You are on page 1of 5

RE S EAR CH | R E P O R T S

resonance (32). In addition, the protocol devel- DUAL CATALYSIS


oped here is general and will affect experiments
with ion traps and other platforms as system
sizes increase, both in full calibrations of the
coupling matrix and in the ability to observe a
Single-electron transmetalation in
single quantity that serves as a proxy for the
entire Hamiltonian. organoboron cross-coupling by
RE FE RENCES AND N OT ES
1. J. I. Cirac, P. Zoller, Nat. Phys. 8, 264–266 (2012).
photoredox/nickel dual catalysis
2. B. A. Cipra, SIAM News 33 (6), 654 (2000); www.siam.org/
pdf/news/654.pdf. John C. Tellis,* David N. Primer,* Gary A. Molander†
3. E. Schneidman, M. J. Berry 2nd, R. Segev, W. Bialek, Nature
440, 1007–1012 (2006).
The routine application of Csp3-hybridized nucleophiles in cross-coupling reactions remains
4. S. Liu, L. Ying, S. Shakkottai, in 48th Annual Allerton
Conference on Communication, Control, and Computing (IEEE, an unsolved challenge in organic chemistry. The sluggish transmetalation rates observed
New York, 2010), pp. 570–576. for the preferred organoboron reagents in such transformations are a consequence of
5. H. T. Diep, Ed., Frustrated Spin Systems (World Scientific, the two-electron mechanism underlying the standard catalytic approach. We describe a
Singapore, 2005).
mechanistically distinct single-electron transfer-based strategy for the activation of
6. R. Feynman, Int. J. Theor. Phys. 21, 467–488 (1982).
7. S. Lloyd, Science 273, 1073–1078 (1996). organoboron reagents toward transmetalation that exhibits complementary reactivity
8. Nature Physics, Insight Issue: “Quantum Simulation” 8, 264 patterns. Application of an iridium photoredox catalyst in tandem with a nickel catalyst
(2012). effects the cross-coupling of potassium alkoxyalkyl- and benzyltrifluoroborates with an
9. A. Aspuru-Guzik, P. Walther, Nat. Phys. 8, 285–291
array of aryl bromides under exceptionally mild conditions (visible light, ambient temperature,

Downloaded from http://science.sciencemag.org/ on March 14, 2018


(2012).
10. R. Blatt, C. F. Roos, Nat. Phys. 8, 277–284 (2012). no strong base). The transformation has been extended to the asymmetric and
11. I. Bloch, J. Dalibard, S. Nascimbene, Nat. Phys. 8, 267–276 stereoconvergent cross-coupling of a secondary benzyltrifluoroborate.

T
(2012).
12. P. Hauke, F. M. Cucchietti, L. Tagliacozzo, I. Deutsch, he immense impact of transition metal– are used, thereby limiting functional group tol-
M. Lewenstein, Rep. Prog. Phys. 75, 082401 (2012). catalyzed cross-coupling has been well rec- erance and augmenting deleterious side reac-
13. D. Porras, J. I. Cirac, Phys. Rev. Lett. 92, 207901
(2004).
ognized, with the Suzuki-Miyaura reaction tions (5). Stoichiometric Ag and Cu salts have
14. K. Kim et al., Phys. Rev. Lett. 103, 120502 (2009). in particular emerging as a preferred meth- been shown to improve transmetalation efficien-
15. K. Kim et al., Nature 465, 590–593 (2010). od for the construction of C-C bonds in both cy in some systems (6–8), although the mech-
16. R. Islam et al., Nat. Commun. 2, 377 (2011). industrial and academic settings (1). Tradition- anism by which the acceleration is achieved is
17. J. W. Britton et al., Nature 484, 489–492 (2012).
18. R. Islam et al., Science 340, 583–587 (2013).
ally, cross-coupling reactions employ a three- unclear (9), thus limiting their widespread ap-
19. P. Richerme et al., Phys. Rev. A 88, 012334 (2013). step catalytic cycle (Fig. 1): (i) oxidative addition plication. Often, the only viable alternative to
20. Materials and methods are available as supplementary of an organic halide at Pd0, (ii) transmetalation overcome a slow transmetalation is to abandon
material on Science Online. of an organometallic nucleophile to an organo- the readily available boronic acids and make use
21. J. Schachenmayer, B. P. Lanyon, C. F. Roos, A. J. Daley,
Phys. Rev. X 3, 031015 (2013).
palladium(II) electrophile, and (iii) reductive elim- of more reactive organometallic reagents. Thus,
22. P. Hauke, L. Tagliacozzo, Phys. Rev. Lett. 111, 207202 ination from a diorganopalladium(II) species, alkylboranes, alkylzincs, or the corresponding
(2013). releasing the coupled product and regenerating Grignard reagents—all of which lack functional
23. M. van den Worm, B. C. Sawyer, J. J. Bollinger, M. Kastner, the Pd0 catalyst (1, 2). Although these methods group tolerance and are unstable to air—are
New J. Phys. 15, 083007 (2013).
24. Z.-X. Gong, L.-M. Duan, New J. Phys. 15, 113051 (2013).
are highly effective for Csp2-Csp2 coupling, exten- often used for alkyl cross-coupling (1).
25. M. Knap et al., Phys. Rev. Lett. 111, 147205 (2013). sion to Csp3 centers has proven challenging be- The challenge of alkylboron transmetalation
26. A. Khromova et al., Phys. Rev. Lett. 108, 220502 cause of lower rates of oxidative addition and was recognized to arise directly from mecha-
(2012). transmetalation, as well as the propensity of nistic limitations inherent in the two-electron
27. P. Jurcevic et al., Nature 511, 202–205 (2014).
28. S. Olmschenk et al., Phys. Rev. A 76, 052314 (2007).
the alkylmetallic intermediates to undergo facile nature of the conventional process, wherein re-
29. G. Tóth, C. Knapp, O. Gühne, H. J. Briegel, Phys. Rev. Lett. 99, b-hydride elimination (2). Recent advances in activity is inversely proportional to heterolytic
250405 (2007). ligand technology and the use of alternative C-B bond strength, thus predisposing Csp3 nu-
30. E. E. Edwards et al., Phys. Rev. B 82, 060412 (2010). metals, such as nickel, have greatly expanded cleophiles for failure in cross-coupling reactions
31. B. Yoshimura, W. C. Campbell, J. K. Freericks, http://arxiv.org/
abs/1402.7357 (2014).
the scope of the electrophilic component, extend- (10, 11). Rather than attempting to override the
32. R. R. Ernst, G. Bodehausen, A. Wokaun, Principles of Nuclear ing even to sterically hindered and unactivated inherent biases of the conventional transmetala-
Magnetic Resonance in One and Two Dimensions (Clarendon alkyl substrates, and have largely succeeded in tion pathway, we anticipated that development
Press, Oxford, 1987). retarding problematic b-hydride elimination (3). of an activation mode based on single-electron
ACKN OW LEDG MEN TS
Despite the progress achieved in advancement transfer (SET) chemistry would constitute a more
We thank J. Freericks, B. Yoshimura, E. Edwards, S. Will,
of the other fundamental steps, transmetalation efficient strategy for engaging this class of re-
Z.-X. Gong, M. Foss-Feig, and A. Gorshkov for helpful discussions. has remained largely unchanged since the in- agents (Fig. 1). Trends in homolytic C-B bond
This work is supported by the U.S. Army Research Office ception of cross-coupling chemistry. As such, strength (12) dictate that such a reaction mani-
(ARO) award no. W911NF0710576 with funds from the Defense cross-couplings conducted under the traditional fold would exhibit reactivity trends complemen-
Advanced Research Projects Agency Optical Lattice Emulator
Program, ARO award no. W911NF0410234 with funds from the
mechanistic manifold typically result in trans- tary to that of a traditional cross-coupling, with
Intelligence Advanced Research Projects Activity Multi-Qubit metalations that are rate-limiting (4). Csp3-hybridized nucleophiles now ideally primed
Coherent Operations Program, and the NSF Physics Frontier Center To date, strategies aimed at accelerating the for successful implementation.
at the Joint Quantum Institute. rate of transmetalation of Csp3-hybridized boronic The first challenge associated with the realiza-
acid reagents have been largely rudimentary. In tion of this ideal is the oxidative profile of radical
SUPPLEMENTARY MATERIALS most cases, excess base and high temperature capture at a transition metal center (R· + Mn →
www.sciencemag.org/content/345/6195/430/suppl/DC1 R-Mn+1), which necessitates a subsequent reduc-
Materials and Methods
Fig. S1
tion to maintain the redox neutrality of a tra-
Roy and Diana Vagelos Laboratories, Department of ditional transmetalation. Here, application of
Table S1
Chemistry, University of Pennsylvania, Philadelphia, PA
References (33–35)
19104, USA.
visible-light photoredox catalysis (13, 14), was
28 January 2014; accepted 3 June 2014 *These authors contributed equally to this work. †Corresponding envisioned to satisfy the requirements of this
10.1126/science.1251422 author. E-mail: gmolandr@sas.upenn.edu unique series of SETs. Encouragement in this

SCIENCE sciencemag.org 25 JULY 2014 • VOL 345 ISSUE 6195 433


R ES E A RC H | R E PO R TS

regard was found in the work of Sanford and groups (15 and 17). Substrates possessing an 34, indole 35, and quinoline 37, proved to be
Glorius, who had demonstrated that coopera- ortho substituent were well tolerated, as evi- competent partners. Although five-membered
tive catalysis between transition metals and denced by isolation of product 13 in 82% yield. heterocyclic bromides generally exhibited poor
photoredox catalysts is indeed possible (15–17). The reaction also exhibited increased efficiency reactivity, electron-deficient thienyl bromides
Potassium organotrifluoroborates were iden- on a larger scale, as diarylmethane 12 was isolated were coupled in moderate yields, leading to 38
tified as promising partners in this new class of in 97% yield on a 5.5-mmol scale with reduced and 39.
cross-couplings, as previous reports have docu- catalyst loading [1 mol % 4 and 1.5 mol % of Ni Several practical and more sustainable fea-
mented their ability to function as carbon radical (COD)2 and ligand]. tures derive from this approach to cross-coupling.
sources upon single-electron oxidation (18, 19). High levels of versatility and functional group Previous approaches to the cross-coupling of
Furthermore, Akita and co-workers have used tolerance were observed with regard to the aryl benzylboron compounds with aryl halides have
Ir[dFCF 3 ppy] 2 (bpy)PF 6 (dFCF 3 ppy = 2-(2,4- halide partner. Substrates bearing electrophilic required excess (3 equiv) aqueous base and tem-
difluorophenyl)-5-(trifluoromethyl)pyridine; bpy = functional groups that would be incompatible peratures no lower than 60°C (28–30). Further-
bipyridine) as a catalyst for the oxidation of ac- with more highly reactive organometallic nucleo- more, the present reaction makes use of air-stable
tivated potassium organotrifluoroborates, dem- philes were well tolerated. Protic functional and inexpensive bipyridine ligands with low
onstrating the feasibility of their implementation groups, including amide 27, sulfonamide 39, loading of the Ni catalyst. A derivative of photo-
in the proposed single-electron transmetalation phenol 26, pyrazole 32, and -NHBoc 28, could catalyst 4 has recently been made commercially
manifold (20, 21). also be used. Substrates possessing substituents available and is similarly effective in promot-
Studies were initiated with nickel as a result ortho to the halide (25, 37) were tolerated, albeit ing the desired reactivity.
of its high reactivity toward organic halides and in diminished yield. The absence of a strong base The reported cross-coupling reactions gener-
its favorable single-electron redox potentials. permitted the coupling of amino acid derivative ally exhibited levels of efficiency and functional
We anticipated that the combination of a mono- 28 with no observable epimerization, demon- group tolerance equal to or surpassing those of

Downloaded from http://science.sciencemag.org/ on March 14, 2018


meric Ni(0) catalyst 1 and an aryl halide 2 would strating the potential utility of this method for traditional cross-coupling reactions on similar
result in rapid oxidative addition, generating late-stage functionalization of peptides or for use substrates. Most reactions cleanly afforded the
Ni(II) species 3. Concomitantly, visible-light with molecules containing other base-sensitive desired product, with the remaining mass balance
irradiation of Ir[dFCF3ppy]2(bpy)PF6 4 would functional groups. consisting of only unreacted aryl halide. Com-
generate the excited-state complex 5, the reduc- A variety of nitrogen-containing heteroaryl peting homocoupling of the trifluoroborate to
tion potential of which is sufficiently high [elec- bromides—classically challenging yet highly valued afford bibenzyl derivatives was undetectable by
trochemical potential of reduction Ered = +1.21 V substrates because of their prevalence in biolog- crude high-performance liquid chromatography
(22, 23)] to induce single-electron oxidation of an ically active compounds (27)—performed well un- analysis, allowing use of only a slight excess (1.2
activated alkyltrifluoroborate 6 [electrochemical der the optimized reaction conditions. Pyridine equiv) of this reaction partner, which is typical
potential of oxidation Eox = –1.10 V (20)], afford- substrates were coupled in all possible regioiso- in traditional Suzuki-Miyaura cross-couplings.
ing the desired alkyl radical 7 upon fragmen- meric configurations (29, 30, 31, 33). Other Also of note is the compatibility of this reaction
tation. Subsequent capture of the alkyl radical important N-heterocycles, including pyrimidine manifold with functional groups susceptible to
at Ni(II) would then yield high-valent Ni(III)
intermediate 9, which was expected to undergo
reductive elimination to generate the desired cross-
coupled product 10 and Ni(I) complex 11. From
here, reduction of 11 [Ered > –1.10 V (24, 25)] by the
reduced form of the photocatalyst 8 [Eox = +1.37 V
(21)] would regenerate both the Ni(0) catalyst
1 and the Ir catalyst 4, closing the dual cata- δ
lytic cycle.
As a proof of concept, this dual catalytic single-
electron transmetalation approach was applied
to the cross-coupling of benzylic trifluoroborates
and aryl bromides (Fig. 2). Our efforts were quickly
rewarded, as a catalytic system consisting of pho-
tocatalyst 4, Ni(COD)2 (COD = 1,5-cyclooctadiene),
4,4-di-tert-butyl-2,2′-bipyridine (dtbbpy) as lig-
and, and 2,6-lutidine as an additive effected the
cross-coupling of potassium benzyltrifluorobo- hν
rate and bromobenzonitrile in 89% yield upon
exposure to visible light from a 26-W compact
fluorescent light bulb at room temperature for
24 hours. Control reactions performed in the
absence of photocatalyst, Ni catalyst, or light
resulted in no detectable product formation,
confirming the essential role of each of these
components in the dual catalytic process (26).
We next analyzed the scope of the reaction with
regard to both the benzylic trifluoroborate and
aryl halide. As expected, electronic modification
of the trifluoroborate component had a mod-
erate effect on reaction yield, with more electron-
rich, and thus more highly stabilized, radical Fig. 1. Comparison of transmetalation in the palladium-catalyzed Suzuki-Miyaura cross-coupling
precursors (16 and 18) performing better than and the proposed single-electron transmetalation in photoredox/nickel cross-coupling. Ir =
those substituted with electron-withdrawing Ir[dFCF3ppy]2(bpy)PF6, R = generic organic subunit, Ar = aryl group.

434 25 JULY 2014 • VOL 345 ISSUE 6195 sciencemag.org SCIENCE


RE S EAR CH | R E P O R T S

single-electron oxidation or potentially reactive potentials (20, 21) serve as the singular com- of diarylmethane product 12 in 91% isolated
toward the radical intermediates, including phe- monality between these structurally dissimilar yield, with no observable biaryl formation (Fig. 4).
nol, anilide, and thienyl substructures, as well as reagents. Thus, C-C bond formation via single- The ability to engage a Csp3-hybridized organo-
five-membered nitrogen heterocycles. Remark- electron transmetalation proceeded smoothly metallic reagent selectively in a transition metal–
ably, even 4-bromostyrene could be used as an under these extremely mild and unoptimized catalyzed C-C bond-forming reaction in the presence
electrophile without competitive radical capture conditions. The differential reactivity between of an equivalent Csp2-hybridized organometallic
or polymerization, affording diarylmethane 40 the two-electron and single-electron transme- represents a striking reversal of the reactivity
in 61% yield. The surprising compatibility of the talation processes is underscored by the stark hierarchy of previously reported cross-coupling
reaction manifold with this substrate may sug- differences in conditions previously reported reactions. This effectively demonstrates the com-
gest a mechanistic scenario more complex than for (a-alkoxy)alkyltrifluoroborate cross-coupling plementary reactivity patterns observed between
that proposed in Fig. 1. However, the radical (5 equiv CsOH, 105°C, 24 hours) versus those de- the single-electron and two-electron transmetal-
nature of the activation mode is strongly sup- scribed herein (31). The observed reactivity also ation modes.
ported by the previous studies of Akita (20, 21) serves to demonstrate tolerance of substrates Another important implication of the single-
and our own experiments using chiral ligand possessing b-hydrogens—a characteristic that is electron transmetalation manifold is related to
scaffolds and racemic secondary alkyl nucleo- requisite for any general method for the cross- the stereochemical outcome of the alkyl transfer.
philes (see below). coupling of alkyl substructures. Nearly all cross-coupling reactions of stereo-
As a demonstration of the broader potential of To highlight further the differences between the defined nucleophiles reported heretofore have
the application of single-electron transmetalation activation mode reported here and that of tradi- demonstrated the transmetalation event to be
in this dual catalytic cross-coupling, the conditions tional cross-coupling, we performed a competition stereospecific (32). Thus, enantioenriched products
optimized for use with primary benzylic trifluoro- experiment between potassium benzyltrifluoro- may only be accessed from nonracemic and con-
borates were directly applied to the cross-coupling borate and potassium phenyltrifluoroborate. Expo- figurationally stable organometallic reagents,

Downloaded from http://science.sciencemag.org/ on March 14, 2018


of a secondary (a-alkoxy)alkyltrifluoroborate sure of these two nucleophiles to the photoredox which are often difficult to access. Isolated ex-
(Fig. 3). Comparable single-electron oxidation cross-coupling conditions resulted in isolation amples of stereoconvergence in transmetalation

Fig. 2. Photoredox cross-coupling of benzylic trifluoroborates and aryl bromides. All yields are percent isolated yield of pure material after chromatography.
Reactions were performed on aryl halide (0.5 mmol). Boc = tert-butoxycarbonyl, Me = methyl, Ph = phenyl, Ac = acetyl.

SCIENCE sciencemag.org 25 JULY 2014 • VOL 345 ISSUE 6195 435


R ES E A RC H | R E PO R TS

exist, specifically in the context of secondary under slightly modified conditions, racemic tri- REFERENCES AND NOTES
benzylmagnesium reagents, a pyrrolidine-based fluoroborate 44 was engaged in stereoconvergent 1. A. de Meijere, F. Diederich, Eds., Metal-Catalyzed Cross-Coupling
organozinc, and a diastereoconvergent cross- cross-coupling with methyl 3-bromobenzoate, Reactions (Wiley-VCH, Weinheim, Germany, ed. 2, 2004).
2. J. F. Hartwig, Organotransition Metal Chemistry: From Bonding
coupling of substituted cyclohexylzinc reagents affording 1,1-diarylethane product 45 in 52% to Catalysis (University Science, Sausalito, CA, 2010).
(33–36). Stereoconvergence in the former is yield and a promising enantiomeric ratio of 75:25 3. A. Rudolph, M. Lautens, Angew. Chem. Int. Ed. 48, 2656–2670
thought to be enabled by dynamic kinetic res- (Fig. 4). The observed stereoconvergence serves (2009).
olution of the configurationally unstable Grignard as an effective mechanistic probe that supports 4. L. S. Hegedus, B. C. G. Soderberg, Transition Metals in the
Synthesis of Complex Organic Molecules (University Science,
reagent; the origin of selectivity in the latter is the role of the organotrifluoroborate as a car- Sausalito, CA, ed. 3, 2010).
not fully understood. None of these approaches bon radical precursor, provides evidence that 5. H. Doucet, Eur. J. Org. Chem. 2008, 2013–2030 (2008).
constitute a general strategy for stereoconver- the radical is intercepted by the ligated Ni com- 6. D. Imao, B. W. Glasspoole, V. S. Laberge, C. M. Crudden, J. Am.
gent transmetalation beyond the scope of the plex, and suggests that C-C bond formation oc- Chem. Soc. 131, 5024–5025 (2009).
7. G. Zou, Y. K. Reddy, J. R. Falck, Tetrahedron Lett. 42,
directly explored reagents. curs via reductive elimination from Ni. 7213–7215 (2001).
In contrast, the stereochemical outcome of This preliminary result strongly implies that 8. J. Z. Deng et al., Org. Lett. 11, 345–347 (2009).
the single-electron transmetalation is dictated high levels of stereoselectivity are possible in 9. M. Aufiero, F. Proutiere, F. Schoenebeck, Angew. Chem. Int. Ed.
by facial selectivity of the addition of a prochiral the photoredox cross-coupling of secondary alkyl 51, 7226–7230 (2012).
10. Transmetalation rates of organoboron compounds adhere to
alkyl radical to a ligated Ni center. Thus, appli- nucleophiles with appropriate modification of trends established for the corresponding organostannanes.
cation of a chiral ligand framework renders reaction conditions and ligand structure. Refine- 11. J. W. Labadie, J. K. Stille, J. Am. Chem. Soc. 105, 6129–6137 (1983).
this process asymmetric and provides a general ment of this approach to asymmetric cross-coupling 12. A. Finch, P. J. Gardner, E. J. Pearn, G. B. Watts, Trans. Faraday
reaction manifold in which stereoconvergent will provide a powerful advancement to the field Soc. 63, 1880–1888 (1967).
13. C. K. Prier, D. A. Rankic, D. W. C. MacMillan, Chem. Rev. 113,
transmetalation can be achieved. Well-known by alleviating the need for synthesis of enantio- 5322–5363 (2013).
stereoconvergent cross-couplings of alkyl halides, enriched organometallic reagents. Taken together, 14. D. M. Schultz, T. P. Yoon, Science 343, 1239176 (2014).

Downloaded from http://science.sciencemag.org/ on March 14, 2018


which putatively undergo a similar mechanistic our findings effectively validate the single-electron 15. D. Kalyani, K. B. McMurtrey, S. R. Neufeldt, M. S. Sanford,
step, provide guidance for selection of appro- transmetalation manifold and dual photoredox/ J. Am. Chem. Soc. 133, 18566–18569 (2011).
16. Y. Ye, M. S. Sanford, J. Am. Chem. Soc. 134, 9034–9037 (2012).
priate ligand scaffolds to maximize the stereo- cross-coupling cycle as a viable alternative to 17. B. Sahoo, M. N. Hopkinson, F. Glorius, J. Am. Chem. Soc. 135,
selectivity of the radical capture (37). Indeed, conventional cross-coupling of Csp3-hybridized 5505–5508 (2013).
when we used commercially available ligand L1 nucleophiles. 18. J. W. Lockner, D. D. Dixon, R. Risgaard, P. S. Baran, Org. Lett.
13, 5628–5631 (2011).
19. G. A. Molander, V. Colombel, V. A. Braz, Org. Lett. 13,
1852–1855 (2011).
20. Y. Yasu, T. Koike, M. Akita, Adv. Synth. Catal. 354, 3414–3420
(2012).
21. K. Miyazawa, Y. Yasu, T. Koike, M. Akita, Chem. Commun. 49,
7249–7251 (2013).
22. All electrochemical potentials are calculated versus the
saturated calomel electrode (SCE).
23. M. Lowry et al., Chem. Mater. 17, 5712–5719 (2005).
24. C. Cannes, E. Labbé, M. Durandetti, M. Devaud, J. Y. Nédélec,
J. Electroanal. Chem. 412, 85–93 (1996).
25. The reduction of (bpy)NiBr2 is observed as a two-electron
Fig. 3. Photoredox cross-coupling of a secondary (a-alkoxy)alkyltrifluoroborate with 4-bromobenzonitrile. wave at –1.1 V, implying that reduction of (bpy)NiBr to
(bpy)Ni0 occurs at a more positive potential.
26. See supplementary materials on Science Online.
27. M. E. Welsch, S. A. Snyder, B. R. Stockwell, Curr. Opin. Chem.
A Biol. 14, 347–361 (2010).
28. P. Jain, S. Yi, P. T. Flaherty, J. Heterocycl. Chem. 50, (S1),
E166–E173 (2013).
29. A. Flaherty, A. Trunkfield, W. Barton, Org. Lett. 7, 4975–4978 (2005).
30. G. A. Molander, T. Ito, Org. Lett. 3, 393–396 (2001).
31. G. A. Molander, S. R. Wisniewski, J. Am. Chem. Soc. 134,
16856–16868 (2012).
32. L. Li, C.-Y. Wang, R. Huang, M. R. Biscoe, Nat. Chem. 5,
607–612 (2013).
33. T. Hayashi, M. Tajika, K. Tamao, M. Kumada, J. Am. Chem. Soc.
98, 3718–3719 (1976).
34. T. Hayashi et al., J. Am. Chem. Soc. 104, 180–186 (1982).
35. C. J. Cordier, R. J. Lundgren, G. C. Fu, J. Am. Chem. Soc. 135,
B 10946–10949 (2013).
36. K. Moriya, P. Knochel, Org. Lett. 16, 924–927 (2014).
37. H. Q. Do, E. R. R. Chandrashekar, G. C. Fu, J. Am. Chem. Soc.
135, 16288–16291 (2013).

AC KNOWLED GME NTS


Supported by NIH (grant R01 GM-081376) and Pfizer. We thank
Frontier Scientific for providing all of the organoboron precursors
used in our work.

SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/345/6195/433/suppl/DC1
Materials and Methods
Fig. 4. Probing chemo- and stereoselectivity. (A) Competition experiment between potassium Supplementary Text
Figs. S1 to S9
benzyltrifluoroborate and potassium phenyltrifluoroborate under photoredox cross-coupling conditions.
NMR spectra
(B) Stereoconvergent cross-coupling of a racemic trifluoroborate 44 and aryl bromide to afford an References (38–54)
enantioenriched product. Reactions were performed on aryl bromide (0.5 mmol). *Determined by chiral
19 March 2014; accepted 27 May 2014
supercritical fluid chromatography (SFC). †Absolute configuration was assigned as (S) on the basis of Published online 5 June 2014;
data reported in the literature. er = enantiomeric ratio. 10.1126/science.1253647

436 25 JULY 2014 • VOL 345 ISSUE 6195 sciencemag.org SCIENCE


Single-electron transmetalation in organoboron cross-coupling by photoredox/nickel dual
catalysis
John C. Tellis, David N. Primer and Gary A. Molander

Science 345 (6195), 433-436.


DOI: 10.1126/science.1253647originally published online June 5, 2014

A bright outlook for carbon coupling


In contemporary organic chemistry, it is straightforward to forge bonds between unsaturated carbons (i.e., carbons
already engaged in double bonds) using cross-coupling catalysis. The protocol runs into some trouble, however, if one or
both starting carbon centers are saturated (purely single-bonded). Tellis et al. and Zuo et al. independently found that

Downloaded from http://science.sciencemag.org/ on March 14, 2018


combining a second, light-activated catalyst with a nickel cross-coupling catalyst could achieve selective coupling of
saturated and unsaturated reagents (see the Perspective by Lloyd-Jones and Ball). Their methods rely on single-electron
transfer from the light-activated catalyst to the saturated carbon, thereby enhancing its reactivity more effectively than the
twoelectron mechanisms prevailing in traditional protocols.
Science, this issue p. 433, p. 437; see also p. 381

ARTICLE TOOLS http://science.sciencemag.org/content/345/6195/433

SUPPLEMENTARY http://science.sciencemag.org/content/suppl/2014/06/04/science.1253647.DC1
MATERIALS

RELATED http://science.sciencemag.org/content/sci/345/6195/381.full
CONTENT
http://science.sciencemag.org/content/sci/345/6195/437.full

REFERENCES This article cites 47 articles, 1 of which you can access for free
http://science.sciencemag.org/content/345/6195/433#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. 2017 © The Authors, some rights reserved; exclusive
licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. The title
Science is a registered trademark of AAAS.

You might also like