Next Article in Journal
Modeling the Catalyst Activation Step in a Metal–Ligand Radical Mechanism Based Water Oxidation System
Next Article in Special Issue
On the Aqueous Solution Behavior of C-Substituted 3,1,2-Ruthenadicarbadodecaboranes
Previous Article in Journal
Functionalizing NaGdF4:Yb,Er Upconverting Nanoparticles with Bone-Targeting Phosphonate Ligands: Imaging and In Vivo Biodistribution
Previous Article in Special Issue
Dimethyloxonium and Methoxy Derivatives of nido-Carborane and Metal Complexes Thereof
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparing the Acidity of (R3P)2BH-Based Donor Groups in Iridium Pincer Complexes

Department of Chemistry, Philipps-Universität Marburg, Hans-Meerwein-Str. 4, 35032 Marburg, Germany
*
Author to whom correspondence should be addressed.
Submission received: 31 March 2019 / Revised: 28 April 2019 / Accepted: 29 April 2019 / Published: 7 May 2019
(This article belongs to the Special Issue Metal Complexes Containing Boron Based Ligands)

Abstract

:
In the current manuscript, we describe the reactivity of a series of iridium(III) pincer complexes with the general formulae [(PEP)IrCl(CO)(H)]n (n = +1, +2) towards base, where PEP is a pincer-type ligand with different central donor groups, and E is the ligating atom of this group (E = B, C, N). The donor groups encompass a secondary amine, a phosphine-stabilised borylene and a protonated carbodiphosphorane. As all ligating atoms E exhibit an E–H bond, we addressed the question of wether the coordinated donor group can be deprotonated in competition to the reductive elimination of HCl from the iridium(III) centre. Based on experimental and quantum chemical investigations, it is shown that the ability for deprotonation of the coordinated ligand decreases in the order of (R3P)2CH+ > R2NH > (R3P)2BH. The initial product of the reductive elimination of HCl from [(PBP)IrCl(CO)(H)]n (1c), the square planar iridium(I) complex, [(PBP)Ir(CO)]+ (3c), was found to be unstable and further reacts to [(PBP)Ir(CO)2]+ (5c). Comparing the C–O stretching vibrations of the latter with those of related complexes, it is demonstrated that neutral ligands based on tricoordinate boron are very strong donors.

Graphical Abstract

1. Introduction

Tricoordinate boron compounds, BR3, are typically Lewis acids and stabilise their electron deficiency by π-donating substituents, hyperconjugation or dimerisation and formation of two-electron three-centre bonds. In consequence, they can accept electron donation from electron rich metal centres and serve as Z-type ligands [1,2]. More recently, several groups demonstrated that the introduction of π-accepting substituents allows to stabilise an occupied pz-orbital and therewith of a trigonal planar Lewis-base with the general formulae L2BR (III) [3,4,5,6,7,8,9]. Consequently, such compounds are able to serve as electron-donating or L-type ligands, but the coordination chemistry of such nucleophilic boron compounds is rather unexplored [8,9,10].
In particular, the similarity to related carbon compounds of the type L2CH+ (II) and secondary amines (I) caught our attention. Pseudo-tetrahedral, secondary amines (I) can serve as cooperative ligands in homogeneous catalysts (Figure 1), by providing a proton in concerted proton hydride transfers or simply by pre-coordination of the substrate via hydrogen bridge bonds (e.g., in Figure 1, cycle A) [11]. Protonated carbodiphosphoranes of the type (R3P)2CH+ (II) can be deprotonated by strong bases and easily form their deprotonated analogues when coordinated to a metal centre [12]. For the corresponding boron compounds, (R3P)2BH (III), previous studies indicated that the boron-bound hydrogen atom in such ligands is not hydridic [13,14]. Due to the π-accepting nature of the cyanido substituents in compounds like [HB(CN)3], they can be deprotonated [15], which stands in contrast to the reactivity of the majority of hydrogen-containing boron compounds.
Motivated by these observations, we began to study a series of isotypical iridium complexes in their reactivity towards base. Herein, we demonstrate that among this series IIII the carbon-based ligand II is the most acidic ligand, while for the other ligands a competitive reductive elimination is observed. In case of the boron-based ligand, this leads to an unique iridium(I) complex. The comparison with related iridium dicarbonyl complexes reveals strong electron donating properties of donor groups akin to III.

2. Results and Discussion

As the starting point for our study, we choose the isotypical iridium(III) pincer complexes 1a1c to investigate. In this context, we compare the amine based pincer-type complex [{(PPh2CH2CH2)2NH}IrCl(CO)(H)]+ Cl (1a) with the formally carbon(0)- and boron(I)-based complexes [{(dppm)2CH}IrCl(CO)(H)]2+ 2Cl (1b) and [{(dppm)2BH}IrCl(CO)(H)]+ Br (1c) [16]. In principle, the deprotonation of 1a1c can take place at several positions in the complex, but commonly either the central donor group E is deprotonated or the hydrido ligand is abstracted in a reductive elimination (Figure 2).

2.1. Deprotonation vs. Reductive Elimination

The reaction of the cationic complex 1a with one equivalent of LiN(SiMe3)2 results in the formation of a new complex 2a (Figure 3), as judged by the observation of a single resonance at 55.5 ppm in the 31P{1H} NMR spectrum of the reaction mixture (Figure S1). The resonance at −16.12 ppm in the 1H NMR spectrum, corresponding to the hydrido ligand in 1a disappears and the absence of a resonance in this region (0 to −40 ppm) suggests that no hydrido ligand is present in the newly formed 2a (Supplementary Materials). By comparison of NMR spectroscopic data with analogues isopropyl-substituted iridium pincer complexes [17], we concluded that the reductive elimination of HCl is the preferred reaction pathway. Addition of a second equivalent of LiN(SiMe3)2 resulted in the formation of a mixture of complexes and the 31P{1H} NMR spectrum displayed several new singlet resonances as well as a new AB spin system (Supplementary Materials). The latter finding either indicates a conformational change to a facially coordinated ligand with different ligands in trans-position, but this seems to be unlikely for a square pyramidal iridium(I) complex that is already formed with the first equivalent of base. A second possibility involves a β-hydride elimination from the amide ligands and subsequent tautomerisation, as previously observed for different noble metal complexes with this type of ligand [18].
The NMR spectra of the iridium(III) complex 1b at ambient temperature show the presence of the cis- and the trans-isomers (ca. 1:1) as well as small quantities of 3b (ca. 1%) [16]. The 1H NOESY NMR spectrum of 1b at ambient temperature displays exchange correlations between the hydride resonances of cis- and trans-1b, as well as between the resonances of trans-1b and 3b (Figure 4a). These findings suggests the presence of an equilibrium between the two isomers of 1b (Figure 4b). Furthermore one of the isomers (trans-1b) seems to be in an equilibrium with the deprotonated species 1b, even though no additional base is present in the mixture. A broad resonance at 3.51 ppm in the 1H NMR spectrum is assigned to HCl [19], which provides further support for reversible (de)protonation equilibrium. To get further insights about the solution behaviour of 1b, we acquired 1H and 1H{31P} NMR spectra at different temperatures. The ratio of integrals for the hydride resonances enables to estimate the equilibrium constant Kcis/trans at different temperatures. The corresponding Van’t Hoff plot (Figure 4c) displays two regions of linearity between 300 and 270 K (R2 = 0.995) as well as between 260 and 230 K (R2 = 0.937), which might be related to the presence of a second equilibrium or solubility issues at low temperatures. However, a reliable quantification of 3b turned out to be difficult, due to the low concentration at ambient temperature, which decreases even further at lower temperatures. The corresponding exchange rates were accessed by line-shape-analysis of the hydride resonances in the 1H{31P} NMR spectra at different temperatures. An Eyring analysis (Figure 4d) revealed an Gibbs enthalpy of activation Δ G 298 = 69.23 KJ mol 1 for the cis-/trans-isomerisation process.
In view of the primary question, these observations suggest that 1b gets selectively deprotonated at the coordinated donor group. The iridium(III) complex 3b is indeed observed by NMR spectroscopy in reactions with base. As complex 1b, in contrast to 1a and 1c, is dicationic, one would expect a higher acidity of the coordinated donor group, but the acidity of hydrido ligands was previously demonstrated to be increased by several orders of magnitude with an increasing charge of the complex [20].
Addition of an excess base (DBU) to 1b results in the formation of the iridium(I) complex 4b as major product according to the 31P{1H} NMR spectrum of the reaction mixture (Figure 5), which displayed new triplet resonances at 23.4 ppm (2JP,P = 48.5 Hz) and 38.3 ppm (2JP,P = 49.3 Hz). A broad multiplet resonance at 4.01–4.12 ppm with an integral of four in combination with multiplet resonances between 6.9 and 7.8 ppm with an overall integral of 40 protons are observed in the 1H NMR spectrum (Supplementary Materials), while the absence of resonances corresponding to a hydrido ligand or a protonated CDP moiety indicate that a deprotonated pincer ligand is coordinated in 4b. The observation of one band at 1925 cm−1 for the C–O stretching vibration of a carbonyl ligand is in line with an electron-rich mono-carbonyl complex. The composition of the cationic complex [{(dppm)2C}Ir(CO)]+ Cl in 4b was further confirmed by high resolution ESI-MS.
A similar observation to the reaction of 1a is made for the boron-based iridium pincer complex (1c). Treatment of complex 1c with one equivalent LiN(SiMe3)2 leads to the formation of two species according to the 31P{1H} NMR spectrum of the reaction mixture, broadened resonance at −5.6 ppm, as well as a broad resonance at 24.9 and a multiplet at 2.9 ppm, both assignable to the newly formed complex 5c (Figure 6). After removal of all volatiles and washing of the residue with n-hexane, complex 5c is obtained in analytically pure form. The 1H NMR spectrum of 5c shows a complete set of resonances for the dppm arms of the coordinated ligand (Supplementary Materials), while resonances corresponding to a boron-bound hydrogen atom and potential hydrido ligands are absent (Figure 7b). Upon 11B-decoupling a triplet resonance at 3.20 ppm (2JP,H = 23.2 Hz) is observed in the 1H{11B} NMR spectrum, assignable to a boron-bound hydrogen atom, clearly indicating that a reductive elimination is favoured over of the ligand deprotonation. The 11B{1H} NMR spectrum of 5c gives rise to a broadened resonance at −35.4 ppm, which is in agreement with previously reported boron-based donor ligands [8,9,10,13,21]. The identity of 5c was finally confirmed by single crystal X-Ray diffraction experiments (Figure 7a), which revealed a cationic iridium(I) complex with a trigonal bipyramidal environment (τ5 = 0.70) [22]. In addition to the facially coordinated PBP-ligand, two carbonyl ligands are observed, one occupying an equatorial and one an axial coordination site. The Ir–B bond in 5c is with 2.276 Å slightly shorter than in the octahedral iridium(III) complex 1c (dIr−B = 2.285 Å) [16].
As the yield of the dicarbonyl complex 5c was below 50% and no other potential source of carbon monoxide was present in the reaction mixture, we assumed that the formation of 5c proceeds via a square planar iridium(I) intermediate 2c that subsequently reacts in carbonyl transfer step to 5c and unidentified decomposition products (Figure 6). This hypothesis is further verified by an increased yield of 59% in the deprotonation reaction in the presence of carbon monoxide.

2.2. Proton Affinities and Deprotonation Pathways

Quantum chemical investigations using density functional theory (DFT) were performed to get further insights about the reactivity of the reported iridium complexes towards bases. First we confirmed that deprotonation of the coordinated donor group results in an energetic minimum (3a3c) according to the frequency calculation (no imaginary modes) and calculated the proton affinities (PAs) for 3a3c (Table 1 and Figure 8). In agreement with the experimental results, complex 3b exhibits the lowest proton affinity (PA, represents the energy difference between complexes calculated without solvation and counter ions; the energy of free proton is not considered) with 864 kJ·mol−1, while the PAs of the neutral complexes 3a (1129 kJ·mol−1) and 3c (1257 kJ·mol−1) are significantly higher. The low PA of the CDP-group in the coordinated pincer-type ligand indicates that it might be less efficient as internal base in a potential catalyst, but in turn it suggests that protonated CDPs might be potential cooperative groups that facilitate an efficient proton-hydride-transfer from or to the catalyst.
In comparison, the value of 1257 kJ·mol−1 is too high to expect metal-ligand-cooperativity via proton-hydride-transfer, but it clearly suggests that deprotonation of coordinated (R3P)2BH groups should be facile with strong bases in the absence of more acidic sites, which would yield an unprecedented phosphine-stabilized boride.
To elucidate the reductive elimination pathway, we removed a proton from the metal-coordinated hydrido ligand in 1a1c in a gedankenexperiment and performed geometry optimisations. The resulting complexes (6a6c) exhibit elongated iridium chloride distances (Figure 8), but were confirmed as energetic minima by frequency calculations. Although the Ir–Cl distances in 6a6c are in range between a weak bond (2.737 Å) and non-bonding (4.181 Å), the resulting proton affinities may be used as estimate in comparison to 3a3c.
It becomes evident that in case of the amine-based ligand product of ligand- (3a) and metal-deprotonation (6a) exhibit very similar proton affinities (ΔPA = 3 kJ·mol−1), which suggests that both pathways are in principle favourable. The experimentally observed selectivity for the reductive elimination might be kinetically favoured. In case of the protonated CDP-based ligand in 1b the ligand deprotonation is favoured 36 kJ·mol−1 over the deprotonation at the metal site, which again is in line with the experimental observations. Notably, both PAs, of 3b and 6b, are rather low. For the boron-based pincer-type ligand in 1c the deprotonation at the metal centre is clearly favoured.

2.3. Comparison with Related Iridium(I) Dicarbonyl Complexes

In comparison to related trigonal bipyramidal iridium(I) dicarbonyl complexes, 5c exhibits very similar structural features (Table 2). All complexes with two Ph2RP-groups and one carbonyl ligand in the equatorial plane differ in the ligand or donor group in the apical position, trans to the second carbonyl ligand [23,24,25]. With τ5-parameters between 0.58 and 0.75, four of the five complexes are best described as trigonal bipyramidal complexes. In the IR spectrum, two bands for the C–O-stretching frequency are observed for each complex, which in principle allow to estimate the net electron donor ability of the specified donor group in comparison. Like for other dicarbonyl-based ligand parameters [26,27,28], averaging of cis- and trans-influences on symmetric and asymmetric C–O-stretching modes can provide a rough picture of the net donor strength. For the neutral complexes, all values, respectively, indicate an increasing donor ability in the order R3SiCH2 > Cl > Br. The cationic complex with a Ph2RP-group in the apical position gives rise to an increased value of ν ˜ C O (av) = 1996 cm−1, confirming that anionic ligands exhibit stronger donor abilities. An unexpected finding in this context is the low value measured for complex 5c ( ν ˜ C O (av) = 1958 cm−1), which is significantly lower than those of the anionic donor groups (Figure S9). Despite the fact that donor groups based on (Ph3P)2BH are overall neutral, this observation suggests that they are stronger donors than alkyl-ligands, which are known as one of the strongest donors in coordination and organometallic chemistry.

3. Materials and Methods

All experiments were carried out under an atmosphere of purified argon or nitrogen in the MBraun glove boxes LABmaster 130 and UNIlab or using standard Schlenk techniques. THF and diethyl ether were dried over Na/K alloy, n-hexane was dried over LiAlH4, toluene was dried over sodium, dichloromethane was dried over CaH2, methanol was dried over magnesium and ethyl acetate was dried over potassium carbonate. After drying, solvents were stored over appropriate molecular sieves. Deuterated solvents were degassed with freeze-pump-thaw cycles and stored over appropriate molecular sieves under argon atmosphere. Complexes 1a1c synthesised according to previously reported procedures [16].
1H, 13C, 11B and 31P NMR spectra were recorded using Bruker BioSpin GmbH (Rheinstetten, Germany) Avance HD 250, 300 A, DRX 400, DRX 500 and Avance 500 NMR spectrometers at 300 K. 1H and 13C{1H}, 13C-APT (attached proton test) NMR chemical shifts are reported in ppm downfield from tetramethylsilane. The resonance of the residual protons in the deuterated solvent was used as internal standard for 1H NMR spectra. The solvent peak of the deuterated solvent was used as internal standard for 13C NMR spectra. The assignment of resonances in 1H and 13C NMR spectra was further supported by 1H COSY, 1H NOESY, 1H,13C HMQC and 1H,13C HMBC NMR spectra. 11B NMR chemical shifts are reported in ppm downfield from BF3·Et2O and referenced to an external solution of BF3·Et2O in CDCl3. 31P NMR chemical shifts are reported in ppm downfield from H3PO4 and referenced to an external 85 % solution of phosphoric acid in D2O. The following abbreviations are used for the description of NMR data: br (broad), s (singlet), d (doublet), t (triplet), q (quartet), quin (quintet), m (multiplet). FT-IR spectra were recorded by attenuated total reflection of the solid samples on a Bruker Tensor IF37 spectrometer. The intensity of the absorption band is indicated as w (weak), m (medium), s (strong), vs (very strong) and br (broad). HR-ESI mass spectra were acquired with a LTQ-FT mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). The resolution was set to 100,000.
Reactivity of [({Ph2PCH2CH2}2NH)IrCl(CO)(H)]Cl (1a) towards base 20 mg [({Ph2PCH2CH2}2NH)IrCl(CO)(H)]Cl (1a, 27.3 μmol, 1.0 eq.) and 4.6 mg LiHMDS (27.5 μmol, 1.0 eq.) were suspended in 0.6 mL THF-d8. After stirring for 16 h, the resulting light orange suspension was filtered and, after addition of 0.2 ml THF-d8, the first NMR spectra were recorded. [({Ph2PCH2CH2}2NH)Ir(CO)]Cl (2a) was identified as the main product, while small amounts of 1a remained unreacted. Further 4.7 mg of LiHMDS (27.5 μmol, 1.0 eq.) were added, upon which the color changed to a dark orange, and the second set of NMR spectra were recorded.
NMR spectra after addition of 1.0 eq. LiHMDS: 1H NMR (300 MHz, THF-d8, 300 K): δ = 2.61–2.86 (m, 4H, CH2), 3.08–3.46 (m, 4H, CH2), 7.03–7.24 (m, 4H, Harom), 7.25–7.51 (m, 12H, Harom), 7.73–8.03 (m, 4H, Harom) ppm. Neither N–H nor Ir–H resonances could be identified. 31P{1H} NMR (122 MHz, THF-d8, 300 K) δ = 31.7 (s, 1a), 55.5 (br s, 2a) ppm.
NMR spectra after addition of 2.0 eq. LiHMDS: 31P{1H} NMR (122 MHz, THF-d8, 300 K) δ = −3.8 (s), −0.9 (s), 25.0 (s), 31.9 (s, 1a), 36.1 (s), 39.8 (d, JP,P = 291.7 Hz), 52.6 (d, JP,P = 292.3 Hz), 56.1 (br s, 2a) ppm. 1H NMR (300 MHz, THF-d8, 300 K): Due to the multiple decomposition products visible in the 31P{1H} NMR spectrum, no analysis was performed.
Formation of [({dppm}2C)Ir(CO)]Cl (4b) 57 mg [({dppm}2CH)IrCl(CO)(H)]Cl2 (1b, 51.4 μmol, 1.0 eq.) were dissolved in 2 mL deuterated dichloromethane. After addition of 15.3 μL DBU (103 μmol, 2.0 eq.), the solution changed color from colorless to yellow. After removal of the solvent in vacuo, a yellow solid remained, containing [({dppm}2C)Ir(CO)]Cl (4b). 1H NMR (300 MHz, CD2Cl2, 300 K): δ = 4.01–4.12 (m, 4H, CH2), 7.06–7.18 (m, 8H, Harom.), 7.25–7.46 (m, 24H, Harom.), 7.59–7.78 (m, 8H, Harom.) ppm. 13C APT NMR (75 MHz, CD2Cl2, 300 K): δ = 129.0–129.3 (m, Carom.), 131.4 (br s, Carom.), 132.7 (br s, Carom.), 132.9 (t, JC,P = 5.1 Hz, CHarom.), 133.4 (t, JC,P = 7.2 Hz, Carom.) ppm. Neither the carbonyl nor the CH2 resonances were observed. 31P{1H} NMR (122 MHz, CD2Cl2, 300 K) δ = 23.4 (t, 2JC,P = 48.5 Hz), 38.3 (t, 2JC,P = 49.3 Hz) ppm. FT-IR / cm−1: 3050 (w), 2962 (w), 2932 (m), 2925 (m), 2858 (m), 2855 (w), 2013 (w), 1979 (w), 1925 (s, CO), 1646 (s), 1612 (s), 1586 (s), 1481 (m), 1434 (s), 1323 (s), 1207 (w), 1119 (m), 1103 (m), 1097 (s), 1070 (s), 824 (m), 740 (s), 721 (m), 691 (s), 543 (m), 527 (m), 503 (s), 481 (s). HRMS: (ESI+, MeCN/CH2Cl2): 1001.1966 [({dppm}2C)Ir(CO)]+ measured, 1001.1972 calculated, Δ = 0.60 ppm.
Synthesis of [({dppm}2BH)Ir(CO)2]Br (5c) Complex 1c was generated in situ by the reaction of 90.0 mg [IrCl(CO)(PPh3)2] (116 μmol) with 100.0 mg of [(dppm)2BH2]Br (116 μmol, 1.0 eq.) in 5 mL THF. The resulting solution of 1c was cooled to −74 °C and 20.0 mg LiN(SiMe3)2 (116 μmol, 1.0 eq.) dissolved in 2 mL THF were added drop-wise. The reaction mixture was allowed to warm to ambient temperature, the argon atmosphere was replaced by carbon monoxide and the mixture was stirred for further two hours at ambient temperature. All volatiles were removed in vacuo, the residue was washed with 5 mL toluene and dried under high vacuum to yield 74.0 mg of a colorless solid, containing [({dppm}2BH)Ir(CO)2]Br (4c, 68 μmol, 59 %). 31P{1H} NMR (101.3 MHz, CD2Cl2, 300 K): δ = 25.4 (br, 2P, P–B–P), 3.5-2.2 (m, 2P, P–Ir–P) ppm. 11B{1H} NMR (96.3 MHz, CD2Cl2, 300 K): δ = −35.4 (br, 1B, BH) ppm. Only resonances that are change upon 11B-decoupling are reported in the 1H{11B} NMR spectrum. 1H NMR (300 MHz, CD2Cl2, 300 K): δ = 7.51–7.66 (m, 4H, Harom.), 7.40–7.49 (m, 8H, Harom.), 7.08–7.31 (m, 8H, Harom.), 6.82–7.10 (m, 20H, Harom.), 5.42–5.61 (m, 2H, CH2), 4.07–4.16 (m, 2H, CH2) ppm. 1H{11B} NMR (300 MHz, CD2Cl2, 300 K) δ = 3.20 (t, 2JHP = 23.2 Hz, 1H, BH) ppm. 13C{1H} NMR (121.5 MHz, CD2Cl2, 300 K) δ = 134.7 (vt, 4C, Carom.), 133.5 (s, 4C, Carom.), 133.0 (s, 4C, Carom.), 132.0 (s, 4C, Carom.), 131.3 (s, 4C, Carom.), 130.9 (vt, 4C, Carom.), 130.2 (s, 4C, Carom.), 129.2 (s, 4C, Carom.), 129.2 (s, 4C, Carom.), 128.9 (s, 4C, Carom.), 128.7 (s, 4C, Carom.), 128.3 (s, 4C, Carom.), 33.5 (vt, 1C, CH2), 30.3 (vt, 1C, CH2) ppm. FT-IR: ν ˜ /cm−1 = 3050 (w), 3017 (w), 2962 (w), 2823 (w), 2724 (w), 2000 (s, CO), 1916 (s, CO), 1586 (w), 1574 (w), 1483 (m), 1434 (s), 1379 (w), 1333 (w), 1306 (w), 1260 (m), 1094 (s), 1024 (s), 869 (w), 797 (s), 778 (s), 731 (vs), 685 (vs), 616 (w), 554 (m), 523 (s), 480 (s). HRMS (ESI+, MeOH) m/z = 969.1884 [({dppm}2BH)Ir(CO)2]+, calc. 969.1887 (Δ = 0.31 ppm).

4. Conclusions

In the current manuscript, we reported the first iridium(I) complex formally containing phosphine-stabilised borylene as a donor group. The comparison to related iridium(I) dicarbonyl complexes suggests strong donor properties of this type of nucleophilic boron compounds. In an internal competition with a hydrido-ligand, the reactivity towards base reveals that analogous carbon compounds and protonated CDPs are easy to deprotonate, while only strong bases contribute to deprotonate phosphine-stabilized borylenes in the coordination sphere of a central metal atom.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/7/5/61/s1, Figures S1–S12: NMR and IR spectra of compounds 2a, 4b and 5c; Table S1: crystallographic data for compound 5c; xyz-coordinates.

Author Contributions

L.M., C.S. and L.A. performed the experiments. All calculations were made by L.M., R.L. and L.M. wrote the manuscript. R.L. designed and directed the project.

Funding

This work was supported by the Deutsche Forschungsgemeinschaft (LA 2830/3-2, 2830/5-1 and 2830/6-1).

Acknowledgments

R.L. is grateful to S. Dehnen for her continuous support.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
CDPcarbodiphosphorane
DBU1,8-Diazabicyclo[5.4.0]undec-7-ene
dppm1,1-bis(diphenylphosphino)methane
DFTdensity functional theory
ESIelectro spray ionisation
HMDShexamethyldisilazane
NMRnuclear magnetic resonance
HRMShigh resolution mass spectrometry
THFtetrahydrofurane

References

  1. Bouhadir, G.; Bourissou, D. Complexes of ambiphilic ligands: reactivity and catalytic applications. Chem. Soc. Rev. 2016, 45, 1065–1079. [Google Scholar] [CrossRef]
  2. Amgoune, A.; Bourissou, D. σ-Acceptor, Z-type ligands for transition metals. Chem. Commun. 2011, 47, 859–871. [Google Scholar] [CrossRef] [PubMed]
  3. Braunschweig, H.; Dewhurst, R.D.; Hupp, F.; Nutz, M.; Radacki, K.; Tate, C.W.; Vargas, A.; Ye, Q. Multiple complexation of CO and related ligands to a main-group element. Nature 2015, 522, 327–330. [Google Scholar] [CrossRef]
  4. Landmann, J.; Sprenger, J.A.P.; Bertermann, R.; Ignat’ev, N.; Bernhardt-Pitchougina, V.; Bernhardt, E.; Willner, H.; Finze, M. Convenient access to the tricyanoborate dianion B ( CN ) 3 2 and selected reactions as a boron-centred nucleophile. Chem. Commun. 2015, 51, 4989–4992. [Google Scholar] [CrossRef] [PubMed]
  5. Bernhardt, E.; Bernhardt-Pitchougina, V.; Willner, H.; Ignatiev, N. “Umpolung” at boron by reduction of [B(CN)4] and formation of the dianion [B(CN)3]2−. Angew. Chem. Int. Ed. 2011, 50, 12085–12088. [Google Scholar] [CrossRef] [PubMed]
  6. Kinjo, R.; Donnadieu, B.; Celik, M.A.; Frenking, G.; Bertrand, G. Synthesis and characterization of a neutral tricoordinate organoboron isoelectronic with amines. Science 2011, 333, 610–613. [Google Scholar] [CrossRef]
  7. Ruiz, D.A.; Melaimi, M.; Bertrand, G. An efficient synthetic route to stable bis(carbene)borylenes [(L1)(L2)BH]. Chem. Commun. 2014, 50, 7837–7839. [Google Scholar] [CrossRef] [PubMed]
  8. Kong, L.; Li, Y.; Ganguly, R.; Vidovic, D.; Kinjo, R. Isolation of a Bis(oxazol-2-ylidene)-phenylborylene adduct and its reactivity as a boron-centered nucleophile. Angew. Chem. Int. Ed. 2014, 53, 9280–9283. [Google Scholar] [CrossRef] [PubMed]
  9. Kong, L.; Ganguly, R.; Li, Y.; Kinjo, R. Diverse reactivity of a tricoordinate organoboron L2PhB: (L = oxazol-2-ylidene) towards alkali metal, group 9 metal, and coinage metal precursors. Chem. Sci. 2015, 6, 2893–2902. [Google Scholar] [CrossRef] [PubMed]
  10. Ruiz, D.A.; Ung, G.; Melaimi, M.; Bertrand, G. Deprotonation of a Borohydride: Synthesis of a Carbene-Stabilized Boryl Anion. Angew. Chem. Int. Ed. 2013, 52, 7590–7592. [Google Scholar] [CrossRef]
  11. Dub, P.A.; Gordon, J.C. The mechanism of enantioselective ketone reduction with Noyori and Noyori–Ikariya bifunctional catalysts. Dalton Trans. 2016, 45, 6756–6781. [Google Scholar] [CrossRef]
  12. Maser, L.; Herritsch, J.; Langer, R. Carbodiphosphorane-based nickel pincer complexes and their (de)protonated analogues: Dimerisation, ligand tautomers and proton affinities. Dalton Trans. 2018, 47, 10544–10552. [Google Scholar] [CrossRef] [PubMed]
  13. Vondung, L.; Frank, N.; Fritz, M.; Alig, L.; Langer, R. Phosphine-Stabilized Borylenes and Boryl Anions as Ligands? Redox Reactivity in Boron-Based Pincer Complexes. Angew. Chem. Int. Ed. 2016, 55, 14450–14454. [Google Scholar] [CrossRef] [PubMed]
  14. Vondung, L.; Jerabek, P.; Langer, R. Ligands Based on Phosphine-Stabilized Aluminum(I), Boron(I), and Carbon(0). Chem. Eur. J. 2019, 25, 3068–3076. [Google Scholar] [CrossRef] [PubMed]
  15. Landmann, J.; Keppner, F.; Hofmann, D.B.; Sprenger, J.A.; Höring, M.; Zottnick, S.H.; Müller-Buschbaum, K.; Ignat’ev, N.V.; Finze, M. Deprotonation of a Hydridoborate Anion. Angew. Chem. Int. Ed. 2017, 56, 2795–2799. [Google Scholar] [CrossRef]
  16. Maser, L.; Schneider, C.; Vondung, L.; Alig, L.; Langer, R. Quantifying the Donor Strength of Ligand-Stabilized Main Group Fragments. J. Am. Chem. Soc. 2019. [Google Scholar] [CrossRef]
  17. Friedrich, A.; Ghosh, R.; Kolb, R.; Herdtweck, E.; Schneider, S. Iridium olefin complexes bearing dialkylamino/amido PNP pincer ligands: Synthesis, reactivity, and solution dynamics. Organometallics 2009, 28, 708–718. [Google Scholar] [CrossRef]
  18. Schneider, S.; Meiners, J.; Askevold, B. Cooperative aliphatic PNP amido pincer ligands-versatile building blocks for coordination chemistry and catalysis. Eur. J. Inorg. Chem. 2012, 2012, 412–429. [Google Scholar] [CrossRef]
  19. Cheng, F.; Yang, X.; Peng, H.; Chen, D.; Jiang, M. Well-controlled formation of polymeric micelles with a nanosized aqueous core and their applications as nanoreactors. Macromolecules 2007, 40, 8007–8014. [Google Scholar] [CrossRef]
  20. Morris, R.H. Estimating the acidity of transition metal hydride and dihydrogen complexes by adding ligand acidity constants. J. Am. Chem. Soc. 2014, 136, 1948–1959. [Google Scholar] [CrossRef]
  21. Grätz, M.; Bäcker, A.; Vondung, L.; Maser, L.; Reincke, A.; Langer, R. Donor ligands based on tricoordinate boron formed by B–H-activation of bis(phosphine)boronium salts. Chem. Commun. 2017, 53, 7230–7233. [Google Scholar] [CrossRef] [PubMed]
  22. Addison, A.W.; Rao, T.N.; Reedijk, J.; van Rijn, J.; Verschoor, G.C.; Trans, D.; Addison, A.W.; Rao, T.N. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6- dithiaheptane]copper(II) perchlorate. Dalton Trans. J. Chem. Soc. 1984, 1349–1356. [Google Scholar] [CrossRef]
  23. Heins, W.; Mayer, H.A.; Fawzi, R.; Steimann, M. Stereochemical and Electronic Control of Functionalized Tripodal Phosphines. Reactivity of the Adamantane-Type Ir(tripod)(CO)Cl (tripod = cis,cis-1,3,5-(PPh2)3-1,3,5-X3C6H6; X = H, COOMe, CN) Complexes toward H+, H2, CO, and C2H4. Organometallics 1996, 15, 3393–3403. [Google Scholar] [CrossRef]
  24. Fox, D.J.; Duckett, S.B.; Flaschenriem, C.; Brennessel, W.W.; Schneider, J.; Gunay, A.; Eisenberg, R. A Model Iridium Hydroformylation System with the Large Bite Angle Ligand Xantphos: Reactivity with Parahydrogen and Implications for Hydroformylation Catalysis. Inorg. Chem. 2006, 45, 7197–7209. [Google Scholar] [CrossRef] [PubMed]
  25. Passarelli, V.; Pérez-Torrente, J.J.; Oro, L.A. Intramolecular C–H Oxidative Addition to Iridium(I) in Complexes Containing a N,N′-Diphosphanosilanediamine Ligand. Inorg. Chem. 2014, 53, 972–980. [Google Scholar] [CrossRef] [PubMed]
  26. Chianese, A.R.; Li, X.; Janzen, M.C.; Faller, J.W.; Crabtree, R.H. Rhodium and iridium complexes of N-heterocyclic carbenes via transmetalation: Structure and dynamics. Organometallics 2003, 22, 1663–1667. [Google Scholar] [CrossRef]
  27. Kelly III, R.A.; Clavier, H.; Giudice, S.; Scott, N.M.; Stevens, E.D.; Bordner, J.; Samardjiev, I.; Hoff, C.D.; Cavallo, L.; Nolan, S.P. Determination of N-Heterocyclic Carbene (NHC) Steric and Electronic Parameters using the [(NHC)Ir(CO)2Cl] System. Organometallics 2008, 27, 202–210. [Google Scholar] [CrossRef]
  28. Wolf, S.; Plenio, H. Synthesis of (NHC)Rh(cod)Cl and (NHC)RhCl(CO)2 complexes—Translation of the Rh- into the Ir-scale for the electronic properties of NHC ligands. J. Organomet. Chem. 2009, 694, 1487–1492. [Google Scholar] [CrossRef]
Figure 1. (a) Secondary amines (I), protonated carbodiphosphoranes (II) and phosphine-stabilized borylenes (III) in comparison; (b) Secondary amine ligands and their role in cooperative catalysis in comparison to the analogous metal complexes with II and III as ligands.
Figure 1. (a) Secondary amines (I), protonated carbodiphosphoranes (II) and phosphine-stabilized borylenes (III) in comparison; (b) Secondary amine ligands and their role in cooperative catalysis in comparison to the analogous metal complexes with II and III as ligands.
Inorganics 07 00061 g001
Figure 2. Cooperative ligand site vs. redox reactivity—principle reaction pathways of octahedral iridium(III) complexes 1a1c towards base (n = +, 2+). X = Cl (a,b), Br (c)
Figure 2. Cooperative ligand site vs. redox reactivity—principle reaction pathways of octahedral iridium(III) complexes 1a1c towards base (n = +, 2+). X = Cl (a,b), Br (c)
Inorganics 07 00061 g002
Figure 3. Reactivity of 1a towards base.
Figure 3. Reactivity of 1a towards base.
Inorganics 07 00061 g003
Figure 4. (a) Hydride region in the 1H NOESY NMR spectrum at ambient temperature, showing chemical exchange correlations; (b) Equilibrium of the complexes in solution; (c) Van’t Hoff plot for the cis-/trans-isomerisation of 1b; (d) Eyring plot for the cis-/trans-isomerisation of 1b.
Figure 4. (a) Hydride region in the 1H NOESY NMR spectrum at ambient temperature, showing chemical exchange correlations; (b) Equilibrium of the complexes in solution; (c) Van’t Hoff plot for the cis-/trans-isomerisation of 1b; (d) Eyring plot for the cis-/trans-isomerisation of 1b.
Inorganics 07 00061 g004
Figure 5. Reactivity of 1b towards base.
Figure 5. Reactivity of 1b towards base.
Inorganics 07 00061 g005
Figure 6. Reactivity of 1c towards base.
Figure 6. Reactivity of 1c towards base.
Inorganics 07 00061 g006
Figure 7. (a) Molecular structure of the cationic complex in 5c in the solid state (ellipsoids are drawn at 50% probability level; carbon atoms of the phenyl rings, carbon-bound hydrogen atoms, co-crystallized solvent molecules and counter ion are omitted for clarity); (b) Selected NMR spectra of complex 5c.
Figure 7. (a) Molecular structure of the cationic complex in 5c in the solid state (ellipsoids are drawn at 50% probability level; carbon atoms of the phenyl rings, carbon-bound hydrogen atoms, co-crystallized solvent molecules and counter ion are omitted for clarity); (b) Selected NMR spectra of complex 5c.
Inorganics 07 00061 g007
Figure 8. Proton affinities and DFT-optimized structures of 3a3c and 6a6c (G16, B97D/def2-TZVPP).
Figure 8. Proton affinities and DFT-optimized structures of 3a3c and 6a6c (G16, B97D/def2-TZVPP).
Inorganics 07 00061 g008
Table 1. Calculated Proton affinities of complexes 3a3c and 6a6c (G16, B97D/def2-TZVPP).
Table 1. Calculated Proton affinities of complexes 3a3c and 6a6c (G16, B97D/def2-TZVPP).
Donor in 1ReactivityPA(3)/kJ·mol−1ReactivityPA(6)/kJ·mol−1ΔPA/kJ·mol−1
R2NH1a3a11291a6a11263
(Ph2RP)2CH1b3b8641b6b900−36
(Ph2RP)2BH1c3c12571c6c117582
Table 2. Comparison of Iridium(I) dicarbonyl complexes from literature with the new complex 5c.
Table 2. Comparison of Iridium(I) dicarbonyl complexes from literature with the new complex 5c.
ComplexDonorτ5 ν ˜ CO /cm−1 ν ˜ CO (av)/cm−1Ref.
Inorganics 07 00061 i001Ph2RP0.422047, 19441996[23]
Inorganics 07 00061 i002Br0.702023, 19501987[24]
Inorganics 07 00061 i003Cl0.582017, 19441981[24]
Inorganics 07 00061 i004R3SiCH 2 0.752001, 19271964[25]
Inorganics 07 00061 i005(R3P)3BH0.702000, 19161958this work

Share and Cite

MDPI and ACS Style

Maser, L.; Schneider, C.; Alig, L.; Langer, R. Comparing the Acidity of (R3P)2BH-Based Donor Groups in Iridium Pincer Complexes. Inorganics 2019, 7, 61. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7050061

AMA Style

Maser L, Schneider C, Alig L, Langer R. Comparing the Acidity of (R3P)2BH-Based Donor Groups in Iridium Pincer Complexes. Inorganics. 2019; 7(5):61. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7050061

Chicago/Turabian Style

Maser, Leon, Christian Schneider, Lukas Alig, and Robert Langer. 2019. "Comparing the Acidity of (R3P)2BH-Based Donor Groups in Iridium Pincer Complexes" Inorganics 7, no. 5: 61. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics7050061

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