Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2021 Apr 8;143(15):5666–5673. doi: 10.1021/jacs.1c01972

A New Ligand Design Based on London Dispersion Empowers Chiral Bismuth–Rhodium Paddlewheel Catalysts

Santanu Singha 1, Michael Buchsteiner 1, Giovanni Bistoni 1, Richard Goddard 1, Alois Fürstner 1,*
PMCID: PMC8154533  PMID: 33829767

Abstract

graphic file with name ja1c01972_0009.jpg

Heterobimetallic bismuth–rhodium paddlewheel complexes with phenylglycine ligands carrying TIPS-groups at the meta-positions of the aromatic ring exhibit outstanding levels of selectivity in reactions of donor/acceptor and donor/donor carbenes; at the same time, the reaction rates are much faster and the substrate scope is considerably wider than those of previous generations of chiral [BiRh] catalysts. As shown by a combined experimental, crystallographic, and computational study, the new catalysts draw their excellent application profile largely from the stabilization of the chiral ligand sphere by London dispersion (LD) interactions of the peripheral silyl substituents.


The first chiral dirhodium carbene characterized by X-ray crystallography was derived from diazoester 4a and [Rh2(PTTL)4] (1a) (Figure 1).14 This particular paddlewheel complex had been chosen not only for its excellent pedigree in asymmetric catalysis526 but also because the reasons for its efficiency had been subject to debate.2733 Interestingly, the four N-phthalimido substituents on the tert-leucine ligands were found to adopt an α,α,α,α-conformation about the carbene ligand occupying an axial site on the dirhodium core; this chiral calyx is maintained in solution.1,27,28 Under the premise that the enantiodetermining transition state features a similar overall structure, the data allowed the sense of induction in the cyclopropanation of styrene to be explained;1 the moderate level of induction (78% ee) can also be rationalized by the fairly wide aperture of the chiral binding site, as observed in the solid state. Moreover, the two Rh atoms reside in notably different ligand environments: the tert-butyl groups form a narrow but essentially “achiral” pore about Rh2: any competing background reaction at this site will reduce the optical purity of the product.1

Figure 1.

Figure 1

Prior art. The “paddles” in the schematic drawing represent the (substituted) phthalimido groups, R = tBu.

In a first foray to translate these insights into an improved catalyst design, the Rh2 center of 1a was formally replaced by Bi(+2).3446 For its larger radius, this main group element imparts a conical shape onto the ligand sphere, which, in turn, tightens the chiral pocket at the Rh-site. Since Bi(+2) does not decompose the diazo ester, the racemic background reaction is essentially shut off. These effects are thought to synergize; they explain why the heterobimetallic complex [BiRh(PTTL)4] (2a) led to much improved ee’s in many cases.34 Limitations were encountered with less electron-rich donor/acceptor diazo derivatives, which gave rather poor results (see below).34

Initial attempts to remedy this issue by tailoring the phthalimido groups largely met with failure. Although perhalogenated phthalimides have proven advantageous in many cases,4750 the heterobimetallic complex 2b gave almost racemic product. The X-ray structure revealed the likely cause: 2b features an inverted calyx, in which the “chiral” environment envelopes the unreactive Bi(+2) site, whereas the Rh-atom sees the tert-butyl substituents.34 The substituted phthalimido groups are more encumbered and hence flank the Bi(+2) center, simply because the larger cation provides more space.34

An improved design needs to provide rigorous control over the directionality of the chiral ligand sphere while allowing the substitution pattern to be fine-tuned. To this end, we planned to employ London dispersion as a noncovalent but attractive interaction to lock the desirable conformation in place. It is increasingly clear that multiple intramolecular contacts can outweigh steric repulsion and be structure-determining;51,52 although this effect might be more prevalent in (organometallic) catalysis than commonly appreciated,51,5356 examples of deliberate use of London dispersion (LD) as a key design principle for the development of new catalysts with improved application profiles are exceedingly rare.5759 Outlined below we present such a case.

To reach this goal, the tert-leucine-derived ligands of 2a,b, which are too far apart for any LD among them,34 were replaced by phenylglycine derivatives carrying TIPS-groups at the meta-positions of the aromatic ring; the latter were expected to mediate a sufficient number of intramolecular contacts to entail stabilization via dispersion interactions (Scheme 1).6062 Commercial benzyl alcohol 6 was O-silylated prior to metal/halogen exchange and quenching of the dilithio species with TIPSCl. Product 7 was elaborated into tert-butylsulfinyl imine 8,63 which reacted with vinylmagnesium bromide in the presence of ZnMe2 to give 9 as a single isomer (dr ≈ 99:1). The auxiliary was removed and the phthalimide of choice64 introduced before the double bond of 10 was cleaved with RuCl3 cat./NaIO4 to avoid racemization. The resulting acids 11(65) were reacted with [BiRh(OCOCF3)4] in boiling toluene and the released trifluoroacetic acid trapped with K2CO3 in a Soxhlet apparatus to furnish complexes 3 in excellent yield.

Scheme 1.

Scheme 1

Reagents and conditions: (a) TIPSCl, DBU, CH2Cl2, 99%; (b) tBuLi, THF, −78 °C → rt, then TIPSCl, −20 °C → rt; (c) TBAF, THF, −20 °C → rt, 74% (over both steps); (d) PCC, CH2Cl2, 94%; (e) (R)-tBuS(=O)NH2, Ti(OEt)4, THF, 70 °C, 84%; (f) vinylmagnesium bromide, Me2Zn (50 mol %), THF, −78 °C, 87% (dr = 99:1); (g) (i) HCl/1,4-dioxane, MeOH; (ii) aq. NaOH, 90% (iii) (substituted) phthalic anhydride, Et3N, toluene, reflux, 75% (10a), 73% (10b); (h) RuCl3·H2O (5 mol %), NaIO4, CCl4, MeCN, H2O, 74% (11a, 96% ee), 69% (11b, 98% ee); (i) [BiRh(OCOCF3)4], toluene, reflux, 81% (3a), 94% (3b).

Although 3a crystallizes well, the many degrees of rotational freedom of the eight peripheral TIPS groups cause disorder. This complication notwithstanding, a data set meeting good crystallographic standards was obtained (see the Supporting Information (SI)). As expected, complex 3a adopts an α,α,α,α-conformation, in which the silylated aryl rings envelope the Bi-center whereas the phthalimides form a narrow chiral pocket about Rh (Figure 2). The steric demand of the TIPS groups is manifested in innumerous short H···H and H···C contacts in the periphery. The crystallographic data, however, do not allow us to decide whether these contacts mediate LD and are hence stabilizing, or whether they are repulsive as a consequence of steric hindrance.

Figure 2.

Figure 2

Structure of complex 3a in the solid state; disordered parts are shown with dashed bonds, H atoms omitted for clarity; for further information, see the SI.

This aspect was addressed by DFT calculations.66,67 To this end, the structures of 3a,b were computed with and without D3 (BJ) dispersion correction at the PBE level of theory (see the SI).68,69 Details apart, the computations are unambiguous in that dispersion renders the structures notably more compact. In energetic terms, it entails stabilization of 3a and 3b by no less than −9.9 and −11.6 kcal·mol–1, respectively.70,71 LD is therefore clearly a major structure determinant.72 Deconvolution validates the original design concept:73 the TIPS-groups account for ∼32% of ΔEdisp, whereas the contribution of the tBu groups in 3b is smaller but still appreciable (∼12%); in structural terms, however, this extra factor visibly tightens the chiral binding site on the Rh-face in 3b (Figure 3).

Figure 3.

Figure 3

Side-view of the structure of 3b as computed at the PBE/def2-SVP level of theory without (top) and with (bottom) D3 (BJ) dispersion correction

To explore the catalytic performance of the new heterobimetallic complexes, the cyclopropanation of styrene with the fluorinated diazo derivative 4b was chosen as a test reaction (Table 1). Neither [Rh2(PTTL)4] (1a) nor [BiRh(PTTL)4] (2a) gave good results,34 whereas the new complex 3b furnished product 5b (R = Me) with 90% ee. The corresponding trichloroethyl ester proved even more successful:74,75 once again, replacement of 2a by either of the new complexes led to massive improvements. The fact that 3a and 3b both perform very well but 3b is the better of the two is in excellent accord with the conclusions drawn from the DFT calculations: the peripheral −TIPS groups make the larger contribution to the stabilization of the chiral ligand sphere, but the tBu– substituents on the phthalimide residues pay an additional dividend. This notion is further supported by a control experiment with a catalyst analogous to 3a but lacking the peripheral TIPS-groups, which gave product 5b′ with only 24% ee (see the SI). The fact that the outcome is fairly independent of the medium (pentane versus CH2Cl2) can be taken as additional indirect evidence for LD being operative in solution.76 Under the optimized conditions, product 5b′ was formed in essentially quantitative yield with 98% ee as a single diastereomer.

Table 1. Screeninga.

graphic file with name ja1c01972_0006.jpg

Entry R Catalyst Mol % ee
1 Me 1a 1 32%b
2 Me 2a 1 35%b
3 Me 3b 1 90%b
4 CH2CCl3 2a 1 58%
5 CH2CCl3 3a 1 91%
6 CH2CCl3 3b 1 95%
7 CH2CCl3 3b 1 94%c
8 CH2CCl3 3b 1 98%b
9 CH2CCl3 3b 0.1 98%b,e
10 CH2CCl3 3b 0.005 97%b,e
11 CH2CCl3 3b 0.001 n.d.d
a

All reactions were performed in pentane at RT; the yield of product (NMR) was ≥90%.

b

At −10 °C.

c

In CH2Cl2.

d

Incomplete conversion after 24 h.

e

1 mmol scale

Formal replacement of a Rh(+2) center by Bi(+2) usually entails a loss in reactivity.3438 When seen against this backdrop, the rate with which cyclopropane 5a was formed by 3 even at −10 °C is truly remarkable (Figure 4): conversion was complete within 10 min at 0.25 mol % catalyst loading, whereas it took >3 h with the parent complex 2a.77 The fact that the visibly more crowded catalysts lead to notably higher rates—in contrast to what one would expect from a confined space—indicates stabilizing LD interactions with the incoming substrate. The new design hence entails excellent selectivity and reactivity at the same time. The high solubility of 3a,b in the common solvents even at low temperatures is an additional asset.

Figure 4.

Figure 4

Kinetic profiling: Formation of 5a using different BiRh-paddlewheel complexes (0.25 mol %, pentane, −10 °C).

These rewarding results suggested that it should be possible to reduce the loading to partly compensate for the high molecular weights of 3a,b. Indeed, no drop in productivity or optical purity was noticed—without any further optimization of the conditions—when the reaction was performed with only 0.005 mol % of 3b on a mmol scale (Table 1, entry 10).78 The robustness of these conditions was confirmed by the additional example shown in Table 2 (entry 6, 1.3 g scale).79

Table 2. Asymmetric [2 + 1] Cycloadditionsa,b.

graphic file with name ja1c01972_0007.jpg

a

Color code: black, 2a; red, 3b.

b

Pentane, −10 °C, 1 mol % catalyst loading.

c

In CH2Cl2/pentane.

d

At rt.

e

In CH2Cl2

The scope is broad, the level of induction is invariably excellent, and the reactions are typically very clean; therefore, basically quantitative yields were obtained in most cases (Table 2). α-Aryl-α-diazo esters comprising a strong donor substituent perform so well with the parent complex [BiRh(S-PTTL)4] (2a) that there was no real need to resort to 3a,b (entries 1–4).34 However, diazoesters comprising less electron-rich aryl groups strongly benefitted from their use: in all cases investigated, 3b led to significantly higher levels of induction, with ee’s often approaching 99% (entries 5–21). This striking invariance of the results upon substantial modulation of the diazo compound is deemed another favorable distinguishing feature of the new catalyst.80

Excellent results were also obtained with donor/acceptor α-diazo ketones (entries 24, 25), which previously required forcing conditions and recourse to special media.81 With catalyst 3b, these challenging substrates react even at −10 °C. Likewise, donor/donor carbenes are well behaved,2,82,83 again furnishing the corresponding cyclopropanes with outstanding levels of enantio- and diastereoselectivity (entries 26–28).

In addition to (substituted) styrenes, many other terminal alkenes proved suitable, including vinyl benzoate, allyltrimethylsilane, and allyl(pinacol)boronate, some of which have been rarely used before. Even allyl bromide reacts cleanly (entry 14): we are unaware of any prior use in asymmetric cyclopropanation. The same is true for exo-methylenecyclobutane and -hexane, which furnish spirocyclic products (entries 22, 23); in these cases, the superiority of 3b is particularly evident.

Although less comprehensive, the foray into the cyclopropenation of terminal alkynes was no less gratifying (entries 29–32). Most notable is the use of propargyl chloride, which leads to a multifunctional building block that promises rich downstream chemistry. Equally remarkable and largely unprecedented is the fact that cyclopropenation of a terminal alkyne is faster than insertion of the carbene intermediate into an unprotected alcohol (entry 32).8487 To put these results into perspective, one has to note that [Rh2(esp)2]88 or [Rh2(OTpa)4] (OTpa = triphenylacetate) as some of the most proficient achiral catalysts known to date furnished only complex mixtures on reaction with these functionalized alkynes.89

Finally, the new catalysts were briefly examined in reactions other than [2 + 1] cycloadditions. Although a more systematic foray has to await further studies, the preliminary data obtained for prototype C–H and Si–H insertion reactions are promising (Table 3). When compared with the current state-of-the-art,90 the formation of a benzodihydrofuran by intramolecular insertion of a donor/donor carbene into a methyl ether makes a particularly compelling case (entry 2). Whereas 2a performed poorly in the particularly taxing reaction of 4a with Et3SiH, 3b resulted in clean conversion and an excellent overall result.11,9198

Table 3. C–H and Si–H Insertion Reactionsa.

graphic file with name ja1c01972_0008.jpg

a

Color code as shown in Table 2; pentane, −10 °C, 1 mol % catalyst loading.

b

In cyclohexane.

c

With [Rh2(R-PTAD)4]: 56% ee (49%).90

In summary, a new class of heterobimetallic paddlewheel complexes is presented, which take advantage of LD to stabilize the shape and directionality of the calyx that forms the chiral binding pocket. 3a,b as the parent complexes of this series excel in cyclopropanation, cyclopropenation, and certain insertion reactions; as many variations of the conceptually new design can be envisaged, one may have high expectations for future generations of catalysts of this type. Opportunities along these and related lines99,100 are actively pursued in our laboratory.

Acknowledgments

Generous financial support by the Max-Planck-Gesellschaft is gratefully acknowledged. We thank Dr. L. Collins and Mr. S. Auris for exploratory studies, Dr. M. Leutzsch for help with the kinetic experiments, and the Analytical Departments of this Institute for excellent support.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c01972.

  • Experimental Section containing supporting crystallographic data, computational data, characterization data, and NMR spectra of new compounds (PDF)

Accession Codes

CCDC 2063745–2063748 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Author Contributions

§ S.S. and M.B. contributed equally

The authors declare no competing financial interest.

Supplementary Material

ja1c01972_si_001.pdf (14.1MB, pdf)

References

  1. 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. 10.1002/anie.201605502. [DOI] [PubMed] [Google Scholar]
  2. Werlé C.; Goddard R.; Fürstner A. The First Crystal Structure of a Reactive Dirhodium Carbene Complex and a Verstatile New Method for the Preparation of Gold Carbenes by Rhodium-to-Gold Transmetalation. Angew. Chem., Int. Ed. 2015, 54, 15452–15456. 10.1002/anie.201506902. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Werlé C.; Goddard R.; Philipps P.; Farès C.; Fürstner A. Structures of Reactive Donor/Acceptor and Donor/Donor Rhodium Carbenes in the Solid State and Their Implication for Catalysis. J. Am. Chem. Soc. 2016, 138, 3797–3805. 10.1021/jacs.5b13321. [DOI] [PubMed] [Google Scholar]
  4. Tindall D. J.; Werlé C.; Goddard R.; Philipps P.; Farès C.; Fürstner A. Structure and Reactivity of Half-Sandwich Rh(+3) and Ir(+3) Carbene Complexes. Catalytic Metathesis of Azobenzene Derivatives. J. Am. Chem. Soc. 2018, 140, 1884–1893. 10.1021/jacs.7b12673. [DOI] [PubMed] [Google Scholar]
  5. Hashimoto S.; Watanabe N.; Ikegami S. Enantioselective Intramolecular C-H Insertion of α-Diazo-β-keto Esters Catalyzed by Homochiral Rhodium(II) Carboxylates. Tetrahedron Lett. 1990, 31, 5173–5174. 10.1016/S0040-4039(00)97834-1. [DOI] [Google Scholar]
  6. Watanabe N.; Ogawa T.; Ohtake Y.; Ikegami S.; Hashimoto S. Dirhodium(II) Tetrakis[N-phthaloyl-(S)-tert-leucinate]: A Notable Catalyst for Enantiotopically Selective Aromatic Substitution Reactions of α-Diazocarbonyl Compounds. Synlett 1996, 1996, 85–86. 10.1055/s-1996-5336. [DOI] [Google Scholar]
  7. 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. 10.1021/ja983748e. [DOI] [Google Scholar]
  8. 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. 10.1248/cpb.53.1366. [DOI] [PubMed] [Google Scholar]
  9. For reviews, see refs (10−26) and the following:Doyle M. P.; McKervey M. A.; Ye T.. Modern Catalytic Methods for Organic Synthesis with Diazo Compounds: From Cyclopropanes to Ylides; Wiley: New York, 1998. [Google Scholar]
  10. Doyle M. P. Catalytic Methods for Metal Carbene Transformations. Chem. Rev. 1986, 86, 919–939. 10.1021/cr00075a013. [DOI] [PubMed] [Google Scholar]
  11. Ye T.; McKervey M. A. Organic Synthesis with α-Diazo Carbonyl Compounds. Chem. Rev. 1994, 94, 1091–1160. 10.1021/cr00028a010. [DOI] [Google Scholar]
  12. Doyle M. P.; Forbes D. C. Recent Advances in Asymmetric Catalytic Metal Carbene Transformations. Chem. Rev. 1998, 98, 911–935. 10.1021/cr940066a. [DOI] [PubMed] [Google Scholar]
  13. Padwa A.; Weingarten M. D. Cascade Processes of Metallo Carbenoids. Chem. Rev. 1996, 96, 223–269. 10.1021/cr950022h. [DOI] [PubMed] [Google Scholar]
  14. Davies H. M. L.; Antoulinakis E. G. Intermolecular Metal-Catalyzed Carbenoid Cyclopropanations. Org. React. 2001, 57, 1–326. 10.1002/0471264180.or057.01. [DOI] [Google Scholar]
  15. Lebel H.; Marcoux J.-F.; Molinaro C.; Charette A. B. Stereoselective Cyclopropanation Reactions. Chem. Rev. 2003, 103, 977–1050. 10.1021/cr010007e. [DOI] [PubMed] [Google Scholar]
  16. Zhang Z.; Wang J. Recent Studies on the Reactions of α-Diazocarbonyl Compounds. Tetrahedron 2008, 64, 6577–6605. 10.1016/j.tet.2008.04.074. [DOI] [Google Scholar]
  17. Zhang Y.; Wang J. Catalytic [2,3]-Sigmatropic Rearrangement of Sulfur Ylide Derived from Metal Carbene. Coord. Chem. Rev. 2010, 254, 941–953. 10.1016/j.ccr.2009.12.005. [DOI] [Google Scholar]
  18. Ford A.; Miel H.; Ring A.; Slattery C. N.; Maguire A. R.; McKervey M. A. Modern Organic Synthesis with α-Diazocarbonyl Compounds. Chem. Rev. 2015, 115, 9981–10080. 10.1021/acs.chemrev.5b00121. [DOI] [PubMed] [Google Scholar]
  19. DeAngelis A.; Panish R.; Fox J. M. Rh-Catalyzed Intermolecular Reactions of α-Alkyl-α-Diazo Carbonyl Compounds with Selectivity over β-Hydride Migration. Acc. Chem. Res. 2016, 49, 115–127. 10.1021/acs.accounts.5b00425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Gois P. M. P.; Afonso C. A. M. Stereo- and Regiocontrol in the Formation of Lactams by Rhodium-Carbenoid C–H Insertion of α-Diazoacetamides. Eur. J. Org. Chem. 2004, 2004, 3773–3788. 10.1002/ejoc.200400237. [DOI] [Google Scholar]
  21. Davies H. M. L.; Manning J. R. Catalytic C–H Functionalization by Metal Carbenoid and Nitrenoid Insertion. Nature 2008, 451, 417–424. 10.1038/nature06485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Doyle M. P.; Duffy R.; Ratnikov M.; Zhou L. Catalytic Carbene Insertion into C–H Bonds. Chem. Rev. 2010, 110, 704–724. 10.1021/cr900239n. [DOI] [PubMed] [Google Scholar]
  23. Zhu S.-F.; Zhou Q.-L. Transition-Metal-Catalyzed Enantioselective Heteroatom-Hydrogen Bond Insertion Reactions. Acc. Chem. Res. 2012, 45, 1365–1377. 10.1021/ar300051u. [DOI] [PubMed] [Google Scholar]
  24. Davies H. M. L.; Lian Y. The Combined C–H Functionalization/Cope Rearrangement: Discovery and Applications in Organic Synthesis. Acc. Chem. Res. 2012, 45, 923–935. 10.1021/ar300013t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Gillingham D.; Fei N. Catalytic X–H Insertion Reactions Based on Carbenoids. Chem. Soc. Rev. 2013, 42, 4918–4931. 10.1039/c3cs35496b. [DOI] [PubMed] [Google Scholar]
  26. Caballero A.; Díaz-Requejo M. M.; Fructos M. R.; Olmos A.; Urbano J.; Pérez P. J. Catalytic Functionalization of Low Reactive C(sp3)–H and C(sp2)–H Bonds of Alkanes and Arenes by Carbene Transfer from Diazo Compounds. Dalton Trans. 2015, 44, 20295–20307. 10.1039/C5DT03450G. [DOI] [PubMed] [Google Scholar]
  27. Fox and coworkers were the first to propose that an “all up” conformation governs asymmetric cyclopropanations using 1a and related catalysts; see ref (28) and the following:DeAngelis A.; Dmitrenko O.; Yap G. P. A.; Fox J. M. Chiral Crown Conformation of Rh2(S-PTTL)4: Enantioselective Cyclopropanation with α-Alkyl-α-diazoesters. J. Am. Chem. Soc. 2009, 131, 7230–7231. 10.1021/ja9026852. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. 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. 10.1039/c001557a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. See also: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. 10.1021/ja9044955. [DOI] [PubMed] [Google Scholar]
  30. 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. 10.1002/chem.200903231. [DOI] [PubMed] [Google Scholar]
  31. Goto T.; Takeda K.; Shimada N.; Nambu H.; Anada M.; Shiro M.; Ando K.; Hashimoto S. Highly Enantioselective Cyclopropenation Reaction of 1-Alkynes with α-Alkyl-α-diazoesters Catalyzed by Dirhodium(II) Carboxylates. Angew. Chem., Int. Ed. 2011, 50, 6803–6808. 10.1002/anie.201101905. [DOI] [PubMed] [Google Scholar]
  32. Xue Y.-S.; Cai Y.-P.; Chen Z.-X. Mechanism and Stereoselectivity of the Rh(II)-Catalyzed Cyclopropanation of Diazooxindole: A Density Functional Theory Study. RSC Adv. 2015, 5, 57781–57791. 10.1039/C5RA07981K. [DOI] [Google Scholar]
  33. Adly F. G. On the Structure of Chiral Dirhodium(II) Carboxylate Catalysts: Stereoselectivity Relevance and Insights. Catalysts 2017, 7, 347. 10.3390/catal7110347. [DOI] [Google Scholar]
  34. Collins L. R.; Auris S.; Goddard R.; Fürstner A. Chiral Heterobimetallic Bismuth-Rhodium Paddlewheel Catalysts: A Conceptually New Approach to Asymmetric Cyclopropanation. Angew. Chem., Int. Ed. 2019, 58, 3557–3561. 10.1002/anie.201900265. [DOI] [PubMed] [Google Scholar]
  35. For an alternative chiral [BiRh] complex, see:Ren Z.; Sunderland T. L.; Tortoreto C.; Yang T.; Berry J. F.; Musaev D. G.; Davies H. M. L. Comparison of Reactivity and Enantioselectivity between Chiral Bimetallic Catalysts: Bismuth-Rhodium- and Dirhodium-Catalyzed Carbene Chemistry. ACS Catal. 2018, 8, 10676–10682. 10.1021/acscatal.8b03054. [DOI] [Google Scholar]
  36. For achiral [BiRh] complexes, see refs (37−42) and the following:Dikarev E. V.; Gray T. G.; Li B. Heterobimetallic Main-Group-Transition-Metal Paddle-Wheel Carboxylates. Angew. Chem., Int. Ed. 2005, 44, 1721–1724. 10.1002/anie.200462433. [DOI] [PubMed] [Google Scholar]
  37. Dikarev E. V.; Li B.; Zhang H. Tuning the Properties at Heterobimetallic Core: Mixed-Ligand Bismuth–Rhodium Paddlewheel Carboxylates. J. Am. Chem. Soc. 2006, 128, 2814–2815. 10.1021/ja058294h. [DOI] [PubMed] [Google Scholar]
  38. Hansen J.; Li B.; Dikarev E.; Autschbach J.; Davies H. M. L. Combined Experimental and Computational Studies of Heterobimetallic Bi–Rh Paddlewheel Carboxylates as Catalysts for Metal Carbenoid Transformations. J. Org. Chem. 2009, 74, 6564–6571. 10.1021/jo900998s. [DOI] [PubMed] [Google Scholar]
  39. Filatov A. S.; Napier M.; Vreshch V. D.; Sumner N. J.; Dikarev E. V.; Petrukhina M. A. From Solid State to Solution: Advancing Chemistry of Bi–Bi and Bi–Rh Paddlewheel Carboxylates. Inorg. Chem. 2012, 51, 566–571. 10.1021/ic202089p. [DOI] [PubMed] [Google Scholar]
  40. Sunderland T. L.; Berry J. F. Metal–Metal Single Bonds with the Magnetic Anisotropy of Quadruple Bonds: A Systematic Series of Heterobimetallic Bismuth(II)–Rhodium(II) Formamidinate Complexes. Chem. - Eur. J. 2016, 22, 18564–18571. 10.1002/chem.201604007. [DOI] [PubMed] [Google Scholar]
  41. Sunderland T. L.; Berry J. F. Expanding the Family of Heterobimetallic Bi–Rh Paddlewheel Carboxylate Complexes via Equatorial Carboxylate Exchange. Dalton Trans. 2016, 45, 50–55. 10.1039/C5DT03740A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Sunderland T. L.; Berry J. F. The First Bismuth(II)-Rhodium(II) Oxopyridinate Paddlewheel Complexes: Synthesis and Structural Characterization. J. Coord. Chem. 2016, 69, 1949–1956. 10.1080/00958972.2016.1177178. [DOI] [Google Scholar]
  43. Collins L. R.; van Gastel M.; Neese F.; Fürstner A. Enhanced Electrophilicity of Heterobimetallic Bi–Rh Paddlewheel Carbene Complexes: A Combined Experimental, Spectroscopic and Computational Study. J. Am. Chem. Soc. 2018, 140, 13042–13055. 10.1021/jacs.8b08384. [DOI] [PubMed] [Google Scholar]
  44. For general treatises of metal-metal bonded catalysts, see refs (45), (46), and the following:Powers I. G.; Uyeda C. Metal-Metal Bonds in Catalysis. ACS Catal. 2017, 7, 936–958. 10.1021/acscatal.6b02692. [DOI] [Google Scholar]
  45. Berry J. F.; Lu C. C. Metal–Metal Bonds: From Fundamentals to Applications. Inorg. Chem. 2017, 56, 7577–7581. 10.1021/acs.inorgchem.7b01330. [DOI] [PubMed] [Google Scholar]
  46. Takaya J. Catalysis Using Transition Metal Complexes Featuring Main Group Metal and Metalloid Compounds as Supporting Ligands. Chem. Sci. 2021, 12, 1964–1981. 10.1039/D0SC04238B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Yamawaki M.; Tsutsui H.; Kitagaki S.; Anada M.; Hashimoto S. Dirhodium(II) Tetrakis[N-tetrachlorophthaloyl-(S)-tert-leucinate]: A New Chiral Rh(II) Catalyst for Enantioselective Amidation of C-H Bonds. Tetrahedron Lett. 2002, 43, 9561–9564. 10.1016/S0040-4039(02)02432-2. [DOI] [Google Scholar]
  48. Goto T.; Takeda 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. 10.1016/j.tetlet.2011.06.008. [DOI] [Google Scholar]
  49. Tsutsui H.; Yamaguchi Y.; Kitagaki S.; Nakamura S.; Anada M.; Hashimoto S. Dirhodium(II) Tetrakis[N-tetrafluorophthaloyl-(S)-tert-leucinate]: An Exceptionally Effective Rh(II) Catalyst for Enantiotopically Selective Aromatic C–H Insertions of Diazo Ketoesters. Tetrahedron: Asymmetry 2003, 14, 817–821. 10.1016/S0957-4166(03)00075-2. [DOI] [Google Scholar]
  50. This is not only true for dirhodium but also for diruthenium paddlewheel complexes; see:Miyazawa T.; Suzuki T.; Kumagai Y.; Takizawa K.; Kikuchi T.; Kato S.; Onoda A.; Hayashi T.; Kamei Y.; Kamiyama F.; Anada M.; Kojima M.; Yoshino T.; Matsunaga S. Chiral Paddle-wheel Diruthenium Complexes for Asymmetric Catalysis. Nature Catal. 2020, 3, 851–858. 10.1038/s41929-020-00513-w. [DOI] [Google Scholar]
  51. Wagner J. P.; Schreiner P. R. London Dispersion in Molecular Chemistry – Reconsidering Steric Effects. Angew. Chem., Int. Ed. 2015, 54, 12274–12296. 10.1002/anie.201503476. [DOI] [PubMed] [Google Scholar]
  52. Liptrot D. J.; Power P. P. London Dispersion Forces in Sterically Crowded Inorganic and Organometallic Molecules. Nat. Rev. Chem. 2017, 1, 0004. 10.1038/s41570-016-0004. [DOI] [Google Scholar]
  53. For established catalysts, in which dispersion was recognized as being critically important, see refs (54−56) and the following:Lyngvi E.; Sanhueza I. A.; Schoenebeck F. Dispersion Makes the Difference: Bisligated Transition States Found for the Oxidative Addition of Pd(PtBu3)2 to Ar-OSO2R and Dispersion-Controlled Chemoselectivity in Reactions with Pd[P(iPr)(tBu2)]2. Organometallics 2015, 34, 805–812. 10.1021/om501199t. [DOI] [Google Scholar]
  54. Saper N. I.; Ohgi A.; Small D. W.; Semba K.; Nakao Y.; Hartwig J. F. Nickel-catalysed anti-Markovnikov Hydroarylation of Unactivated Alkenes with Unactivated Arenes Facilitated by Noncovalent Interactions. Nat. Chem. 2020, 12, 276–283. 10.1038/s41557-019-0409-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Yepes D.; Neese F.; List B.; Bistoni G. Unveiling the Delicate Balance of Steric and Dispersion Interactions in Organocatalysis Using High-Level Computational Methods. J. Am. Chem. Soc. 2020, 142, 3613–3625. 10.1021/jacs.9b13725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Eschmann C.; Song L.; Schreiner P. R. London Dispersion Interactions Rather than Steric Hindrance Determine the Enantioselectivity of the Corey-Bakshi-Shibata Reduction. Angew. Chem., Int. Ed. 2021, 60, 4823–4832. 10.1002/anie.202012760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. The closest examples are those in which intermolecular London dispersion interactions enhance the interaction of a catalyst with the substrate; see refs (58), (59), and the following:Lu G.; Liu R. Y.; Yang Y.; Fang C.; Lambrecht D. S.; Buchwald S. L.; Liu P. Ligand-Substrate Dispersion Facilitates the Copper-Catalyzed Hydroamination of Unactivated Olefins. J. Am. Chem. Soc. 2017, 139, 16548–16555. 10.1021/jacs.7b07373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Thomas A. A.; Speck K.; Kevlishvili I.; Lu Z.; Liu P.; Buchwald S. L. Mechanistically Guided Design of Ligands that Significantly Improve the Efficiency of CuH-Catalyzed Hydroamination Reactions. J. Am. Chem. Soc. 2018, 140, 13976–13984. 10.1021/jacs.8b09565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Xi Y.; Su B.; Qi X.; Pedram S.; Liu P.; Hartwig J. F. Application of Trimethylgermanyl-Substituted Bisphosphine Ligands with Enhanded Dispersion Interactions to Copper-Catalyzed Hydroboration of Disubstituted Alkenes. J. Am. Chem. Soc. 2020, 142, 18213–18222. 10.1021/jacs.0c08746. [DOI] [PubMed] [Google Scholar]
  60. Dirhodium complexes bearing ligands derived from parent phenylglycine gave poor results in the few documented applications; see refs (61), (62), and the following:Pierson N.; Fernández-Garcia C.; McKervey M. A. Catalytic Asymmetric Oxonium Ylide – [2,3] Sigmatropic Rearrangement with Diazocarbonyl Compounds: First Use of C2-Symmetry in Rh(II) Carboxylates. Tetrahedron Lett. 1997, 38, 4705–4708. 10.1016/S0040-4039(97)01003-4. [DOI] [Google Scholar]
  61. 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. 10.1039/c1ob05665d. [DOI] [PubMed] [Google Scholar]
  62. Buck R. T.; Coe D. M.; Drysdale M. J.; Ferris L.; Haigh D.; Moody C. J.; Pearson N. D.; Sanghera J. B. Asymmetric Rhodium Carbene Insertion into the Si–H Bond: Identification of New Dirhodium(II) Carboxylate Catalysts Using Parallel Synthesis Techniques. Tetrahedron: Asymmetry 2003, 14, 791–816. 10.1016/S0957-4166(03)00035-1. [DOI] [Google Scholar]
  63. Robak M. T.; Herbage M. A.; Ellman J. A. Synthesis and Applications of tert-Butanesulfinamide. Chem. Rev. 2010, 110, 3600–3740. 10.1021/cr900382t. [DOI] [PubMed] [Google Scholar]
  64. For the use of tert-butyl phthalimide substituents, see: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. 10.1002/chem.201504817. [DOI] [PubMed] [Google Scholar]
  65. Separation of the enantiomers by HPLC on a chiral stationary phase is an alternative entry into optically pure ligand.
  66. Neese F. Software Update: The Orca Program System, Version 4.0. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2018, 8, e1327. 10.1002/wcms.1327. [DOI] [Google Scholar]
  67. Weigend F.; Ahlrichs R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
  68. Grimme S.; Hansen A.; Brandenburg J. G.; Bannwarth C. Dispersion-Corrected Mean-Field Electronic Structure Methods. Chem. Rev. 2016, 116, 5105–5154. 10.1021/acs.chemrev.5b00533. [DOI] [PubMed] [Google Scholar]
  69. Becke A. D.; Johnson E. R. A Density-Functional Model of the Dispersion Interaction. J. Chem. Phys. 2005, 123, 154101. 10.1063/1.2065267. [DOI] [PubMed] [Google Scholar]
  70. A large number of benchmark studies showed that the D3 dispersion typically provides a lower bound for the coupled cluster dispersion computed using accurate techniques, see refs (68), (71), and the following:Bistoni G. Finding Chemical Concepts in the Hilbert Space: Coupled Cluster Analyses of Noncovalent Interactions. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2020, 10, e1442. 10.1002/wcms.1442. [DOI] [Google Scholar]
  71. Schneider W. B.; Bistoni G.; Sparta M.; Saitow M.; Riplinger C.; Auer A. A.; Neese F. Decomposition of Intermolecular Interaction Energies within the Local Pair Natural Orbital Coupled Cluster Framework. J. Chem. Theory Comput. 2016, 12, 4778–4792. 10.1021/acs.jctc.6b00523. [DOI] [PubMed] [Google Scholar]
  72. The computed and the experimental structure of 3a show good congruence; the comparison is slightly obscured by diglyme coordinated to the axial site at Rh in the X-ray structure and the presence of solute in the unit cell.
  73. ΔEdisp = ΔEdisp(TIPS) + ΔEdisp(tBu) + ΔEdisp(rest); for details, see the SI.
  74. Guptill D. M.; Davies H. M. L. 2,2,2-Trichloroethyl Aryldiazoacetates as Robust Reagents for the Enantioselecive C–H Functionalization of Methyl Ethers. J. Am. Chem. Soc. 2014, 136, 17718–17721. 10.1021/ja5107404. [DOI] [PubMed] [Google Scholar]
  75. Liao K.; Negretti S.; Musaev D. G.; Bacsa J.; Davies H. M. L. Site-selective and Stereoselective Functionalization of Unactivated C–H Bonds. Nature 2016, 533, 230–234. 10.1038/nature17651. [DOI] [PubMed] [Google Scholar]
  76. Schümann J. M.; Wagner J. P.; Eckhardt A. K.; Quanz H.; Schreiner P. R. Intramolecular London Dispersion Interactions Do Not Cancel in Solution. J. Am. Chem. Soc. 2021, 143, 41–45. 10.1021/jacs.0c09597. [DOI] [PubMed] [Google Scholar]
  77. The methyl and the trichloroethyl esters 4a and 4a′ react with almost the same rates; as expected for a donor/acceptor carbene, however, change of the aryl substituents exerts a marked effect; for details, see the SI.
  78. For even lower loadings under optimized conditions, see:Wei B.; Sharland J. C.; Lin P.; Wilkerson-Hill S. M.; Fullilove F. A.; McKinnon S.; Blackmond D. G.; Davies H. M. L. In Situ Kinetic Studies of Rh(II)-Catalyzed Asymmetric Cyclopropanation with Low Catalyst Loadings. ACS Catal. 2020, 10, 1161–1170. 10.1021/acscatal.9b04595. [DOI] [Google Scholar]
  79. The absolute configuration of this and two additional products was determined by X-ray diffraction (Flack absolute structure parameter); see the SI.
  80. In comparison, the ee’s in cyclopropanations with a similar set of α-aryl-α-diazoesters catalyzed by [Rh2(PTAD)4] are much less uniform: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. 10.1016/j.tet.2013.04.075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. 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. 10.1021/ol802614j. [DOI] [PubMed] [Google Scholar]
  82. Bergstrom B. D.; Nickerson L. A.; Shaw J. T.; Souza L. W. Transition Metal Catalyzed Insertion Reactions with Donor/Donor Carbenes. Angew. Chem., Int. Ed. 2021, 60, 6864–6878. 10.1002/anie.202007001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Zhu D.; Chen L.; Fan H.; Yao Q.; Zhu S. Recent Progress on Donor and Donor-Donor Carbenes. Chem. Soc. Rev. 2020, 49, 908–950. 10.1039/C9CS00542K. [DOI] [PubMed] [Google Scholar]
  84. We are aware of a single example using a free carbene generated by light-induced diazo decomposition, which afforded racemic product; see:He F.; Koenigs R. M. Visible Light Mediated, Metal-free Carbene Transfer Reactions of Diazoalkanes with Propargylic Alcohols. Chem. Commun. 2019, 55, 4881–4884. 10.1039/C9CC00927B. [DOI] [PubMed] [Google Scholar]
  85. It is well established that chiral dirhodium paddlewheel complexes are hardly adequate for enantioselective −OH or −NH insertion and Doyle–Kirmse reactions, except for special cases or if chiral cocatalysts are added that largely account for the induction; this inability is inherent and basically rooted in the dissociation of the chiral rhodium moiety from the ylide primarily formed before the enantiodetermining [1,2]-H shift occurs. In line with this notion, the few test reactions performed with complexes 3 also afforded no or only moderate enantioselectivity; cf. the SI. For pertinent literature, see refs (17), (86), (87), and the following:Liang Y.; Zhou H.; Yu Z.-X. Why is Copper(I) Complex More Competent Than Dirhoidum(II) Complex in Catalytic Asymmetric O–H Insertion Reactions? A Computational Study of the Metal Carbenoid O–H Insertion into Water. J. Am. Chem. Soc. 2009, 131, 17783–17785. 10.1021/ja9086566. [DOI] [PubMed] [Google Scholar]
  86. Jana S.; Guo Y.; Koenigs R. M. Recent Perspectives on Rearrangement Reactions of Ylides via Carbene Transfer Reactions. Chem. - Eur. J. 2021, 27, 1270–1281. 10.1002/chem.202002556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Ren Y.-Y.; Zhu S.-F.; Zhou Q.-L. Chiral Proton-Transfer Shuttle Catalysis for Carbene Insertion Reactions. Org. Biomol. Chem. 2018, 16, 3087–3094. 10.1039/C8OB00473K. [DOI] [PubMed] [Google Scholar]
  88. Espino C. G.; Fiori K. W.; Kim M.; Du Bois J. Expanding the Scope of C–H Amination through Catalyst Design. J. Am. Chem. Soc. 2004, 126, 15378–15379. 10.1021/ja0446294. [DOI] [PubMed] [Google Scholar]
  89. These reactions were undertaken in an attempt to form racemic product as needed for accurate ee-determination.
  90. Lamb K. L.; Squitieri R. A.; Chintala S. R.; Kwong A. J.; Balmond E. I.; Soldi C.; Dmitrenko O.; Castineira Reis M.; Chung R.; Addison J. B.; Fettinger J. C.; Hein J. E.; Tantillo D. J.; Fox J. M.; Shaw J. T. Synthesis of Benzodihydrofurans by Asymmetric C–H Insertion Reactions of Donor/Donor Rhodium Carbenes. Chem. - Eur. J. 2017, 23, 11843–11855. 10.1002/chem.201701630. [DOI] [PubMed] [Google Scholar]
  91. For leading references, see refs (92−94) and the following:Buck R. T.; Doyle M. P.; Drysdale M. J.; Ferris L.; Forbes D. C.; Haigh D.; Moody C. J.; Pearson N. D.; Zhou Q.-L. Asymmetric Rhodium Carbenoid Insertion into the Si-H Bond. Tetrahedron Lett. 1996, 37, 7631–7634. 10.1016/0040-4039(96)01679-6. [DOI] [Google Scholar]
  92. Davies H. M. L.; 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. 10.1016/S0040-4039(97)00205-0. [DOI] [Google Scholar]
  93. Kitagaki S.; Kinoshita M.; Takeba M.; Anada M.; Hashimoto S. Enantioselective Si–H Insertion of Methyl Phenyldiazoacetate Catalyzed by Dirhodium(II) Carboxylates Incorporating N-Phthaloyl-(S)-Amino Acids as Chiral Bridging Ligands. Tetrahedron: Asymmetry 2000, 11, 3855–3859. 10.1016/S0957-4166(00)00384-0. [DOI] [Google Scholar]
  94. Sambasivan R.; Ball Z. T. Metallopeptides for Asymmetric Dirhodium Catalysis. J. Am. Chem. Soc. 2010, 132, 9289–9291. 10.1021/ja103747h. [DOI] [PubMed] [Google Scholar]
  95. Zhang Y.-Z.; Zhu S.-F.; Wang L.-X.; Zhou Q.-L. Copper-Catalyzed Highly Enantioselective Carbenoid Insertion into Si–H Bonds. Angew. Chem., Int. Ed. 2008, 47, 8496–8498. 10.1002/anie.200803192. [DOI] [PubMed] [Google Scholar]
  96. Yasutomi Y.; Suematsu H.; Katsuki T. Iridium(III)-Catalyzed Enantioselective Si–H Bond Insertion and Formation of an Enantioenriched Silicon Center. J. Am. Chem. Soc. 2010, 132, 4510–4511. 10.1021/ja100833h. [DOI] [PubMed] [Google Scholar]
  97. Wang J.-C.; Xu Z.-J.; Guo Z.; Deng Q.-H.; Zhou C.-Y.; Wan X.-L.; Che C.-M. Highly Enantioselective Intermolecular Carbene Insertion to C–H and Si–H Bonds Catalyzed by a Chiral Iridium(III) Complex of a D4-Symmetric Halterman Porphyrin Ligand. Chem. Commun. 2012, 48, 4299–4301. 10.1039/c2cc30441d. [DOI] [PubMed] [Google Scholar]
  98. Chen D.; Zhu D.-X.; Xu M.-H. Rhodium(I)-Catalyzed Highly Enantioselective Insertion of Carbenoid into Si–H: Efficient Access to Functional Chiral Silanes. J. Am. Chem. Soc. 2016, 138, 1498–1501. 10.1021/jacs.5b12960. [DOI] [PubMed] [Google Scholar]
  99. Caló F.; Fürstner A. A Heteroleptic Dirhodium Catalyst for Asymmetric Cyclopropanation with α-Stannyl-α-Diazoacetate. ‘Stereoretentive’ Stille Coupling with Formation of Chiral Quarternary Centers. Angew. Chem., Int. Ed. 2020, 59, 13900–13907. 10.1002/anie.202004377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Buchsteiner M.; Martinez-Rodriguez L.; Jerabek P.; Pozo I.; Patzer M.; Nöthling N.; Lehmann C.; Fürstner A. Catalytic Asymmetric Fluorination of Copper Carbene Complexes: Preparative Advances and a Mechanistic Rationale. Chem. - Eur. J. 2020, 26, 2509–2515. 10.1002/chem.202000081. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ja1c01972_si_001.pdf (14.1MB, pdf)

Articles from Journal of the American Chemical Society are provided here courtesy of American Chemical Society

RESOURCES