Next Article in Journal
The Structure and Location of 18-Crown-6 Ether in Zeolites RHO and ZK-5
Previous Article in Journal
Study on the Effects of Biologically Active Amino Acids on the Micellization of Anionic Surfactant Sodium Dodecyl Sulfate (SDS) at Different Temperatures
Previous Article in Special Issue
Complexation Behavior of Pinene–Bipyridine Ligands towards Lanthanides: The Influence of the Carboxylic Arm
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enantiopure Cyclometalated Rh(III) and Ir(III) Complexes Displaying Rigid Configuration at Metal Center: Design, Structures, Chiroptical Properties and Role of the Iodide Ligand

1
Institut Parisien de Chimie Moléculaire (IPCM) UMR CNRS 8232, Sorbonne Université—Campus Pierre et Marie Curie, 4 Place Jussieu, CEDEX 05, 75252 Paris, France
2
Aix Marseille Univ, CNRS, Centrale Marseille, iSm2, 13007 Marseille, France
*
Author to whom correspondence should be addressed.
Chemistry 2022, 4(1), 156-167; https://doi.org/10.3390/chemistry4010014
Submission received: 11 February 2022 / Revised: 8 March 2022 / Accepted: 8 March 2022 / Published: 12 March 2022

Abstract

:
Enantiopure N-heterocyclic carbene half-sandwich metal complexes of the general formula [Cp*M(C^C:)I] (M = Rh, Ir; C^C: = NI-NHC; NI-H = Naphthalimide; NHC = N-heterocyclic carbene) are reported. The rhodium compound was obtained as a single isomer displaying six membered metallacycle and was resolved on chiral column chromatography to the corresponding enantiomers (S)-[Cp*Rh(C^C:)I] (S)-2 and (R)-[Cp*Rh(C^C:)I] (R)-2. The iridium congener, however, furnishes a pair of regioisomers, which were resolved into (S)-[Cp*Ir(C^C:)I] (S)-3 and (R)-[Cp*Ir(C^C:)I] (R)-3 and (S)-[Cp*Ir(C^C:)I] (S)-4 and (R)-[Cp*Ir(C^C:)I] (R)-4. These regioisomers differ from each other, only by the size of the metallacycle; five-membered for 3 and six-membered for 4. The molecular structures of (S)-2 and (S)-4 are reported. Moreover, the chiroptical properties of these compounds are presented and discussed. These compounds display exceptional stable configurations at the metal center in solution with enantiomerization barrier ΔG up to 124 kJ/mol. This is because the nature of the naphthalimide-NHC clamp ligand and the iodide ligand contribute to their configuration’s robustness. In contrast to related complexes reported in the literature, which are often labile in solution.

Graphical Abstract

1. Introduction

Chirality is an ever-fascinating topic and occurs in many fields of science [1,2,3,4]. In the area of transition metal complexes, Brunner [5,6], Gladysz [7,8], von Zelewsky [9], Meggers [10], Constable [11] and others [12,13] have made great contributions to the advancement and comprehension of the elements that control the chirality at metal centers at the molecular and supramolecular levels [11,14]. For instance, octahedral iridium complexes displaying helical chirality (Δ, Λ) show a stable configuration at the metal center [15,16,17,18,19]. Coordination and organometallic complexes with planar chirality also show stable configuration [20,21,22,23,24,25]. In contrast, half-sandwich rhodium and iridium complexes with piano stool geometry, displaying central chirality are labile in solution, as demonstrated by Brunner and co-workers and others [26,27,28,29]. More recently, efforts were devoted to using strongly coordinated N-heterocyclic carbene (NHC) ligands to increase the stability at the metal center; however, only a few examples were reported [30,31,32].
We recently described the synthesis of some half-sandwich iridium complexes displaying stable configuration at the metal center [32]. Such compounds contain Cp*Ir complexes in which the NHC unit is directly attached to a naphthalimide (NI) molecule. The latter strongly chelates the metal center; moreover, an iodide ligand completes the coordination sphere around the metal center. Iodide is known to act as a strong σ-donor and a weak-leaving group [33], thus we anticipated such a combination of ligands should bring rigidity to the configuration of the metal center. In this work, we extended our synthetic procedures to the related rhodium congener. Interestingly only one regioisomer complex is obtained, displaying a six-membered metallacycle. The iridium complexes are obtained as two regioisomers in a rough 1:1 ratio as a mixture of five-membered and six membered iridacycles (Scheme 1).
The racemic complexes were all resolved using chiral column chromatography. The chiroptical optical properties were studied and the configurational stability were investigated by measuring the enantiomerization barrier at T = 60 °C for these complexes and were found to be up to ΔG enantiomerization = 124 kJ/mole, suggesting that these compounds display high configurational stability at room temperature.

2. Results and Discussions

2.1. Synthesis and Characterization

The target compounds 24 were prepared by treatment of two equivalents of the imidazolium salt (1) with the chloro-bridged dimeric precursor [Cp*M(μ-Cl)Cl]2 in the presence of Cs2CO3 and NBu4I, in CH3CN under reflux overnight. When [Cp*Rh(μ-Cl)Cl]2 was used as starting material only one single rhodium carbene compound identified as [Cp*Rh(NI-NHC)I] (2) (NI-H = Naphthalimide) was obtained in good yield as red microcrystalline solid after eluting on a column chromatography. The 1H-NMR of 2 recorded in CDCl3 showed the presence of a singlet assigned to Cp*Rh protons at δ 1.51 ppm and the aromatic protons of the naphthalimide and NHC moiety appeared in the range of δ 7.1 to δ 8.61 ppm.
Starting with iridium precursor [Cp*Ir(μ-Cl)Cl]2, under similar experimental conditions, two regioisomeric complexes 3 and 4 were obtained. These compounds display the same chemical composition but differ only with the size of the iridacycle. The formation of the two isomers results from the different modes of metallacyclization; (a) whether the cyclization process takes place at the arene attached directly to the NHC moiety to give the five-membered metallacycle. (b) Metallacylization occurs at the other arene ring generating the six-membered metallacycle. Both isomers 3 and 4 were obtained in a 1:1 ratio. The 1H-NMR of 4 recorded in CDCl3 displayed a singlet at δ 1.54 ppm, assigned to the Cp*Ir protons, this upfield shift is a diagnosis to the formation of a six-membered metallacylce. The aromatic protons of the naphthalimide and NHC moiety appeared in the range of δ 7.1 to δ 8.57 ppm. These data are similar to those obtained for complex 2. The 1H-NMR of compound 3, on the other hand, displayed a singlet at δ 1.96 ppm to the Cp*Ir protons, which is downfield with respect to the six-membered metallacycle complexes 2 and 4. Moreover, the aromatic protons of the naphthalimide and NHC moiety appeared downfield in the range 7.1 to 9.2 ppm and displayed a different pattern. Furthermore, 2D COSY, ROESY, HSQC, and HMBC experiments carried out on complexes 3 and 4 allowed us to fully characterize both isomers (Figures S1 and S2, 2D ROESY Spectra of 3 and 4). In addition, the identity of these molecules was also confirmed by determining the molecular structures of the enantiopure complexes (S)-2 and (S)-4 (Figure 1) (vide infra).

2.2. Chiral Resolution of Complexes [Cp*M(C^C:)I] (M = Rh, (2); M = Ir (34) and Chiroptical Properties

Half-sandwich rhodium and iridium N-heterocyclic carbene complexes (24) were resolved on the chiral stationary phase (Experimental details are given in the Supplementary Materials). For the rhodium complex, Chiralpak IF column was chosen using heptane/ethanol/dichloromethane (50/30/20) mixture as eluent. The related iridium complexes 3 and 4 were resolved using (S,S)-Whelk-O1 chiral column and using the same solvent mixture as eluent. Remarkably all complexes were stable during the resolution process highlighting their robustness and were obtained with >98% ee. The CD spectra of the enantiomers of 2, 3, and 4 are given in Figure 2, Figure 3 and Figure 4. The optical rotations of both enantiomers were monitored as well (experimental details of their resolution are given in the Supplementary Materials).
The CD curves of both enantiomers (Figure 2) displayed an excellent mirror image relationship. For instance, the CD trace of S-2 displayed a positive band at 200 nm (Δε = +42) and a negative band at 265 nm (Δε = −42), followed by a broad positive band at 390 nm (Δε = +36) and a negative band at 445 nm (Δε = −16). The first eluted complex (CD green trace) was crystallized and provided suitable crystals for X-ray diffraction study. The X-ray molecular structure confirmed the identity of the complex (vide infra). Gratifyingly all enantiopure compounds displayed stable configuration in strongly coordinated CH3CN solution and did not epimerize upon standing for several days in solution as demonstrated by their CD curves (vide infra).
On the other hand, the CD curves of both enantiomers of 3 displayed again an excellent mirror image patterns confirming the enantiomeric relationship between the two complexes (S)-3 and (R)-3. It is noteworthy that the positive and negative Cotton bands above 350 nm appeared weaker than those observed for the enantiomers of 2 and 4 suggesting perhaps the influence of the size of the five- or six-membered metallacycles (vide infra)
The CD spectra of both enantiomers of 4 are shown in Figure 4. The traces showed opposite image absorption bands, confirming the enantiomeric relationship. Moreover, the absorption profile appears to be comparable to those of complex 2 displaying a six-membered metallacycle. For instance, the CD trace of (S)-4 showed a negative cotton band at 267 nm (Δε = −26), and a broad positive band at 371 nm (Δε = +27) comparable to those observed for (S)-2. Crystals of the first eluted sample (S)-4 were obtained for X-ray diffraction. The structure is presented and discussed in the next section.

2.3. Molecular Structures of S-2 and S-4

Single crystals of (S)-2 were obtained by slow diffusion of diethyl ether into a CH2Cl2 solution of the complex. X-ray crystallography confirmed the proposed structure of complex (S)-2 (Figure 5).
The structure shows the formation of the six-membered rhodacycle, which adopts a boat conformation. The absolute configuration around the metal center was determined to be (S) by the refinement of the Flack x parameter (Table 1). The metal center chelated by the (C^C:) NHC-naphthalimide ligand, is symmetrically bound to η-Cp*. Finally, the iodide ligand completes the coordination sphere around the rhodium and confers a distorted tetrahedral geometry. The iridium complex (S)-4 was found to be isostructural to that of the rhodium complex (S)-2 (Figure 6). We note however the nBu-group of the naphthalimide points downward opposite to that observed for (S)-2. The absolute configuration of the complex was assigned by refining the Flack × parameter (Table 1).

2.4. Determination of the Enantiomerization Barriers and Configurational Stability

To evaluate the stability of the enantiopure complex, (R)-2 was dissolved in acetonitrile, heated at 60 °C and then analyzed by chiral HPLC. Complex 2 was chemically stable but underwent racemization. Monitoring of the enantiomeric excess versus time revealed a clean first-order reaction and allowed the determination of an enantiomerization barrier of 115.9 kJ/mol. Half-life time was extrapolated at 25 °C by considering zero enantiomerization entropy: 130 days for 2 at 25 °C in acetonitrile. So, enantiopure 2 can be handled and evaporated without any risk of enantiopurity loss. Similar racemization kinetics studies performed on (S)-3 and (R)-4, provided the enantiomerization barriers in acetonitrile, to be 124.4 kJ/mol and 122.3 kJ/mol, respectively. These values confirm the chemical robustness of these complexes and allowed the estimation of their half-life times at 25 °C, to be 11 and 5 years for 3 and 4, respectively. Highlighting the role of the metallacycle five-membered size versus six-membered size. The configurational stability is weaker in ethanol, with lower enantiomerization barriers of 4 kJ/mol (see Supporting Information for details). The racemization is faster with rhodium than with iridium, and in protic solvents, this facilitates the iodide decoordination as one might expect. These results underline the importance of iodine on the stabilization of the chiral configuration at the metal center. For comparison purposes, we note that Brunner and co-workers have studied the half-sandwich complexes [Cp*M(N^N*)Cl] (M = Rh, Ir) where N^N* is the anion of (+)-2-N-[(S)-1-phenylethylpyrrolcarbaldimine] and showed that such compounds are configurationally unstable at metal centers. Time-dependent integration of 1H NMR signals of the complexes revealed that the epimerization is a first-order reaction. The half-lives in CD2Cl2 solution at −50 °C were 19.4 min for the rhodium complex and 30.6 min for the iridium compound [36].

2.5. Discussion

As mentioned in the introduction of this manuscript, half-sandwich metal complexes continue to attract attention due to their importance in medicinal chemistry [37,38], materials science [39,40] and catalysis [41,42,43]. Of prime importance is the control of the configurational stability at the metal center. For instance, Brunner [26], White [27] and others [29] demonstrated that half-sandwich metal complexes of rhodium and iridium racemize in solution, because they display labile configuration at the metal center.
More recently, efforts were made to use strong donor NHC ligands [44,45] associated with helicene [31] or other elements of chirality to increase the configurational rigidity at the metal center with some success. Thus, our complexes presented in this work display only one stereogenic element of chirality. This is a rare example of chiral half-sandwich metal complexes displaying exceptional rigidity at the metal center. The data obtained from the determination of the free energy of enantiomerization confirm this behavior.
We feel this unique class of complexes owe their rigid configuration to the nature of the NI-NHC (C^C:) ligand that acts as a clamp and strongly chelates the metal center. Moreover, the iodide ligand is known to act as a strong donor ligand but a weak-leaving group [46], which contributes to the robustness of the configuration at the metal center in these complexes. Therefore, our judicious choice to combine strongly NHC-chelating (C^C:) ligands and iodide generate a novel class of chiral cyclometalated complexes with exceptional configuration at the metal center which might show interesting applications in a variety of fields, spanning from medical chemistry, organometallic catalysis to materials science.

3. Materials and Methods

General Synthetic Procedure

All synthetic manipulations were carried out under argon using Schlenk tube techniques. The 1H and 13C, and spectra were recorded in CD2Cl2 and CD3CN using a Bruker Avance 400 and 300 NMR spectrometer at 400.13 and 100.61 and 76.31 MHz, respectively. The azolium salt [IMZ-NI-H][I] (1) was prepared according to our procedure published in the literature [47,48,49].
Synthesis of 2. In a dry Schlenk tube, imidazolium salt (1) (150 mg, 0.33 mmol), [Cp*Rh(μCl)Cl]2 (60 mg, 0.11 mmol), Cs2CO3 (82 mg, 0.253 mmol) and NBu4I (243 mg, 0.66 mmol) were mixed in 15 mL of distilled and degassed acetonitrile, the solution was refluxed overnight. The resulting solution was cooled at r.t., filtered to remove a precipitate and washed with distilled acetonitrile. After evaporation of the solvents, the crude solid was purified by chromatography column (SiO2) with a mixture of CH2Cl2/Acetone (99/1). The desired fractions were then concentrated to give product 2 as orange miscrocrystalline solid in 65% yield.
Chemistry 04 00014 i001
1H NMR (300 MHz, CDCl3) δ 8.61 (dd, J = 8.0, 2.1 Hz, 2H, H9,11napht), 8.21 (d, J = 7.8 Hz, 1H, H12napht), 7.78 (d, J = 2.4 Hz, 1H, H3imid), 7.65 (d, J = 8.0 Hz, 1H, H4napht), 7.28 (q, J = 2.3 Hz, 1H, H2imid), 4.20 (t, J = 7.7 Hz, 2H, H17), 4.06 (d, J = 2.2 Hz, 3H, H16), 1.76 (t, J = 8.2 Hz, 2H, H18), 1.51 (s, 15H, CH3 Cp*), 1.28 (m, 2H, H19), 1.01 (m, J = 6.4, 5.7, 3.8 Hz, 3H, H20).
13C NMR (101 MHz, CD2Cl2) δ 178.02 (C14), 177.58 (C15), 165.12 (C1), 163.22 (Cl0), 144.34 (C11), 141.09 (C5), 129.84 (C9), 128.66 (C6), 128.28 (C7), 125.07 (C2), 122.44 (C8), 119.86 (C3), 116.92 (C13), 112.96 (C4), 99.55 (Cq Cp*), 41.29 (C16), 39.85 (C17), 30.27 (C18), 20.43 (C19), 13.65 (C20), 9.37 (CH3 Cp*). Anal. calcd. for 2 C 51.67, H 4.77, N 6.03; Found: C 51.99, H 4.77, N 5.72.
Synthesis of 3 and 4. In a dry Schlenk tube, imidazolium salt (1) (88 mg, 0.185 mmol), [Cp*Ir(μCl)Cl]2 (50 mg, 0.061 mmol), Cs2CO3 (46 mg, 0.14 mmol) and NBu4I (137 mg, 0.37 mmol) were mixed in 15 mL of distilled and degassed acetonitrile, the solution was refluxed overnight. The resulting solution was cooled at r.t., filtered to remove a precipitate and washed with distilled acetonitrile. After evaporation of the solvents, the crude solid was purified by chromatography column (SiO2) with a mixture of CH2Cl2/Acetone (99/1). The desired fractions were then concentrated to give products 3 and 4 in almost 1/1 ratio.
Chemistry 04 00014 i002
3: A crystalline red powder. Yield 32 mg (33%).1H NMR (500 MHz, CDCl3)). δ 9.09 (s, 1H, H9 napht), 8.48 (dd, J = 8.7; 0.9 Hz, 1H, H10 napht), 8.45 (dd, J = 7.3; 0.9 Hz, 1H, H12 napht), 8.07 (d, J = 2.3 Hz, 1H, H3 imid), 7.71 (dd, J = 8.7, 7.3 Hz, 1H, H11 napht), 7.19 (d, J = 2.3 Hz, 1H, H2 imid), 4.23–4.15 (m, 2H, H17), 3.98 (s, 3H, H16), 1.96 (s, 15H, CH3 Cp*), 1.76–1.70 (m, 2H, H18), 1.50–1.42 (m, 2H, H19), 0.98 (t, J = 7.4 Hz, 3H, H20)13C NMR (125 MHz, Chloroform-d) δ 169.4 (C1), 165.0 (C14), 164.3 (C15), 145.9 (C5), 143.4 (C9), 142.3 (C4), 128.0 (C12), 127.0 (C7), 125.8 (C11), 125.3 (C10), 123.5 (C13), 122.2 (C2), 120.0 (C6), 118.8 (C3), 117.9 (C8), 93.0 (Cq.Cp*), 40.3 (C17), 38.4 (C16), 30.4 (C18), 20.6 (C19), 14.0 (C20), 10.5 (CH3.Cp*). Anal. calcd. for 3 C 45.74, H 4.35, N 5.21; Found: C 45.71, H 4.39, N 5.08.
Chemistry 04 00014 i003
4: A crystalline red powder. Yield 38 mg (39%). 1H NMR (500 MHz, CDCl3) δ 8.57 (d, J = 8.1 Hz, 1H, H9 napht), 8.38 (d, J = 7.8 Hz, 1H, H11 napht), 8.16 (d, J = 7.8 Hz, 1H, H12 napht), 7.69 (d, J = 2.3 Hz, 1H, H3 imid), 7.54 (d, J = 8.1 Hz, 1H, H4 napht), 7.22 (d, J = 2.3 Hz, 1H, H2 imid), 4.18–4.14 (m, 2H, H17), 3.98 (s, 3H, H16), 1.76–1.70 (m, 2H, H18), 1.54 (s, 15H, CH3 Cp*), 1.50–1.42 (m, 2H, H19), 0.98 (t, J = 7.4 Hz, 3H, H20). 13C NMR (125 MHz, CDCl3) δ 165.4 (C14), 163.8 (C15), 158.5 (C1), 158.4 (C10), 145.5 (C11), 141.6 (C5), 130.2 (C12), 129.8 (C9), 128.5 (C6), 128.3 (C7), 124.0 (C2), 122.2 (C8), 118.8 (C3), 116.5 (C13), 112.8 (C4), 93.7 (Cq.Cp*), 40.9 (C16), 40.1 (C17), 30.4 (C18), 20.6 (C19), 14.0 (C20), 9.4 (CH3.Cp*). Anal. calcd. for 4 C 45.74, H 4.35, N 5.21; Found: C 45.65, H 4.30, N 5.09.
X-ray crystal structure determination. A single crystal was selected, mounted and transferred into a cold nitrogen gas stream. Intensity data were collected with a Bruker Kappa-APEX2 system using micro-source Cu-Kα radiation. Unit-cell parameters determination, data collection strategy, integration and absorption correction were carried out with the Bruker APEX2 suite of programs. The structure was solved with SHELXT and refined anisotropically by full-matrix least-squares methods with SHELXL using WinGX. Absolute structure was determined by anomalous scattering effects analysis and chemical absolute configuration was then deduced. The structures were deposited at the Cambridge Crystallographic Data Centre with numbers CCDC 2,145,150 and 2,145,151 which can be obtained free of charge via www.ccdc.cam.ac.uk, accessed on 10 February 2022.

4. Conclusions

In this work, we reported a novel procedure to prepare half-sandwich piano stool complexes of rhodium and iridium. These complexes were resolved using chiral column chromatography into the corresponding enantiomers. The enantiomerization barrier was measured in acetonitrile and ethanol and was up to 124 kJ/mol at 60 °C. These data suggest that our compounds display rigid configuration at the metal center in solution. Our choice to combine a strong chelating (C^C:) ligand with a strong donor/weak-leaving iodide ligand [33,46] generates complexes with strong configuration at the metal center. Due to the presence of an organic chromophore, these complexes might be used to preparer novel luminescent materials or as photocatalysts; the results will be disclosed in due course.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry4010014/s1, Figure S1. 2D ROESY spectrum of 3 recorded in CDCl3; Figure S2. 2D ROESY spectrum of 4 recorded in CDCl3; Resolution of complexes 2, 3 and 4 on chiral stationary phase; UV-vis and CD spectra of complexes 2, 3 and 4; Determination of free energy of enantiomerisation performed in CH3CN and EtOH.

Author Contributions

The synthesis and spectroscopic characterization of the new complexes were performed by A.G. and J.-P.T. The X-ray structural determination was carried out by G.G., M.J. and N.V. performed the chiral separation of the enantiomers and determined the optical rotation and CD spectra. All authors participated in the discussion. Preparation and writing of the manuscript were made by H.A., who also directed the project. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was supported by the CNRS, by the Sorbonne Université, campus Pierre et Marie Curie which we gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. von Zelewsky, A. Stereochemistry of Coordination Compounds; Wiley: Chichester, UK, 1996; p. 254. [Google Scholar]
  2. Constable, E.C. Through a Glass Darkly-Some Thoughts on Symmetry and Chemistry. Symmetry 2021, 13, 1891. [Google Scholar] [CrossRef]
  3. Collet, A.; Crassous, J.; Dutasta, J.P.; Guy, L. Molécules Chirales: Stéreochimie et Propriétes; EDP Sciences: Les Ulis, France, 2006. [Google Scholar]
  4. Amouri, H.; Gruselle, M. Chirality in Transition Metal Chemistry: Molecules, Supramolecular Assemblies and Materials; Wiley: Chichester, UK, 2008. [Google Scholar]
  5. Brunner, H. Optical Activity at an Asymmetrical Manganese Atom. Angew. Chem. Int. Ed. 1969, 8, 382–383. [Google Scholar] [CrossRef]
  6. Brunner, H.; Tsuno, T. Ligand Dissociation: Planar or Pyramidal Intermediates? Acc. Chem. Res. 2009, 42, 1501–1510. [Google Scholar] [CrossRef] [PubMed]
  7. Merrifield, J.H.; Strouse, C.E.; Gladysz, J.A. Synthesis, Optical Resolution, and Absolute-Configuration of Pseudotetrahedral Organorhenium Complexes (Eta-C5H5)Re(NO)(PPh3)(X). Organometallics 1982, 1, 1204–1211. [Google Scholar] [CrossRef]
  8. Mukherjee, T.; Ghosh, S.K.; Wititsuwannakul, T.; Bhuvanesh, N.; Gladysz, J.A. Chiral-at-Metal Ruthenium Complexes with Guanidinobenzimidazole and Pentaphenylcyclopentadienyl Ligands: Synthesis, Resolution, and Preliminary Screening as Enantioselective Second Coordination Sphere Hydrogen Bond Donor Catalysts. Organometallics 2020, 39, 1163–1175. [Google Scholar] [CrossRef]
  9. Mamula, O.; von Zelewsky, A. Supramolecular coordination compounds with chiral pyridine and polypyridine ligands derived from terpenes. Coord. Chem. Rev. 2003, 242, 87–95. [Google Scholar] [CrossRef]
  10. Huang, X.; Meggers, E. Asymmetric photocatalysis with bis-cyclometalated rhodium complexes. Acc. Chem. Res. 2019, 52, 833–847. [Google Scholar] [CrossRef] [PubMed]
  11. Constable, E.C. Stereogenic metal centres-from Werner to supramolecular chemistry. Chem. Soc. Rev. 2013, 42, 1637–1651. [Google Scholar] [CrossRef]
  12. Li, L.P.; Yao, S.Y.; Ou, Y.L.; Wei, L.Q.; Ye, B.H. Diastereoselective Synthesis and Photophysical Properties of Bis-Cyclometalated Ir(III) Stereoisomers with Dual Stereocenters. Organometallics 2017, 36, 3257–3265. [Google Scholar] [CrossRef]
  13. Li, L.P.; Peng, H.L.; Ye, B.H. Thermodynamic Resolution and Enantioselective Synthesis of C-2-Symmetric Bis-sulfoxides Based on Chiral Iridium(III) Complexes. Inorg. Chem. 2019, 58, 12245–12253. [Google Scholar] [CrossRef]
  14. Ehnbom, A.; Ghosh, S.K.; Lewis, K.G.; Gladysz, J.A. Octahedral Werner complexes with substituted ethylenediamine ligands: A stereochemical primer for a historic series of compounds now emerging as a modern family of catalysts. Chem. Soc. Rev. 2016, 45, 6799–6811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Schaffner-Hamann, C.; von Zelewsky, A.; Barbieri, A.; Barigelletti, F.; Muller, G.; Riehl, J.P.; Neels, A. Diastereoselective formation of chiral tris-cyclometalated iridium (III) complexes: Characterization and photophysical properties. J. Am. Chem. Soc. 2004, 126, 9339–9348. [Google Scholar] [CrossRef] [PubMed]
  16. Coughlin, F.J.; Westrol, M.S.; Oyler, K.D.; Byrne, N.; Kraml, C.; Zysman-Colman, E.; Lowry, M.S.; Bernhard, S. Synthesis, separation, and circularly polarized luminescence studies of enantiomers of iridium(III) luminophores. Inorg. Chem. 2008, 47, 2039–2048. [Google Scholar] [CrossRef] [PubMed]
  17. Chepelin, O.; Ujma, J.; Wu, X.; Slawin, A.M.Z.; Pitak, M.B.; Coles, S.J.; Michel, J.; Jones, A.C.; Barran, P.E.; Lusby, P.J. Luminescent, Enantiopure, Phenylatopyridine Iridium-Based Coordination Capsules. J. Am. Chem. Soc. 2012, 134, 19334–19337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Damas, A.; Moussa, J.; Rager, M.N.; Amouri, H. Chiral octahedral bimetallic assemblies with Δ-TRISPHAT as counter anion: Design, anion metathesis, and Cp*Ir as a probe for chiral recognition. Chirality 2010, 22, 889–895. [Google Scholar] [CrossRef] [PubMed]
  19. Groue, A.; Montier-Sorkine, E.; Cheng, Y.P.; Rager, M.N.; Jean, M.; Vanthuyne, N.; Crassous, J.; Lopez, A.C.; Moncada, A.S.; Barbieri, A.; et al. Enantiopure, luminescent, cyclometalated Ir(III) complexes with N-heterocyclic carbene-naphthalimide chromophore: Design, vibrational circular dichroism and TD-DFT calculations. Dalton Trans. 2022, 51, 2750–2759. [Google Scholar] [CrossRef]
  20. Djukic, J.-P.; Hijazi, A.; Flack, H.D.; Bernardinelli, G. Non-racemic (scalemic) planar-chiral five-membered metallacycles: Routes, means, and pitfalls in their synthesis and characterization. Chem. Soc. Rev. 2008, 37, 406–425. [Google Scholar] [CrossRef]
  21. Amouri, H.; Thouvenot, R.; Gruselle, M. Differentiation of planar chiral enantiomers of [Cp*M(2-alkyl-phenoxo)][BF4] {M = Rh, Ir} by the trisphat anion. C. R. Chim. 2002, 5, 257–262. [Google Scholar] [CrossRef]
  22. Amouri, H.; Caspar, R.; Gruselle, M.; Guyard-Duhayon, C.; Boubekeur, K.; Lev, D.A.; Collins, L.S.B.; Grotjahn, D.B. Chiral Recognition and Resolution Mediated by p-p Interactions: Synthesis and X-ray Structure of trans-[(Sp,Sp)-bis(CpRu)-carbazolyl][D-Trisphat]. Organometallics 2004, 23, 4338–4341. [Google Scholar] [CrossRef]
  23. Dubarle-Offner, J.; Axet, M.R.; Chamoreau, L.M.; Amouri, H.; Cooksy, A.L. Enantiomerically Pure, Planar Chiral Cp*Ru Complexes: Synthesis, Molecular Structures, DFT and Coordination Properties. Organometallics 2012, 31, 4429–4434. [Google Scholar] [CrossRef]
  24. Moussa, J.; Chamoreau, L.M.; Amouri, H. Planar Chiral Iridium Complexes with the -TRISPHAT Anion: Toward the First Enantiopure o-Quinone Methide pi-Complex. Chirality 2013, 25, 449–454. [Google Scholar] [CrossRef] [PubMed]
  25. Puig, E.; Gontard, G.; Rager, M.N.; Amouri, H. Optically active Pt-terpyridyl coordination assemblies derived from planar chiral metallothioligands. Inorg. Chim. Acta 2021, 517, 120208. [Google Scholar] [CrossRef]
  26. Brunner, H. Stability of the metal configuration in chiral-at-metal half-sandwich compounds. Eur. J. Inorg. Chem. 2001, 4, 905–912. [Google Scholar] [CrossRef]
  27. Pettinari, C.; Pettinari, R.; Fianchini, M.; Marchetti, F.; Skelton, B.W.; White, A.H. Syntheses, structures, and reactivity of new pentamethylcyclopentadienyl-rhodium(III) and -iridium(III) 4-acyl-5-pyrazolonate complexes. Inorg. Chem. 2005, 44, 7933–7942. [Google Scholar] [CrossRef] [PubMed]
  28. Mimassi, L.; Guyard-Duhayon, C.; Rager, M.N.; Amouri, H. Chiral recognition and resolution of the enantiomers of supramolecular triangular hosts: Synthesis, circular dichroism, NMR, and X-ray molecular structure of [Li subset of(R,R,R)-{Cp*Rh(5-chloro-2,3-dioxopyridine)}(3)][Delta-TRISPHAT]. Inorg. Chem. 2004, 43, 6644–6649. [Google Scholar] [CrossRef] [PubMed]
  29. Feghali, E.; Barloy, L.; Issenhuth, J.T.; Karmazin-Brelot, L.; Bailly, C.; Pfeffer, M. Cyclometalation of (2R,5R)-2,5-Diphenylpyrrolidine and 2-Phenyl-2imidazoline Ligands with Half-Sandwich Iridium(III) and Rhodium(III) Complexes. Organometallics 2013, 32, 6186–6194. [Google Scholar] [CrossRef]
  30. Corberan, R.; Sanau, M.; Peris, E. Highly stable Cp*-Ir(III) complexes with N-heterocyclic carbene ligands as C-H activation catalysts for the deuteration of organic molecules. J. Am. Chem. Soc. 2006, 128, 3974–3979. [Google Scholar] [CrossRef]
  31. Hellou, N.; Jahier-Diallo, C.; Basle, O.; Srebro-Hooper, M.; Toupet, L.; Roisnel, T.; Caytan, E.; Roussel, C.; Vanthuyne, N.; Autschbach, J.; et al. Electronic and chiroptical properties of chiral cycloiridiated complexes bearing helicenic NHC ligands. Chem. Commun. 2016, 52, 9243–9246. [Google Scholar] [CrossRef] [Green Version]
  32. Groue, A.; Tranchier, J.-P.; Rager, M.-N.; Gontard, G.; Jean, M.; Vanthuyne, N.; Pearce, H.R.; Cooksy, A.L.; Amouri, H. Unique Class of Enantiopure N-Heterocyclic Carbene Half-Sandwich Iridium(III) Complexes with Stable Configurations: Probing Five-Membered versus Six-Membered Iridacycles. Inorg. Chem. 2019, 58, 2930–2933. [Google Scholar] [CrossRef]
  33. Fagnou, K.; Lautens, M. Halide effects in transition metal catalysis. Angew. Chem. Int. Ed. 2002, 41, 26–47. [Google Scholar] [CrossRef]
  34. Stanley, K.; Baird, M.C. Demonstration of controlled asymmetric induction in organoiron chemistry. Suggestions concerning the specification of chirality in pseudotetrahedral metal complexes containing polyhapto ligands. J. Am. Chem. Soc. 1975, 97, 6598–6599. [Google Scholar] [CrossRef]
  35. Merrifield, J.H.; Fernandez, J.M.; Buhro, W.E.; Gladysz, J.A. Cleavage of the Rhenium Methyl Bond of (Eta-C5h5)Re(No)(Pph3)(Ch3) by Protic and Halogen Electrophiles-Stereochemistry at Rhenium. Inorg. Chem. 1984, 23, 4022–4029. [Google Scholar] [CrossRef]
  36. Brunner, H.; Kollnberger, A.; Burgemeister, T.; Zabel, M. Optically active transition metal complexes-Part 125. Preparation and epimerization of chiral-at-metal pentamethylcyclopentadienyl-rhodium(III) and iridium(III) half-sandwich complexes. Polyhedron 2000, 19, 1519–1526. [Google Scholar] [CrossRef]
  37. Liu, Z.; Habtemariam, A.; Pizarro, A.M.; Fletcher, S.A.; Kisova, A.; Vrana, O.; Salassa, L.; Bruijnincx, P.C.A.; Clarkson, G.J.; Brabec, V.; et al. Organometallic half-sandwich iridium anticancer complexes. J. Med. Chem. 2011, 54, 3011–3026. [Google Scholar] [CrossRef] [PubMed]
  38. Liu, Z.; Sadler, P.J. Organoiridium Complexes: Anticancer Agents and Catalysts. Acc. Chem. Res. 2014, 47, 1174–1185. [Google Scholar] [CrossRef] [PubMed]
  39. Grotjahn, D.B.; Brown, D.B.; Martin, J.K.; Marelius, D.C.; Abadjian, M.-C.; Tran, H.N.; Kalyuzhny, G.; Vecchio, K.S.; Specht, Z.G.; Cortes-Llamas, S.A.; et al. Evolution of Iridium-Based Molecular Catalysts during Water Oxidation with Ceric Ammonium Nitrate. J. Am. Chem. Soc. 2011, 133, 19024–19027. [Google Scholar] [CrossRef] [PubMed]
  40. Schulze, M.; Steffen, A.; Wurthner, F. Near-IR Phosphorescent Ruthenium(II) and Iridium(III) Perylene Bisimide Metal Complexes. Angew. Chem. Int. Ed. 2015, 54, 1570–1573. [Google Scholar] [CrossRef]
  41. Specht, Z.G.; Cortes-Llamas, S.A.; Tran, H.N.; van Niekerk, C.J.; Rancudo, K.T.; Golen, J.A.; Moore, C.E.; Rheingold, A.L.; Dwyer, T.J.; Grotjahn, D.B. Enabling Bifunctionality and Hemilability of N-Heteroaryl NHC Complexes. Chem. Eur. J. 2011, 17, 6606–6609. [Google Scholar] [CrossRef]
  42. Gomez-Lopez, J.L.; Chavez, D.; Parra-Hake, M.; Royappa, A.T.; Rheingold, A.L.; Grotjahn, D.B.; Miranda-Soto, V. Synthesis and reactivity of bis(protic N-heterocyclic carbene)iridium(III) complexes. Organometallics 2016, 35, 3148–3153. [Google Scholar] [CrossRef]
  43. Peris, E. Smart N-Heterocyclic Carbene Ligands in Catalysis. Chem. Rev. 2018, 118, 9988–10031. [Google Scholar] [CrossRef]
  44. Thongpaen, J.; Schmid, T.E.; Toupet, L.; Dorcet, V.; Mauduit, M.; Basle, O. Directed ortho C-H borylation catalyzed using Cp*Rh(iii)-NHC complexes. Chem. Commun. 2018, 54, 8202–8205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Shibata, T.; Hashimoto, H.; Kinoshita, I.; Yano, S.; Nishioka, T. Unprecedented diastereoselective generation of chiral-at-metal, half sandwich Ir(III) and Rh(III) complexes via anomeric isomerism on “sugar-coated” N-heterocyclic carbene ligands. Dalton Trans. 2011, 40, 4826–4829. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, W.Y.; Banerjee, S.; Hughes, G.M.; Bridgewater, H.E.; Song, J.I.; Breeze, B.; Clarkson, G.J.; Coverdale, J.P.C.; Sanchez-Cano, C.; Ponte, F.; et al. Ligand-centred redox activation of inert organoiridium anticancer catalysts. Chem. Sci. 2020, 11, 5466–5480. [Google Scholar] [CrossRef] [PubMed]
  47. Lanoe, P.H.; Chan, J.; Gontard, G.; Monti, F.; Armaroli, N.; Barbieri, A.; Amouri, H. Deep-Red Phosphorescent Iridium(III) Complexes with Chromophoric N-Heterocyclic Carbene Ligands: Design, Photophysical Properties, and DFT Calculations. Eur. J. Inorg. Chem. 2016, 2016, 1631–1634. [Google Scholar] [CrossRef]
  48. Lanoe, P.H.; Najjari, B.; Hallez, F.; Gontard, G.; Amouri, H. N-Heterocyclic Carbene Coinage Metal Complexes Containing Naphthalimide Chromophore: Design, Structure, and Photophysical Properties. Inorganics 2017, 5, 58. [Google Scholar] [CrossRef] [Green Version]
  49. Lanoe, P.H.; Chan, J.; Groue, A.; Gontard, G.; Jutand, A.; Rager, M.N.; Armaroli, N.; Monti, F.; Barbieri, A.; Amouri, H. Cyclometalated N-heterocyclic carbene iridium(III) complexes with naphthalimide chromophores: A novel class of phosphorescent heteroleptic compounds. Dalton Trans. 2018, 47, 3440–3451. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Preparation of the cyclometalated half-sandwich complexes 24.
Scheme 1. Preparation of the cyclometalated half-sandwich complexes 24.
Chemistry 04 00014 sch001
Figure 1. Optically active molecules (S)-2, (S)-3 and (S)-4/(R)-2, (R)-3 and (R)-4 described in this work and priority rule used (I > Cp > C-Carbene > C-arene) to assign the absolute configuration [34,35].
Figure 1. Optically active molecules (S)-2, (S)-3 and (S)-4/(R)-2, (R)-3 and (R)-4 described in this work and priority rule used (I > Cp > C-Carbene > C-arene) to assign the absolute configuration [34,35].
Chemistry 04 00014 g001
Figure 2. CD spectra of (S)-2 (green) and (R)-2 (red) recorded in CH3CN at 0.251 mM.
Figure 2. CD spectra of (S)-2 (green) and (R)-2 (red) recorded in CH3CN at 0.251 mM.
Chemistry 04 00014 g002
Figure 3. CD spectra of S-3 (green) and R-3 (red) recorded in CH3CN at 0.24 mM.
Figure 3. CD spectra of S-3 (green) and R-3 (red) recorded in CH3CN at 0.24 mM.
Chemistry 04 00014 g003
Figure 4. CD spectra of (S)-4 (green) and (R)-4 (red) recorded in CH3CN at 0.24 mM.
Figure 4. CD spectra of (S)-4 (green) and (R)-4 (red) recorded in CH3CN at 0.24 mM.
Chemistry 04 00014 g004
Figure 5. Molecular structure of (S)-2 with thermal ellipsoids drawn at the 30% probability level, hydrogen atoms were omitted for clarity. The asymmetric unit comprises two molecules with the same configuration and almost identical conformations, differing only in their nBu-groups. One molecule was arbitrarily chosen for representation. Selected average bond distances (Å) and angles (deg): Rh1-I1 2.699(2), Rh1-C1 1.983(5), C1-N1 1.371(7), N1-C5 1.414(6), C5-C10 1.426(7), C10-C11 1.427(7), C11-Rh1 2.026(8), Rh1-C21 2.154(5), Rh1-C22 2.248(8), Rh1-C23 2.289(5), Rh1-C24 2.276(5), Rh1-C25 2.248(5), C1-Rh1-C11 86.5(3), Rh1-C1-N1 126.0(6), Rh1-C11-C10 122.2(12), C1-N1-C5 125.8(4), C11-C10-C5 123.9(5), and N1-C5-C10 120.0(4).
Figure 5. Molecular structure of (S)-2 with thermal ellipsoids drawn at the 30% probability level, hydrogen atoms were omitted for clarity. The asymmetric unit comprises two molecules with the same configuration and almost identical conformations, differing only in their nBu-groups. One molecule was arbitrarily chosen for representation. Selected average bond distances (Å) and angles (deg): Rh1-I1 2.699(2), Rh1-C1 1.983(5), C1-N1 1.371(7), N1-C5 1.414(6), C5-C10 1.426(7), C10-C11 1.427(7), C11-Rh1 2.026(8), Rh1-C21 2.154(5), Rh1-C22 2.248(8), Rh1-C23 2.289(5), Rh1-C24 2.276(5), Rh1-C25 2.248(5), C1-Rh1-C11 86.5(3), Rh1-C1-N1 126.0(6), Rh1-C11-C10 122.2(12), C1-N1-C5 125.8(4), C11-C10-C5 123.9(5), and N1-C5-C10 120.0(4).
Chemistry 04 00014 g005
Figure 6. Molecular structure of (S)-4 with thermal ellipsoids drawn at the 30% probability level, hydrogen atoms were omitted for clarity. Selected bond distances (Å) and angles (deg): Ir1-I1 2.702(1), Ir1-C1 1.982(5), C1-N1 1.367(6), N1-C5 1.430(6), C5-C10 1.422(7), C10-C11 1.422(7), C11-Ir1 2.053(4), Ir1-C21 2.168(5), Ir1-C22 2.230(5), Ir1-C23 2.271(4), Ir1-C24 2.279(4), Ir1-C25 2.246(4), C1-Ir1-C11 86.6(2), Ir1-C1-N1 125.1(3), Ir1-C11-C10 121.0(4), C1-N1-C5 125.8(4), C11-C10-C5 123.8(4), and N1-C5-C10 119.7(4).
Figure 6. Molecular structure of (S)-4 with thermal ellipsoids drawn at the 30% probability level, hydrogen atoms were omitted for clarity. Selected bond distances (Å) and angles (deg): Ir1-I1 2.702(1), Ir1-C1 1.982(5), C1-N1 1.367(6), N1-C5 1.430(6), C5-C10 1.422(7), C10-C11 1.422(7), C11-Ir1 2.053(4), Ir1-C21 2.168(5), Ir1-C22 2.230(5), Ir1-C23 2.271(4), Ir1-C24 2.279(4), Ir1-C25 2.246(4), C1-Ir1-C11 86.6(2), Ir1-C1-N1 125.1(3), Ir1-C11-C10 121.0(4), C1-N1-C5 125.8(4), C11-C10-C5 123.8(4), and N1-C5-C10 119.7(4).
Chemistry 04 00014 g006
Table 1. Summary of crystallographic, data collection and refinement parameters for (S)-[Cp*Rh(C^C:)I] (S)-2 and (S)-[Cp*Ir(C^C:)I] (S)-4.
Table 1. Summary of crystallographic, data collection and refinement parameters for (S)-[Cp*Rh(C^C:)I] (S)-2 and (S)-[Cp*Ir(C^C:)I] (S)-4.
Compound(S)-[Cp*Rh(C^C:)I] (S)-2(S)-[Cp*Ir(C^C:)I] (S)-4
Empirical formulaC30 H33 I RhN3 O2 C30 H33 I Ir N3 O2
Formula weight697.40786.69
Crystal systemMonoclinicOrthorhombic
Space groupP 21P 21 21 21
Unit cell dimensionsa = 17.2315(4) Åa = 8.0943(2) Å
b = 10.6098(2) Åb = 16.2456(4) Å
c = 17.2604(4) Åc = 21.3531(5) Å
α = 90°α = 90°
β = 119.616(1)°β = 90°
γ = 90°γ = 90°
Volume2743.34(11) Å32807.86(12) Å3
Z44
Temperature200(1) K200(1) K
Wavelength1.54178 Å1.54178 Å
θ range for data collection5.10° to 66.62°4.96° to 66.64°
Reflections (all/independent)34,037/968915,376/4959
R(int)2.54%2.28%
Data/parameters/restraints4959/340/09689/679/1
R1 [I > 2σ(I)]2.08%1.79%
wR2 (all data)5.31%4.55%
Flack parameter−0.015(2)0.001(3)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Groué, A.; Tranchier, J.-P.; Gontard, G.; Jean, M.; Vanthuyne, N.; Amouri, H. Enantiopure Cyclometalated Rh(III) and Ir(III) Complexes Displaying Rigid Configuration at Metal Center: Design, Structures, Chiroptical Properties and Role of the Iodide Ligand. Chemistry 2022, 4, 156-167. https://doi.org/10.3390/chemistry4010014

AMA Style

Groué A, Tranchier J-P, Gontard G, Jean M, Vanthuyne N, Amouri H. Enantiopure Cyclometalated Rh(III) and Ir(III) Complexes Displaying Rigid Configuration at Metal Center: Design, Structures, Chiroptical Properties and Role of the Iodide Ligand. Chemistry. 2022; 4(1):156-167. https://doi.org/10.3390/chemistry4010014

Chicago/Turabian Style

Groué, Antoine, Jean-Philippe Tranchier, Geoffrey Gontard, Marion Jean, Nicolas Vanthuyne, and Hani Amouri. 2022. "Enantiopure Cyclometalated Rh(III) and Ir(III) Complexes Displaying Rigid Configuration at Metal Center: Design, Structures, Chiroptical Properties and Role of the Iodide Ligand" Chemistry 4, no. 1: 156-167. https://doi.org/10.3390/chemistry4010014

Article Metrics

Back to TopTop