Next Article in Journal
Experimental Study of CO2 Conversion into Methanol by Synthesized Photocatalyst (ZnFe2O4/TiO2) Using Visible Light as an Energy Source
Next Article in Special Issue
Photophysics and Photochemistry of Iron Carbene Complexes for Solar Energy Conversion and Photocatalysis
Previous Article in Journal
Recent Advances on the Rational Design of Non-Precious Metal Oxide Catalysts Exemplified by CuOx/CeO2 Binary System: Implications of Size, Shape and Electronic Effects on Intrinsic Reactivity and Metal-Support Interactions
Previous Article in Special Issue
Design and Synthesis of Photoactive Iron N-Heterocyclic Carbene Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Natural Product Lepidiline A as an N-Heterocyclic Carbene Ligand Precursor in Complexes of the Type [Ir(cod)(NHC)PPh3)]X: Synthesis, Characterisation, and Application in Hydrogen Isotope Exchange Catalysis

Department of Pure and Applied Chemistry, WestCHEM, University of Strathclyde, Glasgow G1 1XL, Scotland, UK
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(2), 161; https://doi.org/10.3390/catal10020161
Submission received: 20 December 2019 / Revised: 23 January 2020 / Accepted: 26 January 2020 / Published: 1 February 2020
(This article belongs to the Special Issue N‐Heterocyclic Carbenes and Their Complexes in Catalysis)

Abstract

:
A range of iridium(I) complexes of the type [Ir(cod)(NHC)PPh3)]X are reported, where the N-heterocyclic carbene (NHC) is derived from the naturally-occurring imidaozlium salt, Lepidiline A (1,3-dibenzyl-4,5-dimethylimidazolium chloride). A range of complexes were prepared, with a number of NHC ligands and counter-ions, and various steric and electronic parameters of these complexes were evaluated. The activity of the [Ir(cod)(NHC)PPh3)]X complexes in hydrogen isotope exchange reactions was then studied, and compared to established iridium(I) complexes.

Graphical Abstract

1. Introduction

Isotopically labelled compounds are of importance in a range of scientific disciplines. In particular, the isotopes of hydrogen, deuterium, and tritium find application in several key areas. Within the physical sciences, deuterated compounds are routinely used in the elucidation of reaction mechanisms [1]. In the life sciences, deuterated and tritiated compounds are used to determine the pharmacokinetic properties of bioactive molecules [2,3,4]. More recently, active pharmaceutical ingredients themselves have begun to feature deuterium, exploiting the effect of relative C-H/C-D bond strengths to deliver drug molecules with improved metabolic profiles [5,6].
Amongst the synthetic methods available for incorporation of deuterium or tritium into a molecule, hydrogen isotope exchange (HIE) stands out due to its simplicity and generality [7,8,9]. A particularly powerful variant of this process is directed HIE, whereby a Lewis basic group within the molecule directs a metal catalyst to an adjacent site, and a subsequent C-H activation results in a metallocyclic intermediate which facilitates HIE at this adjacent position (Scheme 1).
Iridium(I) complexes are amongst the most effective metal catalysts for directed HIE [10] and, in recent years, we have introduced a suite of highly active Ir(I) catalysts bearing a combination of bulky phosphine and N-heterocyclic carbene ligands (Figure 1). These catalysts are compatible with a broad variety of directing groups, enabling HIE on a wide range of substrates [11,12,13,14,15,16,17,18].
These benchmark complexes 1–2 feature three distinct phosphine ligands. However, our early studies [11,12], which focused on identifying the most effective catalyst system, evaluated a broader range of phosphines based on their steric and electronic properties. In contrast, although we have optimised the NHC ligand for some specific HIE use cases [19,20], there remains scope for a more systematic study of the steric and electronic parameters of this ligand component. The natural product Lepidiline A 3·HCl is an imidazolium chloride salt, featuring two N-benzyl substituents, and methyl groups at the heterocycle’s 4- and 5-positions [21] (Figure 2). We envisaged that this naturally-occurring imidazolium salt could serve as a vehicle to begin to explore the effect of steric and electronic changes in the NHC on the activity of our Ir(I) NHC/phosphine catalysts [22].
Herein we report a study of the complexes 1a, 6a and 7, of the type [Ir(cod)(NHC)PPh3)]PF6 (cod = 1,5-cyclooctadiene), where the NHC ligand is IMes 4, IBn 5, or IBnMe 3, derived from Lepidiline A 3·HCl (Figure 3). These complexes are then used to explore the effect of steric and electronic variance within the NHC through comparison of (a) N-benzyl vs. N-mesityl substituents; and (b) 4,5-H2 vs. 4,5-Me2 substitution, on the reactivity in HIE processes. The effect of the counter on the activity of the IBnMe complexes is also investigated.

2. Results and Discussion

To initiate our studies, we set out to prepare quantities of the desired [Ir(cod)(NHC)PPh3)]PF6 complexes 1a, 6a and 7. Complex 1a was prepared by a modification of previously reported conditions [11,23]. The IBn-containing complex 7 was prepared [23] by using an analogous procedure from the previously reported complex [Ir(cod)(IBn)Cl] [24]. Lastly, the imidazolium chloride salt 3·HCl, representing the natural product Lepidiline A, was targeted to deliver complex 6a.

2.1. Synthesis of Lepidiline A

Surprisingly, to our knowledge, no preparation of this naturally-occurring material has yet been reported. Our synthetic route is summarised in Scheme 2. Chlorination of the commercially available hydroxymethyl imidazolium salt 8 in neat thionyl chloride was easily achieved in excellent yield on a multi-gram scale to deliver 9. Reduction and basification of 9 yielded the free 4,5-dimethylimidazole 10 with good levels of efficiency. Subsequently, double-alkylation of 10 under basic conditions could be achieved using benzyl chloride to give Lepidiline A 3·HCl with good effectiveness, or with benzyl bromide to give 3·HBr in excellent yield. Imidazolium bromide 3·HBr could then undergo salt metathesis to very efficiently afford 3·HPF6 and 3·HBArF.
Following the successful preparation of Lepidiline A 3·HCl and a range of its counterion derivatives, X-ray quality crystals of 3·HCl were grown and the structure solved to reveal a molecular footprint in direct agreement with the spectroscopic data and the reported natural product structure. The crystal structure was found to be a solvate containing one molecule of water per cation (Figure 4, left). The same cationic motif was found in the X-ray structure of 3·HPF6. This latter structure has two crystallographically independent salt units per asymmetric unit (Z’ = 2). Both cations in this structure adopt anti conformations of the benzyl rings with respect to the heteroaromatic ring (Figure 4, right). This is in contrast to the syn conformation found for 3·HCl and in the Ir complexes described below.

2.2. Preparation of Iridium Complexes Based on Lepidiline A 3

With a range of imidazolium salt precursors in hand, formation of the corresponding iridium complexes was investigated. Based on work by Wang and Lin [25], and Liu et al. [26], the [(cod)Ir(NHC)Cl] complexes can be prepared via a transmetallation procedure from the corresponding silver bis-carbene complex. Employing the bromide salt 3·HBr, mixing with silver(I) oxide and sodium iodide in dichloromethane gave silver complex 11 in high yield (Scheme 3). This intermediate could be taken forward into a transmetallation procedure with [(cod)IrCl]2 13 to provide the desired chloro-carbene complex 12 in excellent yield following a short reaction time. Alternatively, it was also found that isolation of 11 could be circumvented via direct addition of [(cod)IrCl]2 13 following in situ formation of the silver complex 11, delivering 12 in a further improved overall yield. With this high-yielding method in place, X-ray quality crystals of 12 were grown, confirming the proposed structure (Figure 5).
With quantities of the NHC chloride complex 12 secured, treatment with triphenylphosphine, followed by silver hexafluorophosphate, delivered target NHC-phosphine complex 6a in a good 75% yield (Table 1, entry 1). In order to probe the effects of varying the counter-ion in these novel complexes, derivatives 6b6d were also prepared in an analogous manner and with excellent isolated yields (Table 1, entries 2−4).
We have previously shown that iridium NHC/phosphine complexes bearing a BArF (tetrakis [3,5-bis(trifluoromethyl)phenyl]borate) counter-ion have a broader solvent applicability than the corresponding PF6 complexes [13]. Based on this, BArF-containing complex 6e was prepared in a one-pot procedure from [(cod)IrCl]2 13 in a good 82% yield (Scheme 4). Complexes 6a6e were all characterized by X-ray crystallography [23].

2.3. Steric and Electronic Characterisation of the Iridium NHC/Phosphine Complexes

Having successfully prepared a range of Ir complexes containing the Lepidiline A-derived NHC 3, attention now turned to evaluating this novel ligand sphere versus the alternative ligand sets in PPh3/IMes complex 1a and PPh3/IBn complex 7. The electronic properties of ligands in a specific complex are often characterised by carbonylation of the complex, whereby the νCO stretching frequency reports on the electron-donating nature of the other ligands on the metal [27]. Recent work on Ir(I) dicarbonyl complexes has revealed that the configuration of the carbonyl ligands, though preferentially cis, is dictated by the ligand size, with larger combinations driving a trans dicarbonyl arrangement [28]. This structural fluidity limits the use of electronic parameters derived from such complexes [28]. However, the synthesis and X-ray characterisation of the carbonlyated analogues of our NHC/phosphine complexes could instead be used to indicate a steric threshold—the ligand size at which Ir(I) carbonyl complexes switch from a cis- to a trans-configuration, and this could be correlated with the well-known steric parameter, percentage buried volume, %Vbur, [29] of the ligands.
Accordingly, we investigated the carbonylation of complexes 1a, 6a, and 7 (Scheme 5). Direct exposure of complexes 6a and 7 to an atmosphere of CO resulted in complexes 14 and 15, where both the NHC/phosphine pair and the two CO ligands are each cis-orientated, as shown by the X-ray crystal structures [23]. The combined ligand %Vbur values (Σ(%Vbur)) were similar for both complexes, as determined by density functional theory (DFT) calculations (Table 2) [23]. As perhaps expected, the enhanced bulk of the NHC ligand in 1a resulted in a complex, 16, with a trans-arrangement of the NHC/P pair and of the two CO ligands, indicating the size and geometry of this complex to be quite distinct from the IBn and IBnMe analogues 14 and 15. Based on these parameters, there is little steric difference evident on moving from the IBn to the IBnMe ligand, with the latter appearing to be a very slightly bulkier system, based on the %Vbur values.
Although these dicarbonyl complexes permitted only a steric, and not electronic, evaluation, literature examples of Ir(I) chloro-carbonyl complexes have demonstrated that the combined influence of various ligand partnerships on the metal’s π-donating ability can be measured via IR spectroscopy, based on the carbonyl stretching frequency, νCO [26,30]. Furthermore, the effect of the ligands on the σ-donating nature of the complex can be measured by 1H nuclear magnetic resonance (NMR) spectroscopic evaluation of the corresponding dihydride complexes [31]. The electronic characteristics of the IMes, IBn, and IBnMe ligands in our complexes were thus explored by preparing the chloro-carbonyl complexes, 17–19 (Table 3). Additionally, from the NHC/PPh3 complexes, in situ generation of the dihydride complexes 20–22, in CD3CN [23], provided insight into the σ-donating properties (Table 3). Based on both electronic parameters, the IBnMe/PPh3 combination was found to be slightly more electron-rich than IBn/PPh3,, and more significantly less electron-rich than IMes/PPh3.

2.4. Evaluation of Iridium Complex Catalytic Activity in HIE

We next chose to focus on the relationship between the IBn/PPh3 and IBnMe/PPh3 complex pair in terms of catalyst activity in HIE reactions. The two precatalysts 6a and 7 were compared via kinetic analysis of the simple reference labelling reaction with acetophenone 23 (Table 4). Catalyst 6a, derived from Lepidiline A, was clearly faster reacting, displaying a pseudo first order rate constant, kobs(6a) = 0.0224 s−1, versus kobs(7) = 0.0145 s−1 for IBn/PPh3 catalyst 7 [23].
This result was supported by DFT modelling of the turnover-limiting [12] C-H activation step (2426, Scheme 6). Based on our electronic characterization of these complexes, the greater reactivity of catalyst 6a over 7 can be rationalised on the basis of an increased π-donating ability of the Ir centre, and not on any appreciable steric difference between the two catalysts 6a and 7.
The variation in reactivity between the IBnMe/PPh3 ligand sphere in 6a and IMes/PPh3 in 1a was most interestingly drawn out in studying the counterion dependence on catalyst reactivity (Table 5). Even at a relatively high and unoptimised catalyst loading of 5 mol%, larger and more weakly coordinating substrates 23, 27, and 28 perform relatively poorly when catalyst 6d, bearing the OTf counter-ion, is used in the HIE reaction. Alternatively, the more strongly binding substrate, 29, is relatively unaffected, and the order of reactivity generally reflects that previously observed when evaluating counter-ion effects with the IMes/PPh3 combination [13]. Presumably, the smaller size of the IBnMe/PPh3 system suffers from competitive coordination of the triflate to the metal centre [32]. Most curiously, the catalyst series 6 is ineffective at labelling the weakly coordinating substrate, nitrobenzene 30, whereas the flagship IMes/PPh3 system 1a labels to ~95% D under similar conditions [11]. This may be due to differences in available intermolecular (for example π-π) interactions of the substrate to the different catalyst systems upon binding. Nonetheless, such effects have the potential to be exploited favourably to induce directing group chemoselectivity in labelling multifunctional molecules.

3. Materials and Methods

See the Supplementary Information for details of all experimental and computational procedures, as well as all crystallographic data.

4. Conclusions

In summary, the first synthesis of the simple imidazolium-based natural product, Lepidiline A, 3·HCl, has been reported, along with a range of other counter-ion derivatives. This has been followed by the development of novel HIE catalysts and the use of a silver-NHC transmetallation approach to HIE catalyst synthesis. Detailed and combined experimental/computational analysis of the resulting iridium complexes 6 and 7 has led to quantitative insight into the parameters responsible for observed differences in the resulting catalytic reactivity of 6a versus the less substituted analogue, 7, and existing flagship HIE catalyst, 1a. Overall, dimethyl substitution on the backbone of the novel NHC was found to impart an electronic influence with a minimal impact on the steric environment. The reactivity differences between the catalysts studied are now being exploited in labelling regioselectivity studies, which will be reported in due course.

Supplementary Materials

All experimental procedures, computational calculations and X-ray structure data are available online at https://www.mdpi.com/2073-4344/10/2/161/s1: all experimental procedures along with corresponding compound characterisation; computational calculations for %Vbur in complexes 1416 and C-H activation of acetophenone with catalysts 6a and 7; X-ray crystallographic parameters for compounds 3·HCl, 3·HPF6, 12, 14–16, and 6a6e; and copies of all NMR spectra.

Author Contributions

Conceptualization, W.J.K. and T.T.; methodology, A.R.C. and M.R.; formal analysis, A.R.C., M.R. and A.R.K.; data curation, A.R.K and D.M.L.; writing—original draft preparation, M.R. and D.M.L.; writing—review and editing, D.M.L. and W.J.K.; supervision, W.J.K; funding acquisition, W.J.K. All authors have read and agreed to the published version of the manuscript.

Funding

We thank the Carnegie Trust (M.R.) for postgraduate studentship funding.

Acknowledgments

We acknowledge the EPSRC UK National Mass Spectrometry Facility at Swansea University for mass spectrometry analyses.

Conflicts of Interest

The authors declare no conflict of interest.

References and Note

  1. Meyer, M.P. New Applications of Isotope Effects in the Determination of Organic Reaction Mechanisms. In Advances in Physical Organic Chemistry; Williams, I., Williams, N., Eds.; Elsevier: London, UK, 2012; Volume 46, pp. 57–114. [Google Scholar]
  2. Isin, E.M.; Elmore, C.S.; Nilsson, G.N.; Thompson, R.A.; Weidolf, L. Use of radiolabeled compounds in drug metabolism and pharmacokinetic studies. Chem. Res. Toxicol. 2012, 25, 532–542. [Google Scholar] [CrossRef] [PubMed]
  3. Penner, N.; Xu, L.; Prakash, C. Radiolabeled absorption, distribution, metabolism, and excretion studies in drug development: why, when and how? Chem. Res. Toxicol. 2012, 25, 513–531. [Google Scholar] [CrossRef] [PubMed]
  4. Atzrodt, J.; Derdau, V.; Kerr, W.J.; Reid, M. Deuterium- and tritium-labelled compounds: Applications in the life sciences. Angew. Chem. Int. Ed. 2018, 57, 1758–1784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Schillerstrom, R. Heavy Drugs. Drug Discov. Dev. 2009, 12, 6–8. [Google Scholar]
  6. Katsnelson, A. Heavy drugs draw heavy interest from pharma backers. Nat. Med. 2013, 19, 656. [Google Scholar] [CrossRef]
  7. Atzrodt, J.; Derdau, V.; Kerr, W.J.; Reid, M. C-H functionalization for hydrogen isotope exchange. Angew. Chem. Int. Ed. 2018, 57, 3022–3047. [Google Scholar] [CrossRef]
  8. Hesk, D. Highlights of C(sp2)-H hydrogen isotope exchange reactions. J. Labelled Compd. Radiopharm. 2019. [Google Scholar] [CrossRef]
  9. Valero, M.; Derdau, V. Highlights of aliphatic C(sp3)-H hydrogen isotope exchange reactions. J. Labelled Compd. Radiopharm. 2019. [Google Scholar] [CrossRef]
  10. Kerr, W.J.; Knox, G.J.; Paterson, L.C. Recent Advances in Iridium(I) Catalysis towards Directed Hydrogen Isotope Exchange. J. Labelled Compd. Radiopharm. 2019. [Google Scholar] [CrossRef]
  11. Brown, J.A.; Irvine, S.; Kennedy, A.R.; Kerr, W.J.; Andersson, S.; Nilsson, G.N. Highly Active Iridium(I) Complexes for Catalytic Hydrogen Isotope Exchange. Chem. Commun. 2008, 1115–1117. [Google Scholar] [CrossRef] [Green Version]
  12. Brown, J.A.; Cochrane, A.R.; Irvine, S.; Kerr, W.J.; Mondal, B.; Parkinson, J.A.; Paterson, L.C.; Reid, M.; Tuttle, T.; Andersson, S.; et al. The Synthesis of Highly Active Iridium(I) Complexes and Their Application in Catalytic Hydrogen Isotope Exchange. Adv. Synth. Catal. 2014, 356, 3551–3562. [Google Scholar] [CrossRef] [Green Version]
  13. Kennedy, A.R.; Kerr, W.J.; Moir, R.; Reid, M. Anion Effects to Deliver Enhanced Iridium Catalysts for Hydrogen Isotope Exchange Processes. Org. Biomol. Chem. 2014, 12, 7927–7931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kerr, W.J.; Mudd, R.J.; Paterson, L.C.; Brown, J.A. Iridium(I)-Catalyzed Regioselective C-H Activation and Hydrogen-Isotope Exchange of Non-Aromatic Unsaturated Functionality. Chem. Eur. J. 2014, 20, 14604–14607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Atzrodt, J.; Derdau, V.; Kerr, W.J.; Reid, M.; Rojahn, P.; Weck, R. Expanded Applicability of Iridium(I) NHC/Phosphine Catalysts in Hydrogen Isotope Exchange Processes with Pharmaceutically-Relevant Heterocycles. Tetrahedron 2015, 71, 1924–1929. [Google Scholar] [CrossRef] [Green Version]
  16. Kerr, W.J.; Lindsay, D.M.; Reid, M.; Atzrodt, J.; Derdau, V.; Rojahn, P.; Weck, R. Iridium-Catalysed Ortho-H/D and -H/T Exchange under Basic Conditions: C–H Activation of Unprotected Tetrazoles. Chem. Commun. 2016, 52, 6669–6672. [Google Scholar] [CrossRef] [Green Version]
  17. Kerr, W.J.; Mudd, R.J.; Owens, P.K.; Reid, M.; Brown, J.A.; Campos, S. Hydrogen Isotope Exchange with Highly Active Iridium(I) NHC/Phosphine Complexes: A Comparative Counterion Study. J. Labelled Compd. Radiopharm. 2016, 59, 601–603. [Google Scholar] [CrossRef] [Green Version]
  18. Kerr, W.J.; Lindsay, D.M.; Owens, P.K.; Reid, M.; Tuttle, T.; Campos, S. Site-Selective Deuteration of N-Heterocycles via Iridium-Catalyzed Hydrogen Isotope Exchange. ACS Catal. 2017, 7, 7182–7186. [Google Scholar] [CrossRef] [Green Version]
  19. Kerr, W.J.; Reid, M.; Tuttle, T. Iridium-Catalyzed C-H Activation and Deuteration of Primary Sulfonamides: An Experimental and Computational Study. ACS Catal. 2015, 5, 402–410. [Google Scholar] [CrossRef] [Green Version]
  20. Kerr, W.J.; Reid, M.; Tuttle, T. Iridium-Catalyzed Formyl-Selective Deuteration of Aldehydes. Angew. Chem. Int. Ed. 2017, 56, 7808–7812. [Google Scholar] [CrossRef]
  21. Cui, B.; Zheng, B.; He, K.; Zheng, Q. Imidazole Alkaloids from Lepidium meyenii. J. Nat. Prod. 2003, 66, 1101–1103. [Google Scholar] [CrossRef]
  22. Curran, D.; Dada, O.; Müller-Bunz, H.; Rothemund, M.; Sánchez-Sanz, G.; Schobert, R.; Zhu, X.; Tacke, M. Synthesis and Cytotoxic Studies of Novel NHC*-Gold(I) Complexes Derived from Lepidiline A. Molecules 2018, 23, 2031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. See the Supplementary Materials for further details.
  24. Cochrane, A.R.; Irvine, S.; Kerr, W.J.; Reid, M.; Andersson, S.; Nilsson, G.N. Application of neutral iridium(I) N-heterocyclic carbene complexes in ortho-directed hydrogen isotope exchange. J. Labelled Compd. Radiopharm. 2013, 46, 451–454. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, H.M.J.; Lin, I.J.B. Facile Synthesis of Silver(I)−Carbene Complexes. Useful Carbene Transfer Agents. Organometallics 1998, 17, 972–975. [Google Scholar] [CrossRef]
  26. Chang, Y.-H.; Fu, C.-F.; Liu, Y.-H.; Peng, S.-M.; Chen, J.-T.; Liu, S.-T. Synthesis, characterization and catalytic activity of saturated and unsaturated N-heterocyclic carbene iridium(I) complexes. Dalton Trans. 2009, 861–867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Tolman, C.R. Steric effects of phosphorus ligands in organometallic chemistry and homogeneous catalysis. Chem. Rev. 1977, 77, 313–348. [Google Scholar] [CrossRef]
  28. Mantilli, L.; Gérard, D.; Besnard, C.; Mazet, C. Structure-Activity Relationship in the Iridium-Catalyzed Isomerization of Primary Allylic Alcohols. Eur. J. Inorg. Chem. 2012, 2012, 3320–3330. [Google Scholar] [CrossRef]
  29. Díez-González, S.; Nolan, S.P. Stereoelectronic parameters associated with N-heterocyclic carbene (NHC) liagnds: A quest for understanding. Coord. Chem. Rev. 2007, 251, 874–883. [Google Scholar]
  30. Nelson, D.J.; Truscott, B.J.; Slawin, A.M.Z.; Nolan, S.P. Synthesis and Reactivity of New Bis(N-heterocyclic carbene) Iridium(I) Complexes. Inorg. Chem. 2013, 52, 12674–12681. [Google Scholar] [CrossRef]
  31. Torres, O.; Martín, M.; Sola, E. Labile N-Heterocyclic Carbene Complexes of Irdium. Organometallics 2009, 28, 863–870. [Google Scholar] [CrossRef]
  32. Smidt, S.P.; Zimmermann, N.; Studer, M.; Pfaltz, A. Enantioselective Hydrogenation of Alkenes with Iridium–PHOX Catalysts: A Kinetic Study of Anion Effects. Chem. Eur. J. 2004, 10, 4685–4693. [Google Scholar] [CrossRef]
Scheme 1. Directed hydrogen isotope exchange.
Scheme 1. Directed hydrogen isotope exchange.
Catalysts 10 00161 sch001
Figure 1. Iridium(I) N-heterocyclic carbene (NHC)/phosphine complexes for directed hydrogen isotope exchange.
Figure 1. Iridium(I) N-heterocyclic carbene (NHC)/phosphine complexes for directed hydrogen isotope exchange.
Catalysts 10 00161 g001
Figure 2. Naturally-occurring imidazolium salt Lepidiline A 3·HCl.
Figure 2. Naturally-occurring imidazolium salt Lepidiline A 3·HCl.
Catalysts 10 00161 g002
Figure 3. Iridium(I) complexes with varying NHC ligands to probe their steric and electronic influence.
Figure 3. Iridium(I) complexes with varying NHC ligands to probe their steric and electronic influence.
Catalysts 10 00161 g003
Scheme 2. Synthesis of Lepidiline A and counter-ion analogues.
Scheme 2. Synthesis of Lepidiline A and counter-ion analogues.
Catalysts 10 00161 sch002
Figure 4. Structures of Lepidiline A (left) and its PF6 derivative (right) as determined by single crystal diffraction. Atoms are shown as 50% probability ellipsoids. H atoms, disorder in the solvent, and anion of 3·HCl and the second independent cation-anion pair of 3·HPF6 have been omitted for clarity.
Figure 4. Structures of Lepidiline A (left) and its PF6 derivative (right) as determined by single crystal diffraction. Atoms are shown as 50% probability ellipsoids. H atoms, disorder in the solvent, and anion of 3·HCl and the second independent cation-anion pair of 3·HPF6 have been omitted for clarity.
Catalysts 10 00161 g004
Scheme 3. Preparation of iridium complex 12.
Scheme 3. Preparation of iridium complex 12.
Catalysts 10 00161 sch003
Figure 5. Molecular structure of 12 as determined by single crystal diffraction. Atoms are shown as 50% probability ellipsoids. H atoms and disorder in the orientation of one phenyl ring have been omitted for clarity.
Figure 5. Molecular structure of 12 as determined by single crystal diffraction. Atoms are shown as 50% probability ellipsoids. H atoms and disorder in the orientation of one phenyl ring have been omitted for clarity.
Catalysts 10 00161 g005
Scheme 4. One-pot synthesis of BArF counter-ion catalyst 6e.
Scheme 4. One-pot synthesis of BArF counter-ion catalyst 6e.
Catalysts 10 00161 sch004
Scheme 5. Preparation of iridium carbonyl complexes 14–16.
Scheme 5. Preparation of iridium carbonyl complexes 14–16.
Catalysts 10 00161 sch005
Scheme 6. Computational evaluation of the C-H activation process with catalyst 6a vs. 7.
Scheme 6. Computational evaluation of the C-H activation process with catalyst 6a vs. 7.
Catalysts 10 00161 sch006
Table 1. Preparation of [(cod)Ir(IBnMe)PPh3]X complexes 6a6d.
Table 1. Preparation of [(cod)Ir(IBnMe)PPh3]X complexes 6a6d.
Catalysts 10 00161 i001
EntryComplexXYield
16aPF675
26bBF487
36cSbF681
46dOTf88
Table 2. Properties of Ir-carbonyl complexes 1416.
Table 2. Properties of Ir-carbonyl complexes 1416.
EntryComplexN-SubstituentRYield%Vbur (PPh3) 2%Vbur (NHC) 2Σ (%Vbur)θ (°) 1
114BnH49%26.928.955.893.2
215BnMe50%27.029.256.289.4 3
92.6 3
316MesH68%27.932.660.5160.0
1 θ represents the angle OC-Ir-CO. 2 Representative %Vbur values are calculated from DFT-optimized [(NHC)(PPh3(Ir(H)2(DCM)2]+ complex cations. Full details are presented in the SI. 3 Carbonyl complex 15 crystallized with two different conformations.
Table 3. Electronic properties of Ir Cl/CO and dihydride complexes.
Table 3. Electronic properties of Ir Cl/CO and dihydride complexes.
Catalysts 10 00161 i002
EntryRR’Cl/CO Complexν(CO)DCMDihydride ComplexδH
1BnH171950 cm−120−21.35 ppm
2BnMe181948 cm−121−21.40 ppm
3MesH191942 cm−122−21.58 ppm
Table 4. Evaluation of complexes 6a and 7 in a benchmark HIE reaction.
Table 4. Evaluation of complexes 6a and 7 in a benchmark HIE reaction.
Catalysts 10 00161 i003
EntryComplexNHCkobs
16aIBnMe0.0224 s−1
27IBn0.0145 s−1
Table 5. Evaluation of counter-ion effects in IBnMe/PPh3 complexes 6.
Table 5. Evaluation of counter-ion effects in IBnMe/PPh3 complexes 6.
Catalysts 10 00161 i004
EntrySubstrate%D
6a, X = PF66b, X = BF46c, X = SbF66d, X = OTf6e, X = BArF
1239696967997
2279187932997
3289887934399
4298289908587
530545416

Share and Cite

MDPI and ACS Style

Cochrane, A.R.; Kennedy, A.R.; Kerr, W.J.; Lindsay, D.M.; Reid, M.; Tuttle, T. The Natural Product Lepidiline A as an N-Heterocyclic Carbene Ligand Precursor in Complexes of the Type [Ir(cod)(NHC)PPh3)]X: Synthesis, Characterisation, and Application in Hydrogen Isotope Exchange Catalysis. Catalysts 2020, 10, 161. https://doi.org/10.3390/catal10020161

AMA Style

Cochrane AR, Kennedy AR, Kerr WJ, Lindsay DM, Reid M, Tuttle T. The Natural Product Lepidiline A as an N-Heterocyclic Carbene Ligand Precursor in Complexes of the Type [Ir(cod)(NHC)PPh3)]X: Synthesis, Characterisation, and Application in Hydrogen Isotope Exchange Catalysis. Catalysts. 2020; 10(2):161. https://doi.org/10.3390/catal10020161

Chicago/Turabian Style

Cochrane, Alison R., Alan R. Kennedy, William J. Kerr, David M. Lindsay, Marc Reid, and Tell Tuttle. 2020. "The Natural Product Lepidiline A as an N-Heterocyclic Carbene Ligand Precursor in Complexes of the Type [Ir(cod)(NHC)PPh3)]X: Synthesis, Characterisation, and Application in Hydrogen Isotope Exchange Catalysis" Catalysts 10, no. 2: 161. https://doi.org/10.3390/catal10020161

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