Next Article in Journal
Dehydrogenation of Formic Acid over a Homogeneous Ru-TPPTS Catalyst: Unwanted CO Production and Its Successful Removal by PROX
Next Article in Special Issue
Synthesis of New C2-Symmetric Six-Membered NHCs and Their Application for the Asymmetric Diethylzinc Addition of Arylaldehydes
Previous Article in Journal
Efficient Degradation of Refractory Organics Using Sulfate Radicals Generated Directly from WO3 Photoelectrode and the Catalytic Reaction of Sulfate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

On the Structure of Chiral Dirhodium(II) Carboxylate Catalysts: Stereoselectivity Relevance and Insights

Biomedical Science Department, Faculty of ESTeM, University of Canberra, Canberra, ACT 2601, Australia
Catalysts 2017, 7(11), 347; https://doi.org/10.3390/catal7110347
Submission received: 2 November 2017 / Revised: 14 November 2017 / Accepted: 17 November 2017 / Published: 20 November 2017
(This article belongs to the Special Issue Asymmetric Catalysis in Organic Synthesis)

Abstract

:
Modern experiments have offered alternative interpretations on the symmetry of chiral dirhodium(II) carboxylate complexes and its relationship to their level of enantioselectivity. So, this contribution is to provide an insight on how the knowledge around the structure of these catalysts has evolved with a particular emphasis on the impact of this knowledge on enantioselectivity prediction and catalyst design.

Graphical Abstract

1. Introduction

Chiral dirhodium(II) paddlewheel complexes are among the most attractive catalysts that found a widespread application in the fields of metal-nitrene and metal-carbene transformations [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15]. This immense interest originates from their exceptional ability to effectively catalyze a broad spectrum of reactions with high levels of chemo-, regio- and stereo-selectivity. These transformations involve aziridinations [16,17,18], C-H insertions [7,19,20], ylide transformations [21,22,23,24,25], Lewis acid-promoted reactions [21,26,27,28,29,30], cross-coupling reactions [31], cyclopropanation and cyclopropenation reactions [32,33,34,35]. In fact, they have proven their potential, particularly, in the field of asymmetric synthesis. As an example to showcase the utility of these complexes, methyl styryldiazoacetate (1), in the presence of a suitable substrate and a dirhodium(II) catalyst, can undergo a range of chemical transformations including cyclopropanation [36,37,38], C-H insertion [39,40], Si-H insertion [41], tandem cyclopropanation/Cope rearrangement [42,43], tandem O-H insertion/[2,3] sigmatropic rearrangement [44], vinylogous reactivity [45,46], tandem C-H insertion/Cope rearrangement [47,48], or [3+2] cycloaddition [49] generating a variety of useful synthetic entities in excellent diastereo- and enantioselectivity (Scheme 1).
Chiral dirhodium(II) carboxylate complexes constitute a very important class of dirhodium(II) catalysts. Their structure consists of two rhodium atoms connected together through a unique dirhodium bridge. This dirhodium axes is connected to four chiral electron withdrawing carboxylate ligands and they have two vacant axial positions available for carbene binding [50,51]. Although several highly enantioselective dirhodium(II) carboxylate catalysts have been reported to date [2,52] and their application spectrum has expanded dramatically, limited evidence is known on how chirality is tailored at the active center of these complexes by its four chiral carboxylate ligands. As a consequence, this article is providing an insight on how the knowledge around the structure of dirhodium(II) carboxylate complexes has evolved. This is with a particular emphasis on the impact of this knowledge on enantioselectivity prediction and catalyst design. To the best of our knowledge, this topic was last reviewed by Hansen and Davies in 2008 [53] and since then subsequent investigations have offered alternative interpretations on the symmetry of chiral dirhodium(II) carboxylate complexes and its relationship to their level of enantioselectivity.

2. Ligand Blocking Groups Arrangements

Generally for chiral dirhodium(II) carboxylate complexes, it is agreed that the ligand blocking groups can adapt either an up (α) or down (β) positions around the catalyst core. It was also assumed that, the ligand’s blocking group cannot be situated in the periphery of the dirhodium(II) core as it would bump into the neighboring ligand [36,53]. Thus, by taking into account the α- and β-arrangements for all four ligands, four possible conformations can be mainly considered: α,α,α,α (C4-symmetry), α,α,β,β (C2-symmetry), α,β,α,β (D2-symmetry) and α,α,α,β (C1-symmetry) (Figure 1) [2,4,36,53,54].

2.1. α,β,α,β-Conformation

Initially, Hansen and Davies suggested that catalysts having two different carbene-formation sites should not lead to high enantiocontrol [53]. In other words, complexes with α,α,β,β- and α,β,α,β-arrangements (having C2- and D2-symmetry, respectively) are the only effective catalysts in terms of enantioselectivity as they possess two equivalent catalyst faces. While complexes with α,α,α,α- and α,α,α,β-conformations, which have non-equivalent catalyst faces, are likely to induce low or no enantioinduction as their more accessible and kinetically active face is achiral. This previous proposal from Davies et al. originates from their discovery of Rh2(S-DOSP)4 catalyst (Scheme 2) [36], which offered extraordinary enantioselectivity in a wide range of chemical transformations (up to 99% ee) [7,13,36,40,42,44,47,48,55,56,57,58,59,60,61,62,63,64,65,66,67,68]. This exceptional enantioselectivity of Rh2(S-DOSP)4 was proposed to originate from a favored α,β,α,β-arrangement of ligands adapted during catalysis. The arrangement that leads to a catalyst with two equivalent rhodium active sites and adequate sterically overburden groups that can limit the trajectories approaching the axial carbene ligand (Scheme 2) [7,13,36,53]. For example, low temperature Rh2(S-DOSP)4-catalyzed cyclopropanation of styrene with methyl styryldiazoacetate resulted in the generation of the corresponding cyclopropane product in more than 50:1 E/Z diastereomeric ratio and 98% ee (Scheme 2a) [13,69]. A general model was then proposed to explain the observed stereochemical outcome [13,36,68,70]. In this model, Rh2(S-DOSP)4 is assumed to favor the α,β,α,β-conformation during catalysis at which, the si-face of the metal-carbene complex is protected by a ligand blocking group. The substrate will approach from the re-face to generate the product in the observed absolute configuration (Scheme 2a).
To support the D2-symmetry hypothesis of Rh2(S-DOSP)4, the authors introduced the second generation prolinate-based complex, Rh2(S-biTISP)2 (Scheme 2) [13,71,72]. As the ligands themselves possess a C2-symmetry, Rh2(S-biTISP)2 was anticipated to adapt the higher D2-symmetrical structure. More importantly, the complex was expected to be more rigid at which, the ligands are conformationally locked in the α,β,α,β-arrangement around the catalyst active centers. For this catalyst, not only both faces are equivalent but also all staggered binding orientations of the axial carbene ligands are identical with respect to the approaching substrate [13,53,71]. As expected by the authors, Rh2(S-biTISP)2 demonstrated excellent enantioselectivity in the enantioselective intermolecular cyclopropanation of styrene with methyl styryldiazoacetate leading to the generation of the cyclopropane product in 98% ee (Scheme 2b) [13,71]. This result very well supported the proposed concept that the observed enantioselectivity of Rh2(S-DOSP)4 originates from an α,β,α,β-conformation adapted during catalysis. However, the use of Rh2(S-biTISP)2 catalyst preferentially resulted in the generation of the (1S,2S)-cyclopropane product (Scheme 2b) which was the opposite to that obtained with the Rh2(S-DOSP)4-catalysed reaction (Scheme 2a). The authors returned this difference in asymmetric induction to the different preferred staggered orientation for the carbene binding which causes different channels of the metal-carbene intermediate to be exposed to the approaching alkene [13].
In the same context, Rh2(R-BTPCP)4 was also reported as a catalyst that exhibited high enantioselectivity levels [73,74,75]. For example, the Rh2(R-BTPCP)4-catalyzed cyclopropanation of styrene with methyl styryldiazoacetate generated the cyclopropane product in 92% ee [73]. Based on its X-ray crystal structure, this enantioselectivity was again returned to the α,β,α,β-distribution of ligands around the dirhodium(II) core (Scheme 3) [73].
This research was further extended and, very recently, Davies et al. reported its analogue, Rh2(R-3,5-(p-tBuPh)TPCP)4, as a very successful catalyst in achieving highly regio-, diastereo- and enantio-selective C-H functionalization at the inactivated C2 position of n-alkanes and terminally substituted n-alkyl compounds (Scheme 4) [52]. Again, X-ray crystallography along with computational studies showed that the catalyst disfavors the existence of two adjacent 3,5-di(tert-butylphenyl)phenyl groups on the same face of the catalyst. Hence the preferred orientation of ligands is α,β,α,β around both catalyst’s active sites (Scheme 4) [52].

2.2. α,α,β,β- and α,α,α,α-Conformations

2.2.1. Global Catalyst Symmetry

The previous proposal that only catalysts with D2- and C2-symmetric structures are effective in terms of enantiocontrol [53] relied also on Hashimoto’s reports for Rh2(S-PTTL)4. In 1993, Hashimoto, Ikegami and co-workers first reported their famous complex, Rh2(S-PTTL)4 [76,77], as a catalyst that manifested (along with its variants) high enantioselectivity levels in wide range of organic transformations [2,54,78,79,80,81,82,83,84]. For example, in enantioselective [2,3]-sigmatropic rearrangement of the cyclic propargylic oxonium ylide shown in Scheme 5a, Rh2(S-PTTL)4 provided the best enantioselectivity of 79% ee of the corresponding allenic bearing benzofuran-3-one product (Scheme 5a) [85].
Hashimoto et al. initially proposed that Rh2(S-PTTL)4 favors the α,α,β,β-conformation during catalysis. This assumption was based on the X-ray crystal structure of its analogue, Rh2(S-PTPA)4, which also showed moderate to good enantioselectivity in some asymmetric transformations [76,79,85,86]. For example, in aromatic C-H insertion of α-diazoketones, Rh2(S-PTPA)4 resulted into the formation of the corresponding 1-alkyl-1-phenyl-2-indanones in up to 95% ee (Scheme 5b) [86]. The X-ray crystal structure of Rh2(S-PTPA)4 showed that the catalyst is adapting the α,α,β,β-conformation in solid state (Scheme 5b) [76,85,86,87].
Based on Hashimoto’s hypothesis, a general model was proposed to account for the observed enantioselectivity with Rh2(S-PTTL)4 [85,86,88,89]. In this model, it was proposed that Rh2(S-PTTL)4 preferentially adapts the α,α,β,β-conformation during catalysis. The carbene orients itself at which, its bulky side is directed away from the catalyst walls (Scheme 5a). The substrate approaches from the re-face to generate the product in the observed absolute configuration.
Later, Davies et al. noticed that, in some reactions, the steric bulk of the ligand affects the enantioselectivity of the catalyst at which the enantioselectivity increased by increasing the steric bulk at the α-carbon of the ligands. For example, in dirhodium(II)-catalyzed enantioselective tandem carbonyl ylide formation/intermolecular 1,3-dipolar cycloaddition of α-diazoketones, the best enantioselectivity was observed when Rh2(S-PTTL)4 was used as a catalyst (74% ee), followed by Rh2(S-PTV)4 (51% ee) and Rh2(S-PTA)4 (35% ee) [90] (Scheme 6a). As a consequence, Davies et al. suggested that the logical way to further develop the Rh2(S-PTTL)4 catalyst is to substitute the tert-butyl group connected to the α-carbon with a much bulkier hydrocarbon group. Accordingly, they introduced Rh2(S-PTAD)4 catalyst which is derived from l-adamantylglycine ligands (Scheme 6b) [37]. As expected, this catalyst succeeded to afford enantioselectivity levels superior to Rh2(S-PTTL)4 in different asymmetric transformations [37,88,91,92,93,94,95]. For example, Rh2(S-PTAD)4-catalyzed cyclopropanation of styrene with α-diazobenzylphosphonate resulted into the generation of the corresponding cyclopropylphosphonate product in high levels of enantioselectivity (99% ee) compared to Rh2(S-DOSP)4 (34% ee), Rh2(S-biTISP)2 (88% ee) and Rh2(S-PTTL)4 (97% ee) (Scheme 6b) [37].
Up to this point, Hansen and Davies theoretical assumption that only catalysts with α,α,β,β- and α,β,α,β-arrangements results in high enantioselectivity and not the others [53] was very well supported by scientific evidence. Nevertheless in 2009, Fox et al. were able to capture an X-ray crystal structure for Rh2(S-PTTL)4 [78,87]. Surprisingly, the obtained structure revealed that the catalyst is adapting the α,α,α,α-conformation in solid state and not the expected α,α,β,β-conformation (Figure 2). Furthermore, the structure showed that the catalyst is having a rectangular chiral cavity (C2- rather than C4-symmetry) with only one ethyl acetate axial ligand coordinated to the rhodium atom from the cavity face.
Based on the new information, Fox et al. introduced a new model to account for the enantioselectivity exhibited by Rh2(S-PTTL)4 in enantioselective cyclopropanation of styrene with α-alkyldiazoesters (Scheme 7) [78,87]. In this model, Rh2(S-PTTL)4 was proposed to keep its α,α,α,α-conformation during catalysis and the carbene is anticipated to align itself within the wide dimension of the catalyst’s rectangular cavity leaving its si-face accessible by the substrate. Also according to Fox model, the tert-butyl groups are crucial to restrict the reactivity to the chiral cavity face of the catalyst and limit the racemic background reaction.
Fox model was very well supported by the results obtained earlier for Rh2(S-PTAD)4 by being more enantioselective than Rh2(S-PTTL)4 [37,91,92,93,95,96]. The results were very relevant as the much bulkier adamantyl groups would more efficiently block the carbene ligands from coordinating to the achiral rhodium face (shrouded by the adamantly groups) compared to the tert-butyl groups of Rh2(S-PTTL)4 leading to the observed more enantioselective outcome. Moreover, Fürstner et al. recently succeeded in capturing an X-ray crystal structure of the Rh2(S-PTTL)4=C(MeOC6H4)CO2Me intermediate (Figure 3) [97]. The obtained structure showed that the entire carbene resides within a long narrow pocket formed by the four N-phthalimido groups of Rh2(S-PTTL)4, that adapt an α,α,α,α-arrangement (C2-symmetric). The overall situation looked, therefore, in good agreement with the model proposed by Fox. But, low temperature NMR data for this intermediate returned a single set of signals for the four chiral carboxylate ligands [97]. This observation was not in a good accord with a C2-symmetric binding site as proposed by the Fox model.
Fox proposal was also challenged by data reported from different research groups. Adly, Ghanem and co-workers reported the X-ray crystal structure of Rh2(S-PTAD)4 [98]. Similar to Rh2(S-PTTL)4, the obtained structure showed that the catalyst is adapting the α,α,α,α-conformation in solid state, however, the catalyst was observed to form a bis(EtOAc) adduct when crystalized from ethyl acetate/n-hexane (Figure 4). This observation contradicted with the assumption proposed by Fox et al. in their model and confirmed that the adamantyl groups are not completely blocking the “achiral” catalyst face. In other words, there is still enough room for a Lewis basic ligand to coordinate to the achiral axial rhodium site. This observation provided direct evidence that both Rh atoms are still accessible by the diazo substrates even after the introduction of the bulkier adamantyl groups [98]. In addition, results have shown that complexes with tert-butyl groups at their α-carbon produced enantioselectivity levels similar and sometimes higher than Rh2(S-PTAD)4 which carries the bulkier adamantyl groups [98].
Also, different reported observations confirmed the flexibility of the PTTL ligands around the rhodium active center, especially in solution and emphasized the effect of the shape and size of the axial ligand on the overall structure of the dirhodium(II)-carbene intermediate. For example, 2D heteronuclear NOESY data reported by the groups of Charette [54] and Duddeck [99], independently, confirmed that Rh2(S-PTTL)4 and analogues have a mobile conformation in solution. This confirms the existence of other conformers with at least one N-phthalimido group is flipped downward. Further and based on X-ray crystallographic data, Rh2(S-PTTL)4(NC-C6H4-CN) adduct was found to adapt a conformation in which one of the ligands is twisted by only about 80° in solid state [87]. In other words, the ligand’s blocking group lies in the periphery of the catalyst, the case that was originally supposed to be problematic due to steric constraints [36,53]. Gardiner and Ghanem [2,100] also succeeded to grow crystals of Rh2(S-PTPA)4(EtOAc)2 and Rh2(S-PTPA)4(MeOH)(H2O) and they found that different conformations exist for each of these adducts at which, all of them feature inequivalent axial coordination sites unlike its original α,α,β,β-structure previously reported by Hashimoto [76,85,86,87]. The Rh2(S-PTPA)4(EtOAc)2 adduct exhibited an α,α,α,β-conformation, while Rh2(S-PTPA)4(MeOH)(H2O) adduct exhibited an α,α,α,α-conformation in solid state.

2.2.2. Local Ligand Symmetry

Following the classical way of catalyst design, all reported chiral dirhodium(II) carboxylate complexes derived from l-amino acid ligands were designed to have a symmetric N-protecting groups for the construction of the chiral ligands [4,53]. However in 2004, Müller and Ghanem reported several Rh2(S-NTTL)4 analogues at which, one hydrogen on the N-protecting group is substituted to generate catalysts carrying asymmetric N-protecting groups for asymmetric cyclopropanation of styrene with dimethyl malonate [101] and Meldrum’s acid [102] (Scheme 8). The authors were surprised when screening results revealed that Rh2(S-4-Br-NTTL)4-catalyzed cyclopropanation proceeded with far improved levels of enantioselectivity compared to Rh2(S-NTTL)4.
Unfortunately, the authors had no success in growing X-ray quality crystals for Rh2(S-4-Br-NTTL)4. Instead, the X-ray crystal structure of Rh2(S-NTTL)4 (Figure 5) was used as a model to account for the enhanced enantioselectivity that accompanied Rh2(S-4-Br-NTTL)4 [102]. The authors proposed that, if Rh2(S-4-Br-NTTL)4 is retaining a similar structure to Rh2(S-NTTL)4 during catalysis (following Fox et al. proposal), the bromo-substituents would be situated at the cavity rim and are likely to exert a strong influence on the enantiofacial discrimination of the approaching substrate leading to the observed enhanced enantioselectivity. The authors also returned the improved performance of the 4-Br substituted catalyst over the 4-Cl analogue to the size of the substituents. The larger bromo- substituents are expected to exert more influence at the cavity rim leading to higher enantioselectivity in the case of Rh2(S-4-Br-NTTL)4.
Later, Adly and Ghanem investigated this further and they introduced a new approach for the design of dirhodium(II) carboxylate complexes. This approach is through the reduction of the N-protecting group symmetry by the inclusion of a bulky substituent at the catalyst’s cavity rim. By the application of their new approach, the authors were able to introduce their catalyst, Rh2(S-tertPTTL)4 (Scheme 9) [89,98,103]. For this catalyst, the planarity of the N-phthalimido groups is maintained but the local symmetry of the N-protecting group was reduced from C2v to Cs by virtue of the substituents. Screening of Rh2(S-tertPTTL)4 in different enantioselective cyclopropanation reactions demonstrated that, generally, Rh2(S-tertPTTL)4 is more enantioselective than Rh2(S-PTTL)4, while returning comparable enantioselectivity to Rh2(S-PTAD)4 (Scheme 9) [98].
The X-ray crystal structure of Rh2(S-tertPTTL)4 showed that the catalyst is adapting the α,α,α,α-conformation in solid state (Scheme 9) featuring the tert-butyl substituents similarly disposed towards the “corner” of the nearly square-shaped cavity. The cavity is approximately square if compared to the reported structure of Rh2(S-PTTL)4. The authors claimed that the added tert-butyl substituent on the N-phthalimido group in Rh2(S-tertPTTL)4 reinforced the overall chiral twist of the cavity, while, at the same time, not placing added steric hindrance to the binding of axial carbene ligand [98]. The authors proposed that, if this geometry is relevant during catalysis, it is likely to be the reason for the observed enhanced enantioselectivity relative to its parent, Rh2(S-PTTL)4.

2.3. α,α,α,β-Conformation

In 2012, Charette et al. explored the replacement of one of the ligands of Rh2(S-TCPTTL)4 (the chlorinated version of Rh2(S-PTTL)4) with N-phthalimido-2-aminoisobutyric acid ligand (PTAiB). This led to the generation of the lower symmetry heteroleptic complex, Rh2(S-TCPTTL)3(PTAiB) [104]. The X-ray crystal structure of this complex revealed that it is adapting the α,α,α,β-conformation in solid state (Scheme 10). Surprisingly, screening results demonstrated that lowering the global symmetry of the catalyst had a beneficial impact on its enantioselectivity. In the cyclopropanation reaction of alkenes with α-nitro diazoacetophenones, Rh2(S-TCPTTL)3(PTAiB) resulted into the formation of the cyclopropane product in 96.4% ee which was higher compared to its parent Rh2(S-TCPTTL)4 (92.4% ee) (Scheme 10).
Also in the same year, Fox et al. reported the mixed ligated complex, Rh2(S-PTTL)3(TPA) (Scheme 11) [105]. The X-ray crystal structure of this complex showed that all of the N-phthalimido groups are on one face of the complex in a structural similarity to Rh2(S-PTTL)4 but with an incomplete chiral pocket (analogous to an α,α,α,β-arrangement). Rh2(S-PTTL)3(TPA) returned its best enantioselectivity in the cyclopropanation of styrene with ethyl α-diazobutanoate (88% ee) which was higher than the observed enantioselectivity with Rh2(S-PTTL)4 in the same reaction (79% ee) (Scheme 11). The scope of this catalyst was expanded to include other substrate classes, namely aliphatic alkynes, silylacetylenes and α-olefins, that were particularly challenging in intermolecular cyclopropanations with α-alkyl-α-diazoesters. Generally, Rh2(S-PTTL)3(TPA) was able to catalyze enantioselective cyclopropanations returning yields and enantioselectivities comparable and sometimes higher than Rh2(S-PTTL)4 [105].

3. Conclusions

From the above demonstration, all models proposed to account for the enantioselectivity observed with different dirhodium(II) carboxylate catalysts relied mainly on the obtained X-ray crystal structures of these catalysts (except for Rh2(S-DOSP)4 and related prolinate-based complexes). However, due to crystal packing, π-stacking and/or existence of axial and lattice solvent molecules within the crystal unit cell, the static pose of a catalyst and its chiral ligands portrayed in its X-ray crystal structure does not validate the adaption of the catalyst to the same conformation in solution and during catalysis. In other words, based on the size and the shape of the axial carbene ligand, as well as the reaction solvent used, the metal-carbene intermediate may adapt a different conformation during catalysis to the one observed in its X-ray crystal structure.
In fact, carboxylate ligands contain conformationally mobile units [13,54,87,98,99] which might give the ability to the dirhodium(II) catalyst to adapt its conformation to the particular substrate undergoing the reaction. Then enantioselection mechanisms can mainly originate from the different preferred reaction channels as a function of the different steric interactions taking place between the approaching substrate and the formed dirhodium(II)-carbene complex. This can explain the compatibility of a particular dirhodium(II) carboxylate catalyst with a particular substrate returning high levels of enantioselectivity, while the same catalyst might not retain the same level of enantioselectivity after switching to a different substrate. As a result, catalyst optimization is always necessary in method development to find the optimum catalyst that can return the highest possible level of enantioselectivity with the particular substrate undergoing the reaction. Based on that, researchers must not only limit their focus on the four main conformations of dirhodium(II) carboxylate complexes and an expansion in the considered conformational space is essential.
Even by relying on the catalyst X-ray crystal structures alone to justify the observed enantioselectivity, it is obvious from the demonstration above that dirhodium(II) carboxylate catalysts with α,α,α,α-, α,α,α,β-, α,α,β,β- and α,β,α,β-ligand arrangements were all reported to lead to high levels of enantioselectivity. As a consequence, the original proposal that catalysts with α,α,β,β- and α,β,α,β-geometries will only result in high enantioselectivity [53] can no longer be considered as valid. In other words, symmetry of the catalyst is not a mandatory prerequisite for achieving high levels of enantioselectivity. Subsequently, the use of high symmetry ligands as the traditional way to produce high symmetry catalysts [53] may not be as essential as previously thought.
All in all, although the knowledge around the structure of dirhodium(II) carboxylate complexes has been expanded dramatically in the last decade, the topic is still in its infancy. With the fast development in X-ray crystallography, computational chemistry and NMR spectroscopy, researchers may prosper in solving the mystery behind this topic. This, in turn, will aid to a great extent in the design of highly stereoselective chiral dirhodium(II) catalysts in the future.

Acknowledgments

The author would like to acknowledge the Chirality research group members at the University of Canberra, Australia, for their support.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Davies, H.M.L.; Manning, J.R. Catalytic C-H functionalization by metal carbenoid and nitrenoid insertion. Nature 2008, 451, 417–424. [Google Scholar] [CrossRef] [PubMed]
  2. Adly, F.G.; Ghanem, A. Enantiomerically Pure Compounds by Enantioselective Synthetic Chiral Metal Complexes. In Asymmetric Synthesis of Drugs and Natural Products; Nag, A., Ed.; CRC Press: Raton, FL, USA, 2018; in press. [Google Scholar]
  3. Deng, Y.; Qiu, H.; Srinivas, H.D.; Doyle, M.P. Chiral Dirhodium(II) Catalysts for Selective Metal Carbene Reactions. Curr. Org. Chem. 2016, 20, 61–81. [Google Scholar] [CrossRef]
  4. Adly, F.G.; Ghanem, A. Chiral Dirhodium (II) Carboxylates and Carboxamidates as Effective Chemzymes in Asymmetric Synthesis of Three-Membered Carbocycles. Chirality 2014, 26, 692–711. [Google Scholar] [CrossRef] [PubMed]
  5. El-Deftar, M.; Adly, F.G.; Gardiner, M.G.; Ghanem, A. Chiral Dirhodium Catalysts: A New Era for Asymmetric Catalysis. Curr. Org. Chem. 2012, 16, 1808–1836. [Google Scholar] [CrossRef]
  6. Trindade, A.F.; Coelho, J.A.S.; Afonso, C.A.M.; Veiros, L.F.; Gois, P.M.P. Fine Tuning of Dirhodium(II): Expolring the Axial Modification. ACS Catal. 2012, 2, 370–383. [Google Scholar] [CrossRef]
  7. Davies, H.M.L.; Morton, D. Guiding principles for site selective and stereoselective intermolecular C-H functionalization by donor/acceptor rhodium carbenes. Chem. Soc. Rev. 2011, 40, 1857–1869. [Google Scholar] [CrossRef] [PubMed]
  8. Davies, H.M.L.; Bois, J.D.; Yu, J.-Q. C-H Functionalization in organic synthesis. Chem. Soc. Rev. 2011, 40, 1855–1856. [Google Scholar] [CrossRef] [PubMed]
  9. Doyle, M.P.; Duffy, R.; Ratnikov, M.; Zhou, L. Catalytic Carbene Insertion into C-H Bonds. Chem. Rev. 2010, 110, 704–724. [Google Scholar] [CrossRef] [PubMed]
  10. Davies, H.M.L.; Hedley, S.J. Intermolecular reactions of electron-rich heterocycles with copper and rhodium carbenoids. Chem. Soc. Rev. 2007, 36, 1109–1119. [Google Scholar] [CrossRef] [PubMed]
  11. Merlic, C.A.; Zechman, A.L. Selectivity in Rhodium(II) Catalyzed Reactions of Diazo Compounds: Effects of Catalyst ElectoPhilicity, Diazo Substitution and Substrate Substitution. From Chemoselectivity to Enantioselectivity. Synthesis 2003, 34, 1137–1156. [Google Scholar] [CrossRef]
  12. Colacot, T.J. An overview on the application of “Doyle catalysts” in asymmetric cyclopropanation, cyclopropenation and C-H insertion reactions. Proc. Ind. Acad. Sci. (J. Chem. Sci.) 2000, 11, 197–207. [Google Scholar] [CrossRef]
  13. Davies, H.M.L. Dirhodium tetra(N-arylsulfonylprolinates) as chiral catalysts for asymmetric transformations of vinyl and aryldiazoacetates. Eur. J. Org. Chem. 1999, 30, 2459–2469. [Google Scholar] [CrossRef]
  14. Davies, H.M.L. Rhodium-stabilized vinylcarbenoid intermediates in organic synthesis. Curr. Org. Chem. 1998, 2, 463–488. [Google Scholar]
  15. Collet, F.; Lescot, C.; Liang, C.; Dauban, P. Studies in catalytic C-H amination involving nitrene C-H insertion. Dalton Trans. 2010, 39, 10401–10413. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, X.-J.; Yan, M.; Huang, D. Catalyzed addition of diazoacetoacetates to imines: Synthesis of highly functionalized aziridines. Org. Biomol. Chem. 2009, 7, 187–192. [Google Scholar] [CrossRef] [PubMed]
  17. Yamawaki, M.; Tanaka, M.; Abe, T.; Anada, M.; Hashimoto, S. Catalytic enantioselective aziridination of alkenes using chiral dirhodium(II) carboxylates. Heterocycles 2007, 72, 709–721. [Google Scholar] [CrossRef]
  18. Catino, A.J.; Nichols, J.M.; Forslund, R.E.; Doyle, M.P. Efficient Aziridination of Olefins Catalyzed by Mixed-Valent Dirhodium(II,III) Caprolactamate. Org. Lett. 2005, 7, 2787–2790. [Google Scholar] [CrossRef] [PubMed]
  19. Davies, H.M.L.; Denton, J.R. Application of donor/acceptor-carbenoids to the synthesis of natural products. Chem. Soc. Rev. 2009, 38, 3061–3071. [Google Scholar] [CrossRef] [PubMed]
  20. Kubiak, R.W.; Mighion, J.D.; Wilkerson-Hill, S.M.; Alford, J.S.; Yoshidomi, T.; Davies, H.M. Enantioselective Intermolecular C-H Functionalization of Allylic and Benzylic sp3 C-H Bonds Using N-Sulfonyl-1, 2, 3-triazoles. Org. Lett. 2016, 18, 3118–3121. [Google Scholar] [CrossRef] [PubMed]
  21. Doyle, M.P. Perspective on Dirhodium Carboxamidates as Catalysts. J. Org. Chem. 2006, 71, 9253–9260. [Google Scholar] [CrossRef] [PubMed]
  22. Doyle, M.P.; McKervey, M.A.; Ye, T. Modern Catalytic Methods for Organic Synthesis with Diazo Compounds: From Cyclopropanes to Ylides; Wiley: New York, NY, USA, 1998. [Google Scholar]
  23. Hodgson, D.M.; Stupple, P.A.; Pierard, F.Y.T.M.; Labande, A.H.; Johnstone, C. Development of dirhodium(II)-catalyzed generation and enantioselective 1,3-dipolar cycloaddition of carbonyl ylides. Chem.-Eur. J. 2001, 7, 4465–4476. [Google Scholar] [CrossRef]
  24. Doyle, M.P.; Forbes, D.C.; Vasbinder, M.M.; Peterson, C.S. Enantiocontrol in the Generation and Diastereoselective Reactions of Catalytically Generated Oxonium and Iodonium Ylides. Metal-Stabilized Ylides as Reaction Intermediates. J. Am. Chem. Soc. 1998, 120, 7653–7654. [Google Scholar] [CrossRef]
  25. Doyle, M.P.; Hu, W. Macrocycle formation from catalytic metal carbene transformations. Synlett 2001, 2001, 1364–1370. [Google Scholar] [CrossRef]
  26. Doyle, M.P.; Phillips, I.M.; Hu, W. A New Class of Chiral Lewis Acid Catalysts for Highly Enantioselective Hetero-Diels-Alder Reactions:  Exceptionally High Turnover Numbers from Dirhodium(II) Carboxamidates. J. Am. Chem. Soc. 2001, 123, 5366–5367. [Google Scholar] [CrossRef] [PubMed]
  27. Doyle, M.P.; Valenzuela, M.; Huang, P. Asymmetric hetero-Diels-Alder reaction catalyzed by dirhodium(II) carboxamidates. Proc. Natl. Acad. Sci. USA 2004, 101, 5391–5395. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, Y.; Wolf, J.; Zavalij, P.; Doyle, M.P. Cationic chiral dirhodium carboxamidates are activated for lewis acid catalysis. Angew. Chem. Int. Ed. 2008, 47, 1439–1442. [Google Scholar] [CrossRef] [PubMed]
  29. Watanabe, N.; Shimada, N.; Anada, M.; Hashimoto, S. Enantio- and diastereoselective hetero-Diels-Alder reactions between 4-methyl-substituted Rawal’s diene and aldhydes catalyzed by chiral dirhodium(II) carboxamidates: Catalytic asymmetric synthesis of (−)-cis-aerrangis lactone. Tetrahedron Asymmetry 2014, 25, 63–73. [Google Scholar] [CrossRef]
  30. Anada, M.; Washio, T.; Shimada, N.; Kitagaki, S.; Nakajima, M.; Shiro, M.; Hashimoto, S. A new dirhodium(II) carboxamidate complex as a chiral Lewis acid catalyst for enantioselective heteo-Diels-Alder reactions. Angew. Chem. Int. Ed. 2004, 43, 2665–2668. [Google Scholar] [CrossRef]
  31. Hansen, J.H.; Parr, B.T.; Pelphrey, P.; Jin, Q.; Autschbach, J.; Davies, H.M.L. Rhodium(II)-Catalyzed Cross-Coupling of Diazo Compounds. Angew. Chem., Int. Ed. 2011, 50, 2544–2548. [Google Scholar] [CrossRef] [PubMed]
  32. Doyle, M.P.; Protopopova, M.N. New aspects of catalytic asymmetric cyclopropanation. Tetrahedron 1998, 54, 7919–7946. [Google Scholar] [CrossRef]
  33. Davies, H.M.L.; Antoulinakis, E.G. Intermolecular Metal-Catalyzed Carbenoid Cyclopropanations. In Organic Reactions; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2001; Volume 57, pp. 1–326. [Google Scholar]
  34. Lebel, H.; Marcoux, J.-F.; Molinaro, C.; Charette, A.B. Stereoselective Cyclopropanation Reactions. Chem. Rev. 2003, 103, 977–1050. [Google Scholar] [CrossRef] [PubMed]
  35. Tsuji, J. Modern Rhodium-Catalyzed Organic Reactions; Wiley-VCH: Weinheim, Germany, 2005. [Google Scholar]
  36. Davies, H.M.L.; Bruzinski, P.R.; Lake, D.H.; Kong, N.; Fall, M.J. Asymmetric Cyclopropanations by Rhodium(II) N-(Arylsulfonyl)prolinate-Catalyzed Decomposition of Vinyldiazomethanes in the presence of Alkenes. Practical Enantioselective Synthesis of the four Stereoisomers of 2-Phenylcyclopropan-1-amino Acid. J. Am. Chem. Soc. 1996, 118, 6897–6907. [Google Scholar] [CrossRef]
  37. Reddy, R.P.; Lee, G.H.; Davies, H.M.L. Dirhodium Tetracarboxylate Derived from Adamantylglycine as Chiral Catalyst for Carbenoid Reactions. Org. Lett. 2006, 8, 3437–3440. [Google Scholar] [CrossRef] [PubMed]
  38. Davies, H.M.L.; Venkataramani, C. Dirhodium Tetraprolinate-Catalyzed Asymmetric Cyclopropanations with High Turnover Numbers. Org. Lett. 2003, 5, 1403–1406. [Google Scholar] [CrossRef] [PubMed]
  39. Davies, H.M.L.; Beckwith, E.J. Catalytic enantioselective C-H activation by means of metal-carbenoid-induced C-H insertion. Chem. Rev. 2003, 103, 2861–2903. [Google Scholar] [CrossRef] [PubMed]
  40. Davies, H.M.; Hansen, T. Asymmetric Intermolecular Carbenoid C−H Insertions Catalyzed by Rhodium (II)(S)-N-(p-Dodecylphenyl) sulfonylprolinate. J. Am. Chem. Soc. 1997, 119, 9075–9076. [Google Scholar] [CrossRef]
  41. Davies, H.M.; Hansen, T.; Rutberg, J.; Bruzinski, P.R. Rhodium(II)(S)-N-(arylsulfonyl) prolinate catalyzed asymmetric insertions of vinyl-and phenylcarbenoids into the Si-H bond. Tetrahedron Lett. 1997, 38, 1741–1744. [Google Scholar] [CrossRef]
  42. Davies, H.M.; Stafford, D.G.; Doan, B.D.; Houser, J.H. Tandem asymmetric cyclopropanation/cope rearrangement. A highly diastereoselective and enantioselective method for the construction of 1,4-cycloheptadienes. J. Am. Chem. Soc. 1998, 120, 3326–3331. [Google Scholar] [CrossRef]
  43. Davies, H.M.L. [3 + 4] annulations between Rhodium-stabilized vinylcarbenoids and dienes. In Advances in Cycloaddition, Volume 5; Harmata, M., Ed.; JAI Press INC: Stamford, CT, USA, 1999. [Google Scholar]
  44. Li, Z.; Davies, H.M. Enantioselective C-C Bond Formation by Rhodium-Catalyzed Tandem Ylide Formation/[2,3]-Sigmatropic Rearrangement between Donor/Acceptor Carbenoids and Allylic Alcohols. J. Am. Chem. Soc. 2009, 132, 396–401. [Google Scholar] [CrossRef] [PubMed]
  45. Yue, Y.; Wang, Y.; Hu, W. Regioselectivity in Lewis acids catalyzed X–H (O, S, N) insertions of methyl styryldiazoacetate with benzyl alcohol, benzyl thiol and aniline. Tetrahedron Lett. 2007, 48, 3975–3977. [Google Scholar] [CrossRef]
  46. Hansen, J.H.; Davies, H.M. Vinylogous reactivity of silver(I) vinylcarbenoids. Chem. Sci. 2011, 2, 457–461. [Google Scholar] [CrossRef]
  47. Hansen, J.H.; Gregg, T.M.; Ovalles, S.R.; Lian, Y.; Autschbach, J.; Davies, H.M. On the mechanism and selectivity of the combined C-H activation/Cope rearrangement. J. Am. Chem. Soc. 2011, 133, 5076–5085. [Google Scholar] [CrossRef] [PubMed]
  48. Davies, H.M.; Jin, Q. Highly Diastereoselective and Enantioselective C-H Functionalization of 1, 2-Dihydronaphthalenes: A Combined C-H Activation/Cope Rearrangement Followed by a Retro-Cope Rearrangement. J. Am. Chem. Soc. 2004, 126, 10862–10863. [Google Scholar] [CrossRef] [PubMed]
  49. Davies, H.M.; Hu, B. Ring expansion of tert-butyl 1-vinylcyclopropane-1-carboxylates to alpha-ethylidenebutyrolactones. J. Org. Chem. 1992, 57, 4309–4312. [Google Scholar] [CrossRef]
  50. Pirrung, M.C.; Liu, H.; Morehead, A.T. Rhodium Chemzymes: Michaelis-Menten Kinetics in dirhodium(II) carboxylate-catalyzed carbenoid reactions. J. Am. Chem. Soc. 2002, 124, 1014–1023. [Google Scholar] [CrossRef] [PubMed]
  51. Nakamura, E.; Yoshikai, N.; Yamanaka, M. Mechanism of C-H Bond Activation/C-C Bond Formation Reaction between Diazo Compound and Alkane Catalyzed by Dirhodium Tetracarboxylate. J. Am. Chem. Soc. 2002, 124, 7181–7192. [Google Scholar] [CrossRef] [PubMed]
  52. Liao, K.; Negretti, S.; Musaev, D.G.; Bacsa, J.; Davies, H.M.L. Site-selective and stereoselective functionalization of inactivated C-H bonds. Nature 2016, 533, 230–234. [Google Scholar] [CrossRef] [PubMed]
  53. Hansen, J.; Davies, H.M.L. High Symmetry Dirhodium(II) Paddlewheel Complexes as Chiral Catalysts. Coord. Chem. Rev. 2008, 252, 545–555. [Google Scholar] [CrossRef] [PubMed]
  54. Lindsay, V.N.G.; Lin, W.; Charette, A.B. Experimental Evidence for the All-Up Reactive Conformation of Chiral Rhodium(II) Carboxylate Catalysts: Enantioselective Synthesis of cis-Cyclopropane α-Amino Acids. J. Am. Chem. Soc. 2009, 131, 16383–16385. [Google Scholar] [CrossRef] [PubMed]
  55. Davies, H.M.L.; Bruzinski, P.R.; Fall, M.J. Effect of diazoalkane structure on the stereoselectivity of rhodium(II) (S)-N-(arylsulfonyl)Prolinate catalyzed cyclopropanations. Tetrahedron Lett. 1996, 37, 4133–4136. [Google Scholar] [CrossRef]
  56. Deng, L.; Giessert, A.J.; Gerlitz, O.O.; Dai, X.; Diver, S.T.; Davies, H.M. Metal carbene-promoted sequential transformations for the enantioselective synthesis of highly functionalized cycloheptadienes. J. Am. Chem. Soc. 2005, 127, 1342–1343. [Google Scholar] [CrossRef] [PubMed]
  57. Lian, Y.; Davies, H.M. Rhodium-catalyzed [3 + 2] annulation of indoles. J. Am. Chem. Soc. 2009, 132, 440–441. [Google Scholar] [CrossRef] [PubMed]
  58. Briones, J.F.; Hansen, J.; Hardcastle, K.I.; Autschbach, J.; Davies, H.M.L. Highly enantioselective Rh2(S-DOSP)4-catalyzed cyclopropenation of alkynes with styryldiazoacetates. J. Am. Chem. Soc. 2010, 132, 17211–17215. [Google Scholar] [CrossRef] [PubMed]
  59. Parr, B.T.; Green, S.A.; Davies, H.M. Rhodium-catalyzed conversion of furans to highly functionalized pyrroles. J. Am. Chem. Soc. 2013, 135, 4716–4718. [Google Scholar] [CrossRef] [PubMed]
  60. Davies, H.M.; Hansen, T.; Churchill, M.R. Catalytic asymmetric C-H activation of alkanes and tetrahydrofuran. J. Am. Chem. Soc. 2000, 122, 3063–3070. [Google Scholar] [CrossRef]
  61. Davies, H.M.; Ren, P. Catalytic Asymmetric C-H Activation of Silyl Enol Ethers as an Equivalent of an Asymmetric Michael Reaction. J. Am. Chem. Soc. 2001, 123, 2070–2071. [Google Scholar] [CrossRef] [PubMed]
  62. Davies, H.M.; Venkataramani, C. Catalytic Enantioselective Synthesis of β2-Amino Acids. Angew. Chem. Int. Ed. 2002, 41, 2197–2199. [Google Scholar] [CrossRef]
  63. Davies, H.M.; Walji, A.M. Direct Synthesis of (+)-Erogorgiaene through a Kinetic Enantiodifferentiating Step. Angew. Chem. Int. Ed. 2005, 44, 1733–1735. [Google Scholar] [CrossRef] [PubMed]
  64. Davies, H.M.L.; Manning, J.R. C-H Activation as a Strategic Reaction:  Enantioselective Synthesis of 4-Substituted Indoles. J. Am. Chem. Soc. 2006, 128, 1060–1061. [Google Scholar] [CrossRef] [PubMed]
  65. Davies, H.M.L.; Dai, X.; Long, M.S. Combined C-H Activation/Cope Rearrangement as a Strategic Reaction in Organic Synthesis:  Total Synthesis of (−)-Colombiasin A and (−)-Elisapterosin B. J. Am. Chem. Soc. 2006, 128, 2485–2490. [Google Scholar] [CrossRef] [PubMed]
  66. Davies, H.M.L.; Lee, G.H. Dirhodium(II) Tetra(N-(dodecylbenzenesulfonyl)prolinate) Catalyzed Enantioselective Cyclopropenation of Alkynes. Org. Lett. 2004, 6, 1233–1236. [Google Scholar] [CrossRef] [PubMed]
  67. Lian, Y.; Davies, H.M. Combined C-H functionalization/Cope rearrangement with vinyl ethers as a surrogate for the vinylogous Mukaiyama aldol reaction. J. Am. Chem. Soc. 2011, 133, 11940–11943. [Google Scholar] [CrossRef] [PubMed]
  68. Davies, H.M.L.; Townsend, R.J. Catalytic asymmetric cyclopropanation of hetereoaryldiazoacetates. J. Org. Chem. 2001, 66, 6595–6603. [Google Scholar] [CrossRef] [PubMed]
  69. Davies, H.M.; Hutcheson, D.K. Enantioselective synthesis of vinylcyclopropanes by rhodium(II)-catalyzed decomposition of vinyldiazomethanes in the presence of alkenes. Tetrahedron Lett. 1993, 34, 7243–7246. [Google Scholar] [CrossRef]
  70. Davies, H.M.L.; Nagashima, T.; Klino, J.L. Stereoselectivity of Methyl Aryldiazoacetate cyclopropanations of 1,1-diarylethylene. Asymmetric synthesis of cyclopropyl analouge of Tamoxifen. Org. Lett. 2000, 2, 823–826. [Google Scholar] [CrossRef] [PubMed]
  71. Davies, H.M.L.; Panaro, S.A. Novel dirhodium tetraprolinate catalysts containing bridging prolinate ligands for asymmetric carbenoid reactions. Tetrahedron Lett. 1999, 40, 5287–5290. [Google Scholar] [CrossRef]
  72. Davies, H.M.L.; Kong, N. Synthesis and evaluation of a novel dirhodium tetraprolinate catalyst containing bridging prolinate ligands. Tetrahedron Lett. 1997, 38, 4203–4206. [Google Scholar] [CrossRef]
  73. Qin, C.; Boyarskikh, V.; Hansen, J.H.; Hardcastle, K.I.; Musaev, D.G.; Davies, H.M.L. D2-Symmetric Dirhodium Catalyst Derived from a 1,2,2-triarylcyclopropanecarboxylate ligand: Design, Synthesis and Application. J. Am. Chem. Soc. 2011, 133, 19198–19204. [Google Scholar] [CrossRef] [PubMed]
  74. Qin, C.; Davies, H.M.L. Enantioselective synthesis of 2-arylbicyclo[1.1.0]butane carboxylates. Org. Lett. 2013, 15, 310–313. [Google Scholar] [CrossRef] [PubMed]
  75. Qin, C.; Davies, H.M.L. Rh2(R-TPCP)4-catalyzed Enantioselective [3 + 2]-cycloaddition between nitrones and vinyldiazoacetates. J. Am. Chem. Soc. 2013, 135, 14516–14519. [Google Scholar] [CrossRef] [PubMed]
  76. Hashimoto, S.; Watanabe, N.; Sato, T.; Shiro, M.; Ikegami, S. Enhancement of Enantioselectivity in Intramolecular C-H Insertion Reactions of α-Diazo-β-Keto Esters Catalyzed by Chiral Dirhodium(II) Carboxylates. Tetrahedron Lett. 1993, 34, 5109–5112. [Google Scholar] [CrossRef]
  77. Tsutsui, H.; Abe, T.; Nakamura, S.; Anada, M.; Hashimoto, S. Practical Synthesis of Dirhodium(II) Tetrakis[N-phthaloyl-(S)-tert-leucinate]. Chem. Pharm. Bull. 2005, 53, 1366–1368. [Google Scholar] [CrossRef] [PubMed]
  78. DeAngelis, A.; Dmitrenko, O.; Yap, G.P.A.; Fox, J.M. Chiral Crown Conformation of Rh2(PTTL)4: Enantioselective Cyclopropanation with α-Alkyl-α-diazoesters. J. Am. Chem. Soc. 2009, 131, 7230–7231. [Google Scholar] [CrossRef] [PubMed]
  79. Minami, K.; Saito, H.; Tsutsui, H.; Nambu, H.; Anada, M.; Hashimoto, S. Highly Enantio- and Diastereoselective Construction of 1,2-Disubstituted Cyclopentane Compounds by Dirhodium(II) Tetrakis[N-phthaloyl-(S)-tert-leucinate]-Catalyzed C-H Insertion Reactions of α-Diazo Esters. Adv. Synth. Catal. 2005, 347, 1483–1487. [Google Scholar] [CrossRef]
  80. Awata, A.; Arai, T. Catalytic Asymmetric Cyclopropanation with Diazooxindole. Synlett 2013, 24, 29–32. [Google Scholar]
  81. Lindsay, V.N.G.; Nicolas, C.; Charette, A.B. Asymmetric Rh(II)-Catalyzed Cyclopropanation of Alkenes with Diacceptor Diazo Compounds: p-Methoxyphenyl Ketone as a General Stereoselectivity Controlling Group. J. Am. Chem. Soc. 2011, 133, 8972–8981. [Google Scholar] [CrossRef] [PubMed]
  82. Goto, T.; Takada, K.; Anada, M.; Ando, K.; Hashimoto, S. Enantio- and diastereoselective cyclopropanation with tert-butyl α-diazopropionate catalyzed by dirhodium(II) tetrakis[N-tetrabromophthaloyl-(S)-tert-Leucinate]. Tetrahedron Lett. 2011, 52, 4200–4203. [Google Scholar] [CrossRef]
  83. Goto, T.; Takada, K.; Shimada, N.; Nambu, H.; Anada, M.; Shiro, M.; Ando, K.; Hashimoto, S. Highly enantioselective cyclopropenation reaction reaction of 1-Alkynes with α-alkyl-α-diazoesters catalyzed by rhodium(II) carboxylates. Angew. Chem. Int. Ed. 2011, 50, 6803–6808. [Google Scholar] [CrossRef] [PubMed]
  84. Xu, X.; Deng, Y.; Yim, D.N.; Zavalij, P.Y.; Doyle, M.P. Enantioselective cis-β-lactam synthesis by intramolecular C-H functionalization from enoldiazoacetamides and derivative donor–acceptor cyclopropenes. Chem. Sci. 2015, 6, 2196–2201. [Google Scholar] [CrossRef] [PubMed]
  85. Tsutsui, H.; Matsuura, M.; Makino, K.; Nakamura, S.; Nakajima, M.; Kitagaki, S.; Hashimoto, S. Enantioselective Tandem Formation and [2,3]-Sigmatropic Rearrangement of Cyclic Propargylic Oxonium Ylides Catalyzed by Dirhodium(II) Tetrakis[Naphthaloyl-(S)-tert-leucinate]. Isr. J. Chem. 2001, 41, 283–295. [Google Scholar] [CrossRef]
  86. Watanabe, N.; Ohtake, Y.; Hashimoto, S.; Shiro, M.; Ikegami, S. Asymmetric creation of quaternary carbon centers by enantiotopically selective aromatic C-H insertion catalyzed by chiral dirhodium(II) carboxylates. Tetrahedron Lett. 1995, 36, 1491–1494. [Google Scholar] [CrossRef]
  87. DeAngelis, A.; Boruta, D.T.; Lubin, J.-B.; Plampin, J.N.; Yap, G.P.A.; Fox, J.M. The chiral crown conformation in paddlewheel complexes. Chem. Commun. 2010, 46, 4541–4543. [Google Scholar] [CrossRef] [PubMed]
  88. Nadeau, E.; Ventura, D.L.; Brekan, J.A.; Davies, H.M.L. Controlling factors for C-H functionalization versus cyclopropanation of dihydronaphthalenes. J. Org. Chem. 2010, 75, 1927–1939. [Google Scholar] [CrossRef] [PubMed]
  89. Adly, F.G.; Maddalena, J.; Ghanem, A. Rh2(S-1,2-NTTL)4: A Novel Rh2(S-PTTL)4 Analog With Lower Ligand Symmetry for Asymmetric Synthesis of Chiral Cyclopropylphosphonates. Chirality 2014, 26, 764–774. [Google Scholar] [CrossRef] [PubMed]
  90. Kitagaki, S.; Anada, M.; Kataoka, O.; Matsuno, K.; Umeda, C.; Watanabe, N.; Hashimoto, S. Enantiocontrol in Tandem Carbonyl Ylide Formation and Intermolecular 1,3-Dipolar Cycloaddition of α-Diazo Ketones Mediated by Chiral Dirhodium(II) Carboxylate Catalyst. J. Am. Chem. Soc. 1999, 121, 1417–1418. [Google Scholar] [CrossRef]
  91. Denton, J.R.; Sukumaran, D.; Davies, H.M.L. EnantioSelective Synthesis of trifluoromethyl-Substituted Cyclopropanes. Org. Lett. 2007, 9, 2625–2628. [Google Scholar] [CrossRef] [PubMed]
  92. Denton, J.R.; Cheng, K.; Davies, H.M.L. Stereoselective construction of nitrile-substituted cyclopropanes. Chem. Commun. 2008, 1238–1240. [Google Scholar] [CrossRef] [PubMed]
  93. Denton, J.R.; Davies, H.M.L. Enantioselective reactions of donor/acceptor carbenoids derived from α-aryl-α-diazoketones. Org. Lett. 2009, 11, 787–790. [Google Scholar] [CrossRef] [PubMed]
  94. Chepiga, K.M.; Qin, C.; Alford, J.S.; Chennamadhavuni, S.; Gregg, T.M.; Olson, J.P.; Davies, H.M.L. Guide to enantioselective dirhodium(II)-catalyzed cyclopropanation with aryldiazoacetates. Tetrahedron 2013, 69, 5765–5771. [Google Scholar] [CrossRef] [PubMed]
  95. Briones, J.F.; Davies, H.M.L. Rh2(S-PTAD)4 catalyzed asymmetric cyclopropanation of aryl alkynes. Tetrahedron 2011, 67, 4313–4317. [Google Scholar] [CrossRef]
  96. Wang, H.; Guptill, D.M.; Varela-Alvarez, A.; Musaev, D.G.; Davies, H.M.L. Rhodium-catalyzed enantioselective cyclopropanation of electron deficient alkenes. Chem. Sci. 2013, 4, 2844–2850. [Google Scholar] [CrossRef] [PubMed]
  97. Werlé, C.; Goddard, R.; Philipps, P.; Farès, C.; Fürstner, A. Stabilization of a Chiral Dirhodium Carbene by Encapsulation and a Discussion of the Stereochemical Implications. Angew. Chem. Int. Ed. 2016, 55, 10760–10765. [Google Scholar] [CrossRef] [PubMed]
  98. Adly, F.G.; Gardiner, M.G.; Ghanem, A. Design and Synthesis of Novel Chiral Dirhodium(II) Carboxylate Complexes for Asymmetric Cyclopropanation Reactions. Chem.-Eur. J. 2016, 22, 3447–3461. [Google Scholar] [CrossRef] [PubMed]
  99. Mattiza, J.T.; Fohrer, J.G.G.; Duddeck, H.; Gardiner, M.G.; Ghanem, A. Optimizing dirhodium(II) tetrakiscarboxylates as chiral NMR auxiliaries. Org. Biomol. Chem. 2011, 9, 6542–6550. [Google Scholar] [CrossRef] [PubMed]
  100. Gardiner, M.G.; Ghanem, A. X-ray crystal structures of bis(EtOAc) and (MeOH)(H2O) adducts of Rh2(S-PTPA)4 complex. Unpublished work. 2010. [Google Scholar]
  101. Müller, P.; Ghanem, A. Rh(II)-Catalyzed Enantioselective Cyclopropanation of Olefins with Dimethyl Malonate via in Situ Generated Phenyliodonium Ylide. Org. Lett. 2004, 6, 4347–4350. [Google Scholar] [CrossRef] [PubMed]
  102. Ghanem, A.; Gardiner, M.G.; Williamson, R.M.; Müller, P. First X-ray Structure of a N-Naphthaloyl-Tethered Chiral Dirhodium(II) Complex: Structural Basis for Tether Substitution Improving Asymmetric Control in Olefin Cyclopropanation. Chem.-Eur. J. 2010, 16, 3291–3295. [Google Scholar] [CrossRef] [PubMed]
  103. Adly, F.G.; Ghanem, A. Polymer monolith-supported dirhodium(II)-catalyzed continuous flow cyclopropanation in capillary format. Tetrahedron Lett. 2016, 57, 852–857. [Google Scholar] [CrossRef]
  104. Lindsay, V.N.G.; Charette, A.B. Design and Synthesis of Chiral Heteroleptic Rhodium(II) Carboxylate Catalysts: Experimental Investigations of Halogen Bond Rigidification Effects in Asymmetric Cyclopropanation. ACS Catal. 2012, 2, 1221–1225. [Google Scholar] [CrossRef]
  105. Boruta, D.T.; Dmitrenko, O.; Yap, G.P.A.; Fox, J.M. Rh2(S-PTTL)3TPA—A mixed-ligand dirhodium(II) catalyst for enantioselective reactions of α-alkyl-α-diazoesters. Chem. Sci. 2012, 3, 1589–1593. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Dirhodium(II)-mediated chemistry of methyl styryldiazoacetate (1) with different substrates.
Scheme 1. Dirhodium(II)-mediated chemistry of methyl styryldiazoacetate (1) with different substrates.
Catalysts 07 00347 sch001
Figure 1. Main arrangements of ligand’s blocking groups around the dirhodium(II) core and their point groups (the sterically blocking groups around the rhodium active sites are depicted as ovals) [2,4,36,53,54].
Figure 1. Main arrangements of ligand’s blocking groups around the dirhodium(II) core and their point groups (the sterically blocking groups around the rhodium active sites are depicted as ovals) [2,4,36,53,54].
Catalysts 07 00347 g001
Scheme 2. Structure, D2-symmetry and facial enantioselectivity in enantioselective cyclopropanation of styrene with methyl styryldiazoacetate of Rh2(S-DOSP)4 and Rh2(S-biTISP)2 as proposed by Davies et al. (a) Rh2(S-DOSP)4-catalyzed reaction [13,36,69], (b) Rh2(S-biTISP)2-catalyzed reaction [13,71].
Scheme 2. Structure, D2-symmetry and facial enantioselectivity in enantioselective cyclopropanation of styrene with methyl styryldiazoacetate of Rh2(S-DOSP)4 and Rh2(S-biTISP)2 as proposed by Davies et al. (a) Rh2(S-DOSP)4-catalyzed reaction [13,36,69], (b) Rh2(S-biTISP)2-catalyzed reaction [13,71].
Catalysts 07 00347 sch002
Scheme 3. Rh2(R-BTPCP)4-catalyzed cyclopropanation of styrene with methyl styryldiazoacetate and the D2-symmetric structure of Rh2(R-BTPCP)4 according to its X-ray crystal structure [73].
Scheme 3. Rh2(R-BTPCP)4-catalyzed cyclopropanation of styrene with methyl styryldiazoacetate and the D2-symmetric structure of Rh2(R-BTPCP)4 according to its X-ray crystal structure [73].
Catalysts 07 00347 sch003
Scheme 4. Rh2(R-3,5-di(p-tBuPh)TPCP)4-catalysed C-H functionalization of pentane and substituted n-alkanes and the D2-symmetric structure of the catalyst according to its X-ray crystal structure. The two 3,5-di(p-tBuPh)C6H3-groups are highlighted in yellow. (X-ray crystal structure adapted with permission from Macmillan Publishers Ltd.: Liao, K.; Negretti, S.; Musaev, D.G.; Bacsa, J.; Davies, H.M.L. Nature 2016, 533, 230–234, Copyright 2017).
Scheme 4. Rh2(R-3,5-di(p-tBuPh)TPCP)4-catalysed C-H functionalization of pentane and substituted n-alkanes and the D2-symmetric structure of the catalyst according to its X-ray crystal structure. The two 3,5-di(p-tBuPh)C6H3-groups are highlighted in yellow. (X-ray crystal structure adapted with permission from Macmillan Publishers Ltd.: Liao, K.; Negretti, S.; Musaev, D.G.; Bacsa, J.; Davies, H.M.L. Nature 2016, 533, 230–234, Copyright 2017).
Catalysts 07 00347 sch004
Scheme 5. (a) Rh2(S-PTTL)4-catalyzed enantioselective tandem propargylic oxonium ylide formation/[2,3]-sigmatropic rearrangement (Facial selectivity of the Rh2(S-PTTL)4-stabilized carbene according to the model proposed by Hashimoto et al. [85]); (b) Rh2(S-PTPA)4-catalyzed enantioselective aromatic C-H insertion reaction and the C2-symmetric structure of the catalyst according to its reported X-ray crystal structure [76,85,86,87].
Scheme 5. (a) Rh2(S-PTTL)4-catalyzed enantioselective tandem propargylic oxonium ylide formation/[2,3]-sigmatropic rearrangement (Facial selectivity of the Rh2(S-PTTL)4-stabilized carbene according to the model proposed by Hashimoto et al. [85]); (b) Rh2(S-PTPA)4-catalyzed enantioselective aromatic C-H insertion reaction and the C2-symmetric structure of the catalyst according to its reported X-ray crystal structure [76,85,86,87].
Catalysts 07 00347 sch005
Scheme 6. (a) Dirhodium(II)-catalyzed enantioselective tandem carbonyl ylide formation/intermolecular 1,3-dipolar cycloaddition of α-diazoketones [90]; (b) Rh2(S-PTAD)4-catalyzed enantioselective synthesis of dimethyl 1,2-diphenylcyclopropylphosphonate [37].
Scheme 6. (a) Dirhodium(II)-catalyzed enantioselective tandem carbonyl ylide formation/intermolecular 1,3-dipolar cycloaddition of α-diazoketones [90]; (b) Rh2(S-PTAD)4-catalyzed enantioselective synthesis of dimethyl 1,2-diphenylcyclopropylphosphonate [37].
Catalysts 07 00347 sch006
Figure 2. (a) Space filling structure of mono(EtOAc) adduct of Rh2(S-PTTL)4 (top view); Schematic illustration of the rectangular binding pocket of Rh2(S-PTTL)4; (b) Space filling structure of mono(EtOAc) adduct of Rh2(S-PTTL)4 (side view). (Adapted from Adly, F.G.; Gardiner, M.G.; Ghanem, A. Chem. Eur. J. 2016, 22, 3447, Copyright 2017, with permission from John Wiley and Sons).
Figure 2. (a) Space filling structure of mono(EtOAc) adduct of Rh2(S-PTTL)4 (top view); Schematic illustration of the rectangular binding pocket of Rh2(S-PTTL)4; (b) Space filling structure of mono(EtOAc) adduct of Rh2(S-PTTL)4 (side view). (Adapted from Adly, F.G.; Gardiner, M.G.; Ghanem, A. Chem. Eur. J. 2016, 22, 3447, Copyright 2017, with permission from John Wiley and Sons).
Catalysts 07 00347 g002
Scheme 7. Facial selectivity of the Rh2(S-PTTL)4-stabilized carbene according to the model proposed by Fox et al. [78,87].
Scheme 7. Facial selectivity of the Rh2(S-PTTL)4-stabilized carbene according to the model proposed by Fox et al. [78,87].
Catalysts 07 00347 sch007
Figure 3. Space-filling model of the Rh2(S-PTTL)4 = C(MeOC6H4)CO2Me unit (top view). (Reprinted from Werlé, C.; Goddard, R.; Philipps, P.; Farès, C.; Fürstner, A. Angew. Chem. Int. Ed. 2016, 55, 10760, Copyright 2017, with permission from John Wiley and Sons).
Figure 3. Space-filling model of the Rh2(S-PTTL)4 = C(MeOC6H4)CO2Me unit (top view). (Reprinted from Werlé, C.; Goddard, R.; Philipps, P.; Farès, C.; Fürstner, A. Angew. Chem. Int. Ed. 2016, 55, 10760, Copyright 2017, with permission from John Wiley and Sons).
Catalysts 07 00347 g003
Figure 4. Space filling structure of bis(EtOAc) adduct of Rh2(S-PTAD)4 viewed along the Rh-Rh axis (a) into the chiral crown cavity, (b) onto the achiral Rh coordination site shrouded by the adamantyl groups. (c) Side view of bis(EtOAc) adduct of Rh2(S-PTAD)4. (Adapted from Adly, F.G.; Gardiner, M.G.; Ghanem, A. Chem. Eur. J. 2016, 22, 3447, Copyright 2017, with permission from John Wiley and Sons).
Figure 4. Space filling structure of bis(EtOAc) adduct of Rh2(S-PTAD)4 viewed along the Rh-Rh axis (a) into the chiral crown cavity, (b) onto the achiral Rh coordination site shrouded by the adamantyl groups. (c) Side view of bis(EtOAc) adduct of Rh2(S-PTAD)4. (Adapted from Adly, F.G.; Gardiner, M.G.; Ghanem, A. Chem. Eur. J. 2016, 22, 3447, Copyright 2017, with permission from John Wiley and Sons).
Catalysts 07 00347 g004
Scheme 8. Dirhodium(II)-catalyzed enantioselective cyclopropanation of styrene with dimethyl malonate [101] and Meldrum’s acid [102].
Scheme 8. Dirhodium(II)-catalyzed enantioselective cyclopropanation of styrene with dimethyl malonate [101] and Meldrum’s acid [102].
Catalysts 07 00347 sch008
Figure 5. Space filling structure of bis(EtOAc) adduct of Rh2(S-NTTL)4 (top and side view). (Reprinted from Ghanem, A.; Gardiner, M.G.; Williamson, R.M.; Müller, P. Chem. Eur. J. 2010, 16, 3291, Copyright 2017, with permission from John Wiley and Sons).
Figure 5. Space filling structure of bis(EtOAc) adduct of Rh2(S-NTTL)4 (top and side view). (Reprinted from Ghanem, A.; Gardiner, M.G.; Williamson, R.M.; Müller, P. Chem. Eur. J. 2010, 16, 3291, Copyright 2017, with permission from John Wiley and Sons).
Catalysts 07 00347 g005
Scheme 9. Rh2(S-tertPTTL)4-catalyzed enantioselective synthesis of dimethyl 1,2-diphenylcyclopropylphosphonate and X-ray crystal structure of Rh2(S-tertPTTL)4 (top view). (X-ray crystal structure adapted from Adly, F.G.; Gardiner, M.G.; Ghanem, A. Chem. Eur. J. 2016, 22, 3447, Copyright 2017, with permission from John Wiley and Sons).
Scheme 9. Rh2(S-tertPTTL)4-catalyzed enantioselective synthesis of dimethyl 1,2-diphenylcyclopropylphosphonate and X-ray crystal structure of Rh2(S-tertPTTL)4 (top view). (X-ray crystal structure adapted from Adly, F.G.; Gardiner, M.G.; Ghanem, A. Chem. Eur. J. 2016, 22, 3447, Copyright 2017, with permission from John Wiley and Sons).
Catalysts 07 00347 sch009
Scheme 10. Enantioselective cyclopropanation reaction of alkenes with α-nitro diazoacetophenones and X-ray crystal structure of Rh2(S-TCPTTL)3(PTAiB). (X-ray crystal structure adapted with permission from Lindsay, V.N.G.; Charette, A.B. ACS Catal. 2012, 2, 1221. Copyright 2017 American Chemical Society).
Scheme 10. Enantioselective cyclopropanation reaction of alkenes with α-nitro diazoacetophenones and X-ray crystal structure of Rh2(S-TCPTTL)3(PTAiB). (X-ray crystal structure adapted with permission from Lindsay, V.N.G.; Charette, A.B. ACS Catal. 2012, 2, 1221. Copyright 2017 American Chemical Society).
Catalysts 07 00347 sch010
Scheme 11. Enantioselective cyclopropanation of styrene with ethyl α-diazobutanoate and structure of Rh2(S-PTTL)3(TPA) according to its X-ray crystal structure [105].
Scheme 11. Enantioselective cyclopropanation of styrene with ethyl α-diazobutanoate and structure of Rh2(S-PTTL)3(TPA) according to its X-ray crystal structure [105].
Catalysts 07 00347 sch011

Share and Cite

MDPI and ACS Style

Adly, F.G. On the Structure of Chiral Dirhodium(II) Carboxylate Catalysts: Stereoselectivity Relevance and Insights. Catalysts 2017, 7, 347. https://doi.org/10.3390/catal7110347

AMA Style

Adly FG. On the Structure of Chiral Dirhodium(II) Carboxylate Catalysts: Stereoselectivity Relevance and Insights. Catalysts. 2017; 7(11):347. https://doi.org/10.3390/catal7110347

Chicago/Turabian Style

Adly, Frady G. 2017. "On the Structure of Chiral Dirhodium(II) Carboxylate Catalysts: Stereoselectivity Relevance and Insights" Catalysts 7, no. 11: 347. https://doi.org/10.3390/catal7110347

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