylphosphine oxide complexes. Spectrochim. Acta A Mol. Spectrosc. 1969, 25, 355–364. [Google Scholar] [CrossRef]
  26. Guo, Z.; Huo, R.; Tan, Y.Q.; Flosbach, N.T.; Wang, N.; Leonhardt, C.; Urbatsch, A.; Deacon, G.B.; Junk, P.C.; Izgorodina, E.I.; et al. A new twist on an old molecule: A rotameric isomer of bis(pentafluorophenyl)mercury. J. Coord. Chem. 2021, 74, 2947–2958. [Google Scholar] [CrossRef]
  27. Nyburg, S.C.; Faerman, C.H. A revision of van der Waals atomic radii for molecular crystals: N, O, F, S, Cl, Se, Br and I bonded to carbon. Acta Crystallogr. B Struct. 1985, 41, 274–279. [Google Scholar] [CrossRef]
  28. Grachova, E.; Linti, G.; Vologzhanina, A. CCDC 877465: EBIXEV Experimental Crystal Structure Determination. In CSD Communications; University of Cambridge: Cambridge, UK, 2016. [Google Scholar]
  29. Dunitz, J.D.; Taylor, R. Organic Fluorine Hardly Ever Accepts Hydrogen Bonds. Chem. Eur. J. 1997, 3, 89–98. [Google Scholar] [CrossRef]
  30. Bondi, A. van der Waals volumes and radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  31. Antiñolo, A.; Carrillo-Hermosilla, F.; Díez-Barra, E.; Fernández-Baeza, J.; Fernández-López, M.; Lara-Sánchez, A.; Moreno, A.; Otero, A.; Maria Rodriguez, A.; Tejeda, J. New functionalized bis(pyrazol-1-yl)methane ligands. Synthesis, spectroscopic characterization of early and late transition metal complexes containing a functionalized N,N or P,P-chelate bis(5-diphenylphosphinopyrazol-1-yl)methane ligand. J. Chem. Soc. Dalton Trans. 1998, 3737–3744. [Google Scholar] [CrossRef]
  32. Bhat, S.A.; Mague, J.T.; Balakrishna, M.S. Gold(i) complexes of bisphosphines with bis(azol-1-yl)methane backbone: Structure of a rare dinuclear gold(i) complex [(Au2Cl){CH2(1,2-C3H2N2PPh2)2}3Cl]. Dalton Trans. 2015, 44, 17696–17703. [Google Scholar] [CrossRef]
  33. Bhat, S.A.; Sonawane, S.C.; Mague, J.T.; Balakrishna, M.S. Synthesis and characterization of Mo(0) and W(0) complexes of bis(azol-1-yl)methane based bisphosphines. J. Coord. Chem. 2021, 74, 2253–2262. [Google Scholar] [CrossRef]
  34. Geiger, D.K.; DeStefano, M.R. Conformational differences and intermolecular C-H...N interactions in three polymorphs of a bis(pyridinyl)-substituted benzimidazole. Acta Cryst. C 2016, 72, 867–874. [Google Scholar] [CrossRef]
  35. Pietraszkiewicz, M.; Mal, S.; Pietraszkiewicz, O.; Górski, K.; Chelwani, N. Photoluminescent Tetrazolate-based Eu(III) Complexes: An Outstanding Impact of Aromatic Phosphine Oxide Co-ligands on the Photoluminescence Quantum Yields. Z Naturforsch. 2014, 69, 239–247. [Google Scholar] [CrossRef]
  36. Li, Y.; Lu, L.-Q.; Das, S.; Pisiewicz, S.; Junge, K.; Beller, M. Highly Chemoselective Metal-Free Reduction of Phosphine Oxides to Phosphines. J. Am. Chem. Soc. 2012, 134, 18325–18329. [Google Scholar] [CrossRef] [PubMed]
  37. Díez-Barra, E.; de la Hoz, A.; Sánchez-Migallón, A.; Juan, T. Phase Transfer Catalysis without Solvent. Synthesis of Bisazolylalkanes. Heterocycles 1992, 34, 1365–1373. [Google Scholar] [CrossRef]
  38. Qin, L.; Ren, X.; Lu, Y.; Li, Y.; Zhou, J. Intermolecular Mizoroki–Heck Reaction of Aliphatic Olefins with High Selectivity for Substitution at the Internal Position. Angew. Chem. Int. Ed. 2012, 51, 5915–5919. [Google Scholar] [CrossRef]
  39. Agilent Technologies Ltd. CrysAlisPRO, Version 39; Agilent Technologies Ltd.: Yarnton, UK, 2021. [Google Scholar]
  40. Sheldrick, G. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Sheldrick, G. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Chart 1. Selected complexes of bis(pentafluorophenyl)mercury(II) with σ- and π-donor moieties I [6], II [6], III [7], IV [9], V [10], VI [11] and VII [12].
Chart 1. Selected complexes of bis(pentafluorophenyl)mercury(II) with σ- and π-donor moieties I [6], II [6], III [7], IV [9], V [10], VI [11] and VII [12].
Crystals 13 00530 ch001
Chart 2. Bis(phosphane oxide) ligands used in this study.
Chart 2. Bis(phosphane oxide) ligands used in this study.
Crystals 13 00530 ch002
Chart 3. Selected mercurial complexes I [17], II [19], III [20] and IV [21] with different phosphane oxide groups.
Chart 3. Selected mercurial complexes I [17], II [19], III [20] and IV [21] with different phosphane oxide groups.
Crystals 13 00530 ch003
Scheme 1. Synthesis of [{Hg(C6F5)2}2{Ph2P(O)}2CH2] (1Hg).
Scheme 1. Synthesis of [{Hg(C6F5)2}2{Ph2P(O)}2CH2] (1Hg).
Crystals 13 00530 sch001
Figure 1. Molecular structure of 1Hg (a) asymmetric unit and (b) perspective view of the complex. All hydrogen atoms are omitted for clarity. Displacement ellipsoids are drawn at the 30% probability level. Selected bond lengths (Å) and bond angles (o): Hg1-O1 2.5373(14), P1-O1 1.4895(15) P1-C13 1.8121(15), P1-C14 1.810(2), P1-C20 1.796(2). Hg1-C1 2.063(2), Hg1-C7 2.071(2), C1-Hg-C7 175.06 (9), O1-Hg1-C7 86.56(7), O1-Hg1-C1 97.38(6), P1-C13-P1′ 120.42(16).
Figure 1. Molecular structure of 1Hg (a) asymmetric unit and (b) perspective view of the complex. All hydrogen atoms are omitted for clarity. Displacement ellipsoids are drawn at the 30% probability level. Selected bond lengths (Å) and bond angles (o): Hg1-O1 2.5373(14), P1-O1 1.4895(15) P1-C13 1.8121(15), P1-C14 1.810(2), P1-C20 1.796(2). Hg1-C1 2.063(2), Hg1-C7 2.071(2), C1-Hg-C7 175.06 (9), O1-Hg1-C7 86.56(7), O1-Hg1-C1 97.38(6), P1-C13-P1′ 120.42(16).
Crystals 13 00530 g001
Scheme 2. Synthesis of ligand 2.
Scheme 2. Synthesis of ligand 2.
Crystals 13 00530 sch002
Figure 2. Molecular structure of 2. All hydrogen atoms except CH and lattice acetonitrile are omitted for clarity. Displacement ellipsoids are drawn at the 30% probability level. Selected bond lengths (Å) and bond angles (°): P1-O1 1.4773(11), P2-O2 1.4783(11), P1-N3 1.6373(13), P1-N4 1.6407(11), P2-N7 1.6444(15), P2-N8 1.6339(11), P1-C2 1.8056(12), P2-C13 1.8025(12), C1-N1 1.4589(16), C1-N5 1.4618(16). N5-C1-N1 112.10(11), O1-P1-N3 117.86(7), O1-P1-N4 110.42(6), O1-P1-C2 110.31(6), O2-P2-N8 110.79(6), O2-P2-N7 119.71(8), O2-P2-C13 110.07(6).
Figure 2. Molecular structure of 2. All hydrogen atoms except CH and lattice acetonitrile are omitted for clarity. Displacement ellipsoids are drawn at the 30% probability level. Selected bond lengths (Å) and bond angles (°): P1-O1 1.4773(11), P2-O2 1.4783(11), P1-N3 1.6373(13), P1-N4 1.6407(11), P2-N7 1.6444(15), P2-N8 1.6339(11), P1-C2 1.8056(12), P2-C13 1.8025(12), C1-N1 1.4589(16), C1-N5 1.4618(16). N5-C1-N1 112.10(11), O1-P1-N3 117.86(7), O1-P1-N4 110.42(6), O1-P1-C2 110.31(6), O2-P2-N8 110.79(6), O2-P2-N7 119.71(8), O2-P2-C13 110.07(6).
Crystals 13 00530 g002
Scheme 3. Synthesis of complex 2Hg.
Scheme 3. Synthesis of complex 2Hg.
Crystals 13 00530 sch003
Figure 3. Molecular structure of 2Hg (a) asymmetric unit and (b) perspective view of 1D polymer. Selected bond lengths (Å) and bond angles (°): Hg1-O1 2.966(3), Hg2-O2 2.979(4), P1-O1 1.475(3), P2-O2 1.479(4), Hg1-C29 2.061(5), Hg2-C30 2.055(5), C29-Hg1-O1 92.77(14), C30-Hg2-O2 82.05(18), C30′-Hg2-O2 97.95(18), C29′-Hg1-C29 180.0, C30′-Hg2-C30 180.0.
Figure 3. Molecular structure of 2Hg (a) asymmetric unit and (b) perspective view of 1D polymer. Selected bond lengths (Å) and bond angles (°): Hg1-O1 2.966(3), Hg2-O2 2.979(4), P1-O1 1.475(3), P2-O2 1.479(4), Hg1-C29 2.061(5), Hg2-C30 2.055(5), C29-Hg1-O1 92.77(14), C30-Hg2-O2 82.05(18), C30′-Hg2-O2 97.95(18), C29′-Hg1-C29 180.0, C30′-Hg2-C30 180.0.
Crystals 13 00530 g003
Scheme 4. Synthesis of [Hg(C6F5)2{2-PPh2(O)C6H4}2O] (3Hg).
Scheme 4. Synthesis of [Hg(C6F5)2{2-PPh2(O)C6H4}2O] (3Hg).
Crystals 13 00530 sch004
Figure 4. Molecular structure of 3Hg. All hydrogen atoms are omitted for clarity. Displacement ellipsoids are drawn at the 30% probability level. Selected bond lengths (Å) and bond angles (°): Hg1-O1 2.711(3), Hg1-C1 2.070(5), Hg1-C7 2.064(5), P1-O1 1.493(3), P2-O3 1.486(3). C1-Hg1-O1 88.11(15), C7-Hg1-O1 95.62(14), C7-Hg1-C1 175.15(17).
Figure 4. Molecular structure of 3Hg. All hydrogen atoms are omitted for clarity. Displacement ellipsoids are drawn at the 30% probability level. Selected bond lengths (Å) and bond angles (°): Hg1-O1 2.711(3), Hg1-C1 2.070(5), Hg1-C7 2.064(5), P1-O1 1.493(3), P2-O3 1.486(3). C1-Hg1-O1 88.11(15), C7-Hg1-O1 95.62(14), C7-Hg1-C1 175.15(17).
Crystals 13 00530 g004
Table 1. Hydrogen-bond geometry of 1Hg (Å, °).
Table 1. Hydrogen-bond geometry of 1Hg (Å, °).
D–H‧‧‧AD–HH‧‧‧AD‧‧‧AD–H‧‧‧A
C15–H15‧‧‧F6 [i]0.962.4963.164127
C13-H13···F3 [ii]0.952.4713.398176
[i]-x,y,1.5-z and [ii] 1-x,y,1.5-z.
Table 2. Hydrogen-bonds geometry (Å, °) of 2.
Table 2. Hydrogen-bonds geometry (Å, °) of 2.
D–H···AD–HH···AD···AD–H···A
C22–H22B‧‧‧C3 [i]0.972.7893.651148
C1-H1B···O2 [ii]0.972.6423.412137
[i] x,y,z and [ii] x,y,z.
Table 3. Hydrogen-bond geometry (Å, °) of 2Hg. Cg1 is C30-C35.
Table 3. Hydrogen-bond geometry (Å, °) of 2Hg. Cg1 is C30-C35.
D–H···AD–HH···AD···AD–H···A
C43–H43A‧‧‧F1 [i]0.962.4383.365162
C23-H23B···F2 [ii]0.962.3733.261154
C10-H10B···Cg1 [iii]0.973.0433.831139.3
C22A-H22A···N6 [iv]0.972.6933.492140
[i] x,y,z [ii] x,y,z [iii] 1-x,1-y,1-z [iv] x,y,z.
Table 4. Hydrogen-bonds geometry (Å, °) of 3Hg.
Table 4. Hydrogen-bonds geometry (Å, °) of 3Hg.
D–H‧‧‧AD–HH‧‧‧AD‧‧‧AD–H‧‧‧A
C24–H24‧‧‧F4 [i]0.932.4833.232138
C17-H17···F10 [ii]0.932.5933.423149
C35-H35···F9 [iii]0.932.4073.283157
C45-H45···F9 [iv]0.932.5613.151122
[i] 1-x,-y,2-z [ii] x,y,z [iii] x,y,z, [iv] x,y,z.
Table 5. Crystallographic Data.
Table 5. Crystallographic Data.
Compound21Hg2Hg3Hg
Empirical formulaC23H46N8O2P2[C49H22F20Hg2O2P2][C35H46F10HgN8O2P2][C48H28F10HgO3P2]
Formula weight528.621485.781063.331105.23
Temperature/K150123100100
Crystal systemMonoclinicMonoclinicTriclinicTriclinic
Space groupP21/nC2/cP-1P-1
a/Å9.3556(3)14.0695(2)11.480(2)8.510(17)
b/Å12.0193(4)17.0277(3)12.900(3)11.680(2)
c/Å25.8957(8)19.1180(3)15.330(3)21.020(4)
α/°9090110.03(3)84.47(3)
β/°94.663(3)92.6470(10)97.81(3)85.62(3)
γ/°909095.36(3)83.13(3)
Volume/Å32902.28(16)4575.24(13)2089.2(8)2060.2(7)
Z4422
ρcalcg/cm31.2102.1571.6901.782
μ/mm−10.1846.8973.8473.903
F(000)1144.02808.01056.01080.0
Mo Kα radiation/Synchrotron, λ/ÅMo Kα (λ = 0.71073)Mo Kα (λ = 0.71073)Synchrotron(λ = 0.71073)Mo Kα (λ = 0.71073)
Crystal size/mm30.258 × 0.115 × 0.0850.076 × 0.064 ×0.030.089 × 0.068 × 0.0560.1 × 0.1 × 0.08
2θ range for datacollection/°4.764 to 67.2827.038 to 63.7822.874 to 50.6961.95 to 58.188
Reflections collected32458294043744541307
Independent reflections9353654875728939
Data/restraints/parameters9353/27/3826548/0/3837572/228/7118942/0/577
Goodness-of-fit on F21.0181.0351.0681.060
R1 [a]0.04810.01930.03720.0376
wR2 [b]0.12870.04340.09650.1078
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

Rangarajan, S.; Beaumont, O.A.; Guo, Z.; Balakrishna, M.S.; Deacon, G.B.; Blair, V.L. Synthesis and Structural Studies of Complexes of Bis(pentafluorophenyl)mercury with Di(phosphane oxide) Ligands. Crystals 2023, 13, 530. https://doi.org/10.3390/cryst13030530

AMA Style

Rangarajan S, Beaumont OA, Guo Z, Balakrishna MS, Deacon GB, Blair VL. Synthesis and Structural Studies of Complexes of Bis(pentafluorophenyl)mercury with Di(phosphane oxide) Ligands. Crystals. 2023; 13(3):530. https://doi.org/10.3390/cryst13030530

Chicago/Turabian Style

Rangarajan, Shalini, Owen A. Beaumont, Zhifang Guo, Maravanji S. Balakrishna, Glen B. Deacon, and Victoria L. Blair. 2023. "Synthesis and Structural Studies of Complexes of Bis(pentafluorophenyl)mercury with Di(phosphane oxide) Ligands" Crystals 13, no. 3: 530. https://doi.org/10.3390/cryst13030530

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