2020 Science 367, 1246-1251 Enantioselective Remote C-H Activation Directed by A Chiral Cation

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Corrected 13 March 2020. See full text.

RES EARCH

ORGANIC CHEMISTRY a single case functionalize at the arene ortho


position (32). Only very recently did Yu and
Enantioselective remote C–H activation directed co-workers achieve enantioselective desym-
metrization through direct arylation at the
by a chiral cation arene meta position (Fig. 1C) (33), taking ad-
vantage of an ingenious relay strategy via the
Georgi R. Genov*, James L. Douthwaite*, Antti S. K. Lahdenperä, David C. Gibson, Robert J. Phipps† ortho position, although relatively high load-
ings of the chiral norbornene mediator (CTM,
Chiral cations have been used extensively as organocatalysts, but their application to rendering transition 20 to 50 mol %) were required. C–H borylation
metal–catalyzed processes enantioselective remains rare. This is despite the success of the analogous reactions have the useful attribute that the
charge-inverted strategy in which cationic metal complexes are paired with chiral anions. We report new C–B bond can undergo numerous diverse
here a strategy to render a common bipyridine ligand anionic and pair its iridium complexes with a chiral transformations (34, 35), but so far, enantio-
cation derived from quinine. We have applied these ion-paired complexes to long-range asymmetric control in arene borylation has been realized
induction in the desymmetrization of the geminal diaryl motif, located on a carbon or phosphorus center, only in two recent reports, from Shi, Hartwig,
by enantioselective C–H borylation. In principle, numerous common classes of ligand could likewise and co-workers (30) and Xu, Ke, and co-workers
be amenable to this approach. (31). In both cases, the chiral information is
covalently incorporated into the ligand scaf-

I
fold in the conventional manner and a directing
on-pairing has been put to extensive despite the obvious potential presented by sev- group guides borylation to the ortho position.
use as a key design feature in the field eral privileged classes of chiral cation. Given By contrast, the creation of chirality over long

Downloaded from http://science.sciencemag.org/ on March 17, 2020


of asymmetric catalysis (1). In the 1980s, the success of these motifs as chiral controllers ranges, where the enantiotopic site is far from
pioneering studies on enantioselective in asymmetric organocatalysis (vide supra), a the new stereocenter, is an outstanding chal-
phase-transfer catalysis paired a chiral general strategy to integrate them with tran- lenge in which catalyst designs that incorpo-
cation with a reactive anionic intermediate sition metal catalysis would likely have broad rate noncovalent interactions offer numerous
in the enantiodetermining transition state impact in the field of asymmetric catalysis. opportunities (36–38).
(2), with cinchona alkaloid-derived cations In an important advance, which compel- We recently developed anionic bipyridine
dominating as effective and readily accessi- lingly demonstrates this potential, Ooi and ligands that bear a remote sulfonate group to
ble scaffolds (3). The numerous subsequent co-workers incorporated a chiral cation cova- impart control of regioselectivity in iridium-
developments in this area have had enormous lently into the structure of a phosphine ligand, catalyzed C–H borylation via noncovalent in-
impact in the field of asymmetric organo- resulting in highly stereocontrolled formation teractions with the substrate (39–41). Throughout
catalysis, encapsulating such important trans- of contiguous all-carbon quaternary stereocen- these studies, a single ligand scaffold consist-
formations as Michael and aldol additions, as ters under palladium catalysis (Fig. 1A, right) ently gave the optimal regiocontrol. In one
well as Mannich, fluorination, alkylation, and (22, 23). At the outset of this project, we en- particular study, we attributed the high regio-
oxidative cyclization reactions, to name but a visioned a potentially more generally appli- selectivity for borylation at the arene meta
few (4–7) (Fig. 1A, left). Over the past decade, cable approach whereby an anionic handle is position to the existence of a hydrogen bond
the inverse strategy of using a chiral anion to incorporated into a common ligand scaffold, between the substrate and the sulfonate group
associate with a cationic reaction intermediate providing the key point of interaction with of the ligand in the regiodetermining transition
has also proven extremely successful (1, 8, 9). the chiral cation (Fig. 1B). Judicious place- state for C–H activation (Fig. 1D) (40). We hy-
This latter strategy has been effective not only ment of this anionic group would be crucial pothesized that exchange of the achiral tetra-
in an organocatalytic context (10, 11) but also to success—not close enough to the metal cen- butylammonium counterion of the ligand for
in powerful combination with transition metal ter to disrupt reactivity but not so far that the a chiral cation might allow enantioselective,
catalysts (12–14), cleverly capitalizing on the chiral environment imparted by the cation desymmetrizing C–H activation in a prochiral
relatively common occurrence of cationic tran- would be ineffective. Various chiral cations substrate (as in Fig. 1E). Herein, we demon-
sition metal complexes in catalytic cycles. By could be introduced in the final step by sim- strate that, using this approach, remote, enan-
contrast, it is far rarer to encounter anionic ple ion exchange, allowing for rapid catalyst tioselective C–H borylation can be achieved
transition metal complexes as key intermedi- optimization. In pioneering work, Ooi and for formation of chiral-at-carbon and chiral-
ates. As such, the charge-inverted approach of co-workers previously demonstrated the pro- at-phosphorus compounds, showcasing the
pairing a chiral cation with an anionic transition ductive combination of cationic ligands with thus far unexplored approach of combining
metal catalyst has only been demonstrated in a chiral anions, as demonstrated effectively in a chiral cation with an anionic ligand for a
handful of pioneering cases, notably asymmetric enantioselective allylic alkylation (24, 25). We reactive transition metal.
oxidation reactions involving anionic diphos- envisaged that, in principle, a wide variety of We commenced our studies with symmetri-
phatobisperoxotungstate (15) and peroxomo- privileged ligand scaffolds for transition metal cal benzhydrylamide 2a (Fig. 2A). Numerous
lybdate (16) complexes as catalysts (Fig. 1A, catalysis could be rendered anionic, creating ion-paired ligands L·1, possessing a variety
center) (17–21). Owing to this scarcity of anionic exciting opportunities to explore the use of chi- of chiral cations 1a to 1i, could be readily ob-
metal complexes in the most commonly used ral cations as chiral controllers in a wealth of tained through counterion exchange. The
processes, the broader potential of uniting powerful transition metal–catalyzed reactions. chiral cations were all derived from dihydro-
chiral cations with the versatile reactivity of In seeking a rigorous and relevant test of the quinine (DHQ) with varying N-benzyl substitu-
transition metals has remained underexplored, above-described approach, we targeted a trans- tion. At room temperature in tetrahydrofuran
formation that lies at the cutting edge of what (THF) as solvent, low but encouraging levels
is currently possible in enantioselective cataly- of enantioselectivity were obtained with 3,5-
sis. Although enantioselective, desymmetrizing dimethoxy benzyl and 3,5-di-tert-butyl groups
Department of Chemistry, University of Cambridge, Lensfield C–H activation of arenes has been extensively [L·1a and L·1b, 31 and 30% enantiomeric
Road, Cambridge, CB2 1EW, UK.
*These authors contributed equally to this work. explored with palladium (26, 27), rhodium excess (ee)]. We next investigated placing sub-
†Corresponding author. Email: rjp71@cam.ac.uk (28, 29), and iridium (30, 31) catalysis, all but stituted aromatic rings at the 3- and 5-positions

Genov et al., Science 367, 1246–1251 (2020) 13 March 2020 1 of 6


Corrected 13 March 2020. See full text.
RES EARCH | R E P O R T

Fig. 1. Strategy for


incorporating chiral
cations with transition
metals. (A) Applications
of chiral cations in
asymmetric catalysis.
Me, methyl; Ph, phenyl;
R, substituent. (B) General
strategy for integration of
transition metal (TM)
catalysis with chiral
cations. M, metal.
(C) State-of-the-art
methodology for enantiose-
lective remote C–H
activation of arenes.
Boc, butyloxycarbonyl;
Ac, acetyl; BNDHP,
1,1′-binaphthyl-2,2′-diyl
hydrogenphosphate.

Downloaded from http://science.sciencemag.org/ on March 17, 2020


(D) Our prior work
controlling regioselectivity
in arene borylation.
Bu, butyl. (E) Summary
of this work.

of the quaternizing N-benzyl group. Encourag- point, we investigated aryl groups in these with H2O2 (see inset in Fig. 2A). This deriva-
ingly, L·1c (4-CF3C6H4) gave increased enantio- positions to extend the reach even further, but tization aided separation from any remain-
selectivity (39% ee) and L·1d (3,4,5-F3C6H2) both of these proved detrimental (L·1h and ing starting material or difunctionalized material.
resulted in a further improvement (52% ee). L·1i). Thus, we shifted our attention to other The undesired borylation of monoborylated 3a′
Focusing attention on the meta positions of reaction parameters with L·1g. A solvent eval- to give a symmetrical diborylated by-product did
the outer arenes of the teraryl system, we then uation identified cyclopentyl methyl ether occur to varying degrees in the reactions, being
evaluated a series of substituents (L·1e to (CPME) as being optimal, in that the reaction unavoidable at higher conversions. We thus
L·1i). Trifluoromethyl (L·1e) and methoxy temperature could be reduced to −10°C while carried out careful experiments to establish
(L·1f) substitution again gave increases (both high reactivity was maintained, resulting in iso- whether kinetic resolution may be occurring
60% ee), but the biggest gain came from the lation of 3a in 72% yield and with 96% ee, fol- in such instances, resulting in possible en-
tert-butyl substituted L·1g (73% ee). At this lowing oxidation to the corresponding phenol hancement of the observed ee of 3a′ at the

Genov et al., Science 367, 1246–1251 (2020) 13 March 2020 2 of 6


Corrected 13 March 2020. See full text.
RES EARCH | R E P O R T

Downloaded from http://science.sciencemag.org/ on March 17, 2020

Fig. 2. Enantioselective desymmetrizing C–H borylation of benzhydryla- enantioselectivity. Yield values refer to isolated yields. Regioisomeric ratios
mides. (A) Reaction optimization. COD, 1,5-cyclooctadiene; rt, room were determined from the crude 1H–nuclear magnetic resonance (NMR)
temperature; tBu, tert-butyl. (B) Scope of enantioselective borylation spectrum before isolation. Enantiomeric excesses determined by chiral
using L·1g in substrates bearing no regioselectivity challenge. Et, ethyl. high-performance liquid chromatography (HPLC) or supercritical fluid
(C) Examples in which the catalyst is controlling regioselectivity and chromatography (SFC) analysis.

Genov et al., Science 367, 1246–1251 (2020) 13 March 2020 3 of 6


Corrected 13 March 2020. See full text.
RES EARCH | R E P O R T

end of the reaction. Evaluating ee of 3a′ at control both of these important selectivity enantioselectivity (85% ee). In contrast to this,
various levels of conversion as well as sub- factors for a substrate that possessed ortho- the control borylation with standard borylation
mitting racemic 3a′ to the enantioselective substituted aromatic rings (Fig. 2C) (40). We ligand dtbpy resulted in a 1.6:1 ratio of regio-
borylation conditions with L·1g showed that were concerned that the introduction of ortho isomers (fig. S5). An ortho-bromo substrate
there is no appreciable kinetic resolution oc- substituents may substantially change the pre- performed similarly (3l), as did an ortho-CF3
curring (figs. S1 and S2). Finally, we evaluated ferred substrate conformation, potentially af- (3m) and ortho-OCF3 (3o). We also examined
a ligand paired with a Maruoka-type chiral fecting crucial interactions with the chiral a meta-fluoro substrate, which presents regio-
cation which gave racemic product, a variant of cation. Also, it was possible that the complex selectivity challenges using standard ligands
L·1g in which the quinine hydroxyl group is chiral cation might disrupt the regioselectivity owing to the small size of the fluorine atom
methylated, which gave a reduced ee of 72%, that we had previously observed when using (42), but with L·1g, high regioselectivity was
and a variant of L·1g in which the stereo- tetrabutylammonium as the cation. However, observed (3n). In addition, we carried out
chemistry of the quinine hydroxyl group is we were delighted to find that an ortho-chloro preliminary experiments with nonsymmetri-
inverted, which gave only 11% ee (see sup- substrate gave the meta-borylated product 3k cal substrates to assess the viability of using
plementary materials and table S1 for full op- with excellent regioselectivity [10:1 regioiso- the reaction in kinetic resolution mode. These
timization details). A survey of N-protecting meric ratio (rr)] and only a small reduction in showed that it is indeed viable, although
groups demonstrated that trifluoroacetyl is
optimal, although acetyl also performed well
(fig. S3).
We proceeded to examine the scope of the
reaction in terms of versatile substituents on

Downloaded from http://science.sciencemag.org/ on March 17, 2020


the substrate aryl rings (Fig. 2B). Postreaction
derivatization of the introduced boronic acid
pinacol ester (BPin) group facilitated purifi-
cation, and we used oxidation with hydrogen
peroxide to give the corresponding phenols.
We were pleased to find that halide substi-
tution was very well tolerated in the enantio-
selective borylation. Chloro-substituted (3b),
bromo-substituted (3d), and iodo-substituted
(3e) arenes all delivered excellent levels of en-
antioselectivity, the latter being of particular
note because it would likely be incompatible
with palladium catalysis and is a testament to
the mild conditions and functional group to-
lerance of iridium-catalyzed borylation. The
N-trifluoroacetyl group could be replaced by
acetyl with little drop in ee, as demonstrated
on substrate 3c. The absolute stereochemistry
of compound 3e was determined by x-ray crys-
tallographic analysis, and all other compounds
were assigned by analogy with this. Further
variation of substituents revealed that trifluor-
omethoxy (3f), ester (3g), and nitrile (3h) were
all well accommodated at the 3-position of the
substrates. We also examined vicinally dichlo-
rinated (3i) and difluorinated (3j) substrates,
which both worked effectively. Substrates bear-
ing electron-donating substituents exhibited
lower reactivity under our conditions—3-
methoxy gave no conversion and 3-methyl gave
<5% conversion, likely owing to the reaction
temperatures being lower than those typically
used in C–H borylation. Performing the reac-
tions at room temperature gave conversion, but
with moderate enantioselectivity (fig. S4). The
substrates examined so far have all presented
no regioselectivity challenge, owing to the well-
established preference for C–H borylation at
the least hindered arene position (42). Given
that the sulfonated bipyridine ligand scaffold
was originally designed for the purpose of
controlling regioselectivity in substrates that
would typically be nonselective, we were keen Fig. 3. Substrate scope of the enantioselective C–H borylation of diaryl phosphinamides. Yield values
to evaluate whether L·1g would be able to refer to isolated yields.

Genov et al., Science 367, 1246–1251 (2020) 13 March 2020 4 of 6


Corrected 13 March 2020. See full text.
RES EARCH | R E P O R T

further investigations and optimization are nitrogen, was borylated to give 3p with 90% both regioselectivity and enantioselectivity
likely required to enable this to be a general ee using ligand L·1g, which had been op- in this substrate class, we tested an ortho-
procedure (fig. S6). timal for the benzhydrylamide substrate class substituted symmetrical phosphinamide but
At this stage, we envisaged that a com- (Fig. 3). X-ray crystallographic analysis of 3p found that both outcomes were poor (fig. S8).
pelling demonstration of the potential of this showed that this product had analogous ab- We speculate that this may arise owing to the
approach would be to successfully apply it to solute stereochemistry to that obtained in the ortho-substituted aromatic ring and bulky
a different class of compound entirely. For amide series, relative to the position of the NH nature of the quaternary phosphorus center,
this purpose, we identified symmetrical diaryl- hydrogen-bond donor. Experiments stopped at relative to the benzhydrylamide, having a
phosphinamides, which contain a prochiral, various conversions demonstrated that sec- conformational impact on the substrate that
configurationally stable phosphorus atom at ondary kinetic resolution to form diborylated adversely affects crucial substrate-ligand
the heart of the compound. We reasoned that product is not contributing to the observed interactions.
such substrates would test our chiral cation– high enantioselectivity (fig. S7). N-substitution For both classes of compounds demon-
directed C–H borylation strategy in tackling was found not to be limited to aromatic strated, the C–H borylation products typi-
an additional prominent challenge to synthetic moieties, as demonstrated by N-tert-butyl sub- cally possess three versatile functional groups
chemists—that of how to synthesize P-chiral stituted 3q (95% ee). As in the amide sub- on the aromatic rings for further elaboration
compounds in a catalytic asymmetric manner strate class, a variety of useful functional groups into complex scaffolds, at the heart of which
(43). Although there are several recently re- were tolerated on the aromatic ring, encom- lies the newly formed stereocenter. By virtue
ported methods for enantioselective desym- passing bromide (3r), ester (3s), iodide (3t), of the desymmetrization strategy used, two
metrizing C–H activation of phosphinamides trifluoromethoxy (3u), trifluoromethyl (3v), of these functional groups must necessarily
using chiral Pd and Rh complexes, both result and nitrile (3w). In some cases, yields are be identical, and we sought to demonstrate
in ortho-functionalized products (44, 45).

Downloaded from http://science.sciencemag.org/ on March 17, 2020


modest owing to poor substrate solubility that site selectivity between these in the pro-
Given the broad utility of P-chiral compounds under the reaction conditions (as in 3s). There duct should be possible in many instances by
in catalysis as well as increasingly in medicinal are numerous established avenues for the ma- electronic differentiation arising from intro-
chemistry, we envisaged that remote desym- nipulation of the phosphinamide functional duction of the new substituent. In the first
metrization would be of substantial practical group in a stereospecific manner, such as to example, we carried out borylation and oxi-
utility (46). We were pleased to observe that a tertiary phosphine oxides, which have been dation of dichloride 2b to give the phenol 3b
symmetrical phosphinamide, bearing a para- amply demonstrated elsewhere (44). To test with good yield and high enantioselectivity
methoxy phenyl group on the phosphinamide whether the catalyst may be able to influence (Fig. 4A, upper scheme). By carrying out

Fig. 4. Product elaboration and further experiments. (A) Use of arene electronics to control site-selective derivatization of reaction products. HATU,
hexafluorophosphate azabenzotriazole tetramethyl uronium. (B) Use of a pseudoenantiomeric chiral cation to form (R)-3a. (C) Control experiments to probe
ligand-substrate interactions. % conv., % conversion. (D) Control experiments to probe ligand-cation interactions. Yield values refer to isolated yields.

Genov et al., Science 367, 1246–1251 (2020) 13 March 2020 5 of 6


Corrected 13 March 2020. See full text.
RES EARCH | R E P O R T

Suzuki-Miyaura coupling on 3b in the presence scaffold. In principle, numerous widely used 33. H. Shi, A. N. Herron, Y. Shao, Q. Shao, J.-Q. Yu, Nature 558,
of one equivalent of tetrabutylammonium transition metal–catalyzed reactions could be 581–585 (2018).
34. J.-Y. Cho, M. K. Tse, D. Holmes, R. E. Maleczka Jr.,
hydroxide, we were able to achieve >20:1 site amenable to this approach, as evidenced by M. R. Smith III, Science 295, 305–308 (2002).
selectivity for cross-coupling on the nonphe- the numerous common ligand classes that 35. T. Ishiyama et al., J. Am. Chem. Soc. 124, 390–391
nolic aromatic ring. We anticipate that this have been sulfonated for the purpose of en- (2002).
36. B. Kim et al., J. Am. Chem. Soc. 138, 7939–7945 (2016).
is a result of the highly electron-rich nature gendering water solubility (49). We anticipate 37. A. J. Metrano, S. J. Miller, Acc. Chem. Res. 52, 199–215
of the in situ–generated phenolate disfavor- that wider incorporation of chiral cations into (2019).
ing oxidative addition to the C–Cl bond on the mainstream transition metal catalysis could 38. S. Y. Hsieh, Y. Tang, S. Crotti, E. A. Stone, S. J. Miller, J. Am.
Chem. Soc. 141, 18624–18629 (2019).
same ring. In the second case, we carried out have broad implications in asymmetric or- 39. H. J. Davis, M. T. Mihai, R. J. Phipps, J. Am. Chem. Soc. 138,
borylation of diester 2g followed by cyana- ganic synthesis. 12759–12762 (2016).
tion to obtain 6 (Fig. 4A, lower scheme) (47). 40. H. J. Davis, G. R. Genov, R. J. Phipps, Angew. Chem. Int. Ed. 56,
13351–13355 (2017).
Careful treatment of 6 with NaOH selectively RE FERENCES AND NOTES 41. M. T. Mihai, H. J. Davis, G. R. Genov, R. J. Phipps, ACS Catal. 8,
hydrolyzed the ester on the same ring as the 1. K. Brak, E. N. Jacobsen, Angew. Chem. Int. Ed. 52, 534–561 3764–3769 (2018).
nitrile owing to electronic factors that can be (2013). 42. J. F. Hartwig, Chem. Soc. Rev. 40, 1992–2002 (2011).
readily rationalized and predicted using sub- 2. U. H. Dolling, P. Davis, E. J. J. Grabowski, J. Am. Chem. Soc. 43. M. Dutartre, J. Bayardon, S. Jugé, Chem. Soc. Rev. 45,
106, 446–447 (1984). 5771–5794 (2016).
stituent Hammett parameters (48), giving 7 in 3. H.-G. Park, B.-S. Jeong, in Cinchona Alkaloids in Synthesis and 44. Z.-J. Du et al., J. Am. Chem. Soc. 137, 632–635 (2015).
>20:1 rr after amide coupling. Catalysis, C. E. Song, Ed. (Wiley, 2009), pp. 131–169. 45. Y. Sun, N. Cramer, Angew. Chem. Int. Ed. 56, 364–367
To emphasize the practicality of our process, 4. T. Ooi, K. Maruoka, Angew. Chem. Int. Ed. 46, 4222–4266 (2007). (2017).
5. S. Shirakawa, K. Maruoka, Angew. Chem. Int. Ed. 52, 46. D. A. DiRocco et al., Science 356, 426–430 (2017).
we demonstrated borylation using ligand L·1j, 4312–4348 (2013). 47. C. W. Liskey, X. Liao, J. F. Hartwig, J. Am. Chem. Soc. 132,
possessing a diastereomeric chiral cation de- 6. M. Uyanik, H. Okamoto, T. Yasui, K. Ishihara, Science 328, 11389–11391 (2010).

Downloaded from http://science.sciencemag.org/ on March 17, 2020


rived from quinidine, the pseudoenantiomer 1376–1379 (2010). 48. C. Hansch, A. Leo, R. W. Taft, Chem. Rev. 91, 165–195
7. M. T. Oliveira, J.-W. Lee, ChemCatChem 9, 377–384 (2017). (1991).
of quinine. This proceeded smoothly, giving 8. R. J. Phipps, G. L. Hamilton, F. D. Toste, Nat. Chem. 4, 49. K. H. Shaughnessy, Chem. Rev. 109, 643–710 (2009).
(R)-3a with 90% ee (Fig. 4B). Next, we per- 603–614 (2012).
formed experiments to probe the hydrogen- 9. M. Mahlau, B. List, Angew. Chem. Int. Ed. 52, 518–533 AC KNOWLED GME NTS
(2013). We are grateful to A. Bond (University of Cambridge) for solving
bonding interaction of substrate with ligand. 10. S. Mayer, B. List, Angew. Chem. Int. Ed. 45, 4193–4195 (2006). and refining the x-ray crystal structures, the EPSRC UK National
The N-methylated variant (2x) of successful 11. G. L. Hamilton, T. Kanai, F. D. Toste, J. Am. Chem. Soc. 130, Mass Spectrometry Facility at Swansea University, M. Gaunt for
substrate 2d underwent no borylation under 14984–14986 (2008). useful discussion and use of equipment, and C. Hunter for useful
the optimized conditions at −10°C, and the 12. G. L. Hamilton, E. J. Kang, M. Mba, F. D. Toste, Science 317, discussion. We also thank A. Turner (AstraZeneca) and P. Seden
496–499 (2007). (Syngenta) for useful discussion. Funding: We are grateful to
temperature had to be raised to 10°C to obtain 13. S. Mukherjee, B. List, J. Am. Chem. Soc. 129, 11336–11337 The Royal Society for a University Research Fellowship (R.J.P.,
product, which was found to have only 8% ee (2007). UF130004), the EPSRC (EP/N005422/1), the European Research
(Fig. 4C). This outcome highlights the impor- 14. S. Liao, B. List, Angew. Chem. Int. Ed. 49, 628–631 Council under the Horizon 2020 Program (Starting Grant no.
(2010). 757381), the EPSRC and AstraZeneca for a CASE studentship
tance of the hydrogen-bond donor in the 15. X. Ye et al., Angew. Chem. Int. Ed. 55, 7101–7105 (2016). (J.L.D.), The Emil Aaltonen Foundation for a fellowship (A.S.K.L.),
substrate for both reactivity and selectivity, in 16. L. Zong et al., Nat. Commun. 7, 13455 (2016). and the EPSR, and Syngenta for a CASE studentship (D.C.G.).
line with our initial hypothesis (Fig. 1E). We 17. R. C. D. Brown, J. F. Keily, Angew. Chem. Int. Ed. 40, Author contributions: G.R.G. developed the catalysts and
4496–4498 (2001). reactions and performed and analyzed experiments. J.L.D.
also performed an experiment in which ion- 18. C. Wang, L. Zong, C.-H. Tan, J. Am. Chem. Soc. 137, developed the reactions and performed and analyzed experiments.
paired ligand L·1g was replaced with neutral 10677–10682 (2015). A.S.K.L and D.C.G. performed and analyzed experiments. R.J.P.
5,5′-dimethylbipyridine (8) together with 19. L. Zong, C.-H. Tan, Acc. Chem. Res. 50, 842–856 (2017). conceived and supervised the project and wrote the manuscript,
20. S. Paria, H.-J. Lee, K. Maruoka, ACS Catal. 9, 2395–2399 with input from all authors. Competing interests: The authors
the optimal chiral cation as its bromide salt (2019). declare no competing interests. Data and materials availability:
(Br·1g). The product was racemic, demon- 21. K. Ohmatsu, T. Ooi, Top. Curr. Chem. 377, 31 (2019). The supplementary materials contain additional spectral and
strating the requirement for ligand and chiral 22. K. Ohmatsu, N. Imagawa, T. Ooi, Nat. Chem. 6, 47–51 chromatographic data. Crystallographic data are available free of
(2014). charge from the Cambridge Crystallographic Data Centre under
cation to be associated to achieve enantioin- 23. K. Ohmatsu, S. Kawai, N. Imagawa, T. Ooi, ACS Catal. 4, reference numbers CCDC 1959892 and
duction. We also ran a reaction in which ligand 4304–4306 (2014). CCDC 1960549.
L·NBu4, bearing achiral tetrabutylammonium 24. K. Ohmatsu, M. Ito, T. Kunieda, T. Ooi, Nat. Chem. 4, 473–477
(2012).
as the cation, was used in conjunction with 25. K. Ohmatsu, M. Ito, T. Kunieda, T. Ooi, J. Am. Chem. Soc. 135,
SUPPLEMENTARY MATERIALS
Br1g. In this case, 58% ee was obtained in 590–593 (2013). science.sciencemag.org/content/367/6483/1246/suppl/DC1
the product, consistent with some degree of 26. X.-F. Cheng et al., J. Am. Chem. Soc. 135, 1236–1239 (2013). Materials and Methods
27. L. Chu, K.-J. Xiao, J.-Q. Yu, Science 346, 451–455 (2014). Figs. S1 to S8
counterion exchange occurring between the Tables S1 and S2
28. B. Ye, N. Cramer, Science 338, 504–506 (2012).
two, leading to moderate enantioinduction. 29. B. Ye, N. Cramer, Acc. Chem. Res. 48, 1308–1318 (2015). HPLC and SFC Traces
We have demonstrated a strategy for pairing 30. B. Su, T.-G. Zhou, P.-L. Xu, Z.-J. Shi, J. F. Hartwig, NMR Spectra
Angew. Chem. Int. Ed. 56, 7205–7208 (2017). References (50–63)
privileged chiral cations with an iridium-
31. X. Zou et al., J. Am. Chem. Soc. 141, 5334–5342 (2019).
bipyridine complex, enabled by incorporation 32. C. G. Newton, S.-G. Wang, C. C. Oliveira, N. Cramer, Chem. Rev. 8 November 2019; accepted 19 February 2020
of an anionic sulfonate group into the ligand 117, 8908–8976 (2017). 10.1126/science.aba1120

Genov et al., Science 367, 1246–1251 (2020) 13 March 2020 6 of 6


Enantioselective remote C−H activation directed by a chiral cation
Georgi R. Genov, James L. Douthwaite, Antti S. K. Lahdenperä, David C. Gibson and Robert J. Phipps

Science 367 (6483), 1246-1251.


DOI: 10.1126/science.aba1120originally published online March 12, 2020

Asymmetry on the plus side


Numerous positively charged metal catalysts have been paired with chiral negative ions to select for just one of
two mirror-image products. Genov et al. now report a potentially general strategy to invert the charges in this paradigm.
Because intrinsically negative metal catalysts are comparatively rare, the authors appended a sulfonate group to the
common bipyridyl ligand. Iridium complexes of this ligand paired with chiral positive ions could borylate just one of two

Downloaded from http://science.sciencemag.org/ on March 17, 2020


aryl rings appended to carbon or phosphorus centers with high enantioselectivity.
Science, this issue p. 1246

ARTICLE TOOLS http://science.sciencemag.org/content/367/6483/1246

SUPPLEMENTARY http://science.sciencemag.org/content/suppl/2020/03/11/367.6483.1246.DC1
MATERIALS

REFERENCES This article cites 61 articles, 6 of which you can access for free
http://science.sciencemag.org/content/367/6483/1246#BIBL

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Use of this article is subject to the Terms of Service

Science (print ISSN 0036-8075; online ISSN 1095-9203) is published by the American Association for the Advancement of
Science, 1200 New York Avenue NW, Washington, DC 20005. The title Science is a registered trademark of AAAS.
Copyright © 2020 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of
Science. No claim to original U.S. Government Works

You might also like