Вы находитесь на странице: 1из 7

R ES E A RC H

ORGANIC CHEMISTRY results in a para-selective reaction to form de-


aromatized intermediate Int-I (31–37). Two
equivalents of DBU (1,8-diazabicyclo[5.4.0]undec-
Heterobiaryl synthesis by contractive 7-ene) eliminate first the triflyl anion to form
phosphonium ion Int-II, and then methyl acry-

C–C coupling via P(V) intermediates late to form heteroaryl phosphine 2a in good
yield. Pyridine was chosen as the second cou-
pling partner in stage B with phosphine 2a as
Michael C. Hilton1, Xuan Zhang1*, Benjamin T. Boyle1*, Juan V. Alegre-Requena1, a nucleophile, resulting in bis-heteroaryl phos-
Robert S. Paton1,2†, Andrew McNally1† phonium salt 3a, with complete regiocontrol.
Several nucleophiles are known to initiate phos-
Heterobiaryls composed of pyridine and diazine rings are key components of pharmaceuticals phorus ligand coupling, including alkoxides,
and are often central to pharmacological function. We present an alternative approach Grignard reagents, and acidic alcohol solutions
to metal-catalyzed cross-coupling to make heterobiaryls using contractive phosphorus (22, 25–29); for stage C, we found the latter to
C–C couplings, also termed phosphorus ligand coupling reactions. The process starts by be most effective and two equivalents of HCl in
regioselective phosphorus substitution of the C–H bonds para to nitrogen in two successive EtOH at 80°C to be optimal, forming hetero-
heterocycles; ligand coupling is then triggered via acidic alcohol solutions to form the biaryl 4a in excellent yield with diphenylphos-
heterobiaryl bond. Mechanistic studies imply that ligand coupling is an asynchronous process phine oxide as a by-product (see table S1). We
involving migration of one heterocycle to the ipso position of the other around a central did not observe products from heteroaryl-phenyl
pentacoordinate P(V) atom. The strategy can be applied to complex drug-like molecules or phenyl-phenyl coupling, nor ethoxylation of
either heterocycle, in this protocol.

Downloaded from http://science.sciencemag.org/ on November 15, 2018


containing multiple reactive sites and polar functional groups, and also enables convergent
coupling of drug fragments and late-stage heteroarylation of pharmaceuticals.
Mechanistic investigation

R
To investigate the reasons for selective heterocycle-
eactions that couple two aromatic rings to Another serious problem arises from the lack of heterocycle coupling and the kinetics of the
make biaryls are among the most widely methods to prepare the cross-coupling precur- ligand-coupling process, we performed a series
used processes in the pharmaceutical in- sors. Although simple heteroaryl halides are com- of experimental and computational studies. We
dustry (1, 2). Coupling of pyridines and mercially available or can be straightforwardly hypothesized that ethanol attacks the phospho-
diazines results in heterobiaryls, a privi- prepared, direct and selective halogenation of rus center and a P(V) species is formed. Subject-
leged pharmacophore found in commercial drugs pyridine and diazine derivatives encountered ing salt 3a to a solution of DCl in d4-methanol
as well as numerous therapeutic candidates, such during drug development remains an unsolved results in successive shifts of pyridine proton
as the examples shown in Fig. 1A (3–5). These challenge (18, 19). Similarly, synthesizing nu- resonances per equivalent of acid by 1H NMR
heterocycles often play a key role in drug-receptor cleophilic coupling partners such as heteroaryl (nuclear magnetic resonance) and 31P NMR
binding and impart other important properties boronic acids, stannanes, and organozinc or spectroscopy, and indicates that both pyridines
such as net polarity, aqueous solubility, and re- magnesium compounds is challenging, and are protonated (see figs. S17 and S18). However,
sistance to oxidative metabolism. Most con- they are often prepared from the correspond- no P(V) intermediates were detected in a 31P NMR
ceivable aryl-aryl couplings can be accomplished ing heteroaryl halides to begin with (20). Cross- study under the reaction conditions. Computa-
using metal-catalyzed cross-coupling reactions; dehydrogenative couplings of heteroarenes have tional studies do predict that intramolecular
these processes feature exceptional chemoselec- shown some promise but are currently limited ligand coupling occurs from P(V) intermediate
tivity, precise regioselectivity, and sufficient ro- to specific pyridine combinations and are not Int-III in a stepwise fashion (see below) and
bustness to be applied to both drug discovery applicable in complex settings (21). that there is a substantial barrier-lowering effect
and manufacture (6–8). However, the same syn- (DG‡) upon successive protonation of Int-III
thetic prowess is not transferable to heteroaryl- Reaction development (Fig. 2A) (38). Transition state energies consider-
heteroaryl coupling, particularly for complex The limitations of current heterocycle coupling ably favor pyridine-pyridine coupling over pyridine-
substrates. An alternative strategy that addresses methods can potentially be overcome by con- phenyl coupling for each protonation state;
the shortcomings in this fundamental bond con- tractive phosphorus C–C couplings, often termed DGreact values show that the process is similarly
struction would therefore offer new opportunities phosphorus ligand coupling reactions; a test exergonic and irreversible in each case, reinforcing
to incorporate heterobiaryls into therapeutic system is shown in Fig. 1C (22–24). The strategy the conclusion that selective pyridine-pyridine
candidates. does not rely on heteroaryl halides or partners coupling results from kinetic differences in the
For de novo synthesis of heterobiaryls, a sche- such as boronic acids, but instead regioselectively ligand-coupling transition state rather than ther-
matic for metal-catalyzed cross-couplings is shown substitutes the C–H bond in each heterocyclic modynamic factors. The intrinsic reaction co-
in Fig. 1B (9–15). A minimum of three steps are coupling partner by successive C–P bond forma- ordinate (IRC; Fig. 2B) shows no involvement
required, and there are challenges in the coupling tions to produce a bis-azaarene phosphonium of alkoxy lone pairs and negligible changes to
step, such as catalyst poisoning and decompo- salt; phosphorus ligand coupling is then trig- the other three equatorial P–C bonds. In the C–C
sition of starting materials (16). Furthermore, gered to form the heterobiaryl bond via a P(V) bond-forming transition structure [TS-I·2H]2+,
drug-like molecules and intermediates often intermediate. Heteroaryl-heteroaryl coupling has a single P–C bond breaks, allowing one ligand
have multiple reactive sites and a high proportion previously been observed at phosphorus centers, to migrate to the ipso-carbon of another (Fig.
of polar functional groups, such as basic amines, but an inability to transform a generic set of pyr- 2C). The intermediate formed in this key step
that interfere with catalytic processes and cause idines and diazine precursors into the required ([Int-IV·2H]2+) is a dearomatized adduct char-
a considerable number of them to fail (15, 17). bis-azaarene phosphonium salts has restricted acteristic of nucleophilic aromatic substitution,
these processes to specialized cases (25–29). In which is predicted to collapse irreversibly (DG =
our test system, stage A combined the first het- –39 kcal/mol) and with considerable ease (see
1
Department of Chemistry, Colorado State University, Fort erocycle, 2-phenylpyridine, with Tf2O at low figs. S9 to S11). This stepwise ligand coupling is
Collins, CO 80523, USA. 2Chemistry Research Laboratory, temperature to form an intermediate pyridin- therefore mechanistically distinct from the con-
University of Oxford, Oxford OX1 3TA, UK.
*These authors contributed equally to this work.
ium triflyl salt (not shown); adding fragment- certed cleavage of two s-bonds during reduc-
†Corresponding author. Email: robert.paton@chem.ox.ac.uk able phosphine 1 (prepared on large scale from tive elimination at, for example, Pd(II) or in
(R.S.P.); andy.mcnally@colostate.edu (A.M.) diphenyl phosphine and methyl acrylate) (30) dihydrogen formation from PH5. This latter

Hilton et al., Science 362, 799–804 (2018) 16 November 2018 1 of 6


R ES E A RC H | R E S EA R C H A R T I C LE

detail is important because concerted coupling apical positions: s-electron–withdrawing alkoxy s-electron–withdrawing power of the axial donor
of apical-equatorial substituents from a (D3h) and heteroaryl groups preferentially occupy the ligand and weakening the P–C bond (dP–C(py)
trigonal bipyramidal compound is symmetry- apical sites (fig. S5). Weaker and more polar apical increases from 1.95 Å in [Int-III·H]+ to 1.99 Å in
forbidden (fig. S6) (24): in contrast, this stepwise- P–L bonds favor migration in nucleophilic 1,2- [Int-III·2H]2+), whereas equatorial P–C bonds
coupling mechanism permits, and indeed favors, rearrangements, in which an equatorial ligand are largely unchanged (dP–C(py) is 1.87 Å in [Int-
the migration of an apical ligand to an equatorial acts as the electrophilic acceptor (fig. S7), lead- III·H]+ and 1.86 Å in [Int-III·2H]2+). Computed
one ([TS-I·2H]2+) (fig. S7). ing to ligand coupling. Phenyl ligands are un- values of C–O coupling from Int-[Int-III·2H]2+
The computed structures of P(V) intermedi- favorable for both donor and acceptor roles in are also disfavored relative to pyridine-pyridine
ates, such as [Int-III·2H]2+, are characterized ligand coupling: Apical positions (donors) fa- coupling (DG ‡(C–O) = 18 kcal/mol versus DG ‡(py-py) =
by stronger, shorter (dP–C(py) = 1.86 Å) bonds to vor more s-electron–withdrawing substituents, 14 kcal/mol; fig. S8) (31).
equatorial ligands and weaker, longer (dP–C(py) = whereas pyridyl substituents are superior accep- Figure 2D examines the effect of phosphorus
1.99 Å) bonds to those in apical positions. This is tors. This explains the complete absence of bi- electrophilicity on the rate of heterobiaryl forma-
a result of three-center, four-electron bonding be- phenyl and phenyl-heterobiaryl coupled products. tion. The low energy barrier for ligand coupling in
tween the apical ligands and the central phos- N-protonation decreases the activation barrier [Int-III·2H]2+ implies that the rate-determining
phorus atom. Accordingly, the relative stability considerably, from 30 to 20 kcal/mol, increasing step precedes this event and involves attack of
of P(V) stereoisomeric forms can be readily the electrophilicity of the equatorial pyridyl group. the alcoholic solvent at the phosphonium center.
predicted on the basis of each ligand’s capacity Successive N-protonation further reduces the We prepared a set of salts with substituted aryl
to stabilize the buildup of electron density at the activation barrier to 14 kcal/mol by increasing the groups that would change the electrophilicity at

Downloaded from http://science.sciencemag.org/ on November 15, 2018


A Me
MeO2S
Fig. 1. Important
heterobiaryl-containing
N N
Me N drugs and synthetic
H Cl
N strategies. (A) Hetero-
N N
N O OMe biaryls in drugs.
N N N (B) Heterobiaryls via
N
HN metal-catalyzed cross-
Me N OMe coupling reactions.
HN
Cl R denotes a general
Imatinib (Gleevec) O organic group; Hal,
Etoricoxib (Arcoxia) MK-1064
- Myelogenous Leukemia - Pain and inflammation - Insomnia halogen substituent.
(C) Test system for
B Heterobiaryl synthesis: Metal-catalyzed cross-coupling schematic heterobiaryl synthesis
via phosphorus ligand
H Hal
coupling reactions. Ph,
Heterobiaryls phenyl; Me, methyl; Et,
R1 R1 ethyl; Tf, trifluorometh-
R2 ylsulfonyl; DBU,
N N
N 1,8-diazabicyclo[5.4.0]
H B(OH)2 R1 undec-7-ene; rt,
X room temperature.
X X N
R2 R2
N N

C Heterobiaryl synthesis: Initial study into a strategy using phosphorus ligand coupling reactions

CO2Me CO2Me CO2Me


H DBU PPh2
1
PPh2 Ph2P
H PPh2

OTf OTf
N Ph A N Ph

N Ph N Ph CO2Me
2a, 71%
Tf Int-I Int-II

N Ph N
A

C OTf B
B P

N H
C N Ph
N
4a, 88% 3a, 83%

Hilton et al., Science 362, 799–804 (2018) 16 November 2018 2 of 6


R ES E A RC H | R E S EA R C H A R T I C LE

phosphorus; rate data show faster heterobiaryl coupling process (Fig. 3). The reaction is com- were formed via the three-step sequence, with
formation as the electrophilicity of the phospho- pletely selective for the 4-position of pyridines lower yields in the coupling step relative to pyri-
nium increases, in line with the above hypothesis. in the vast majority of cases studied, unless a dine examples.
Further experimental verification of the low bar- 4-substituent is present, which switches selectiv-
riers for ligand coupling is shown in Fig. 2E. Acidic ity to the 2-position. A variety of 4,4′-bipyridines Reaction guidelines
alcohol solutions are inefficient for heterobiaryl are accessible using this strategy (4b–4f); func- During these studies, we have established a gen-
formation at lower temperatures (table S1); how- tional groups such as esters, trifluoromethyl eral set of reaction guidelines and limitations.
ever, when ethoxide is used as a nucleophile for groups, and methoxy groups are accommodated, First, when coupling 2-substituted pyridines to
facile addition to the phosphonium ion (fig. S24), as are halides that would normally be active in 3-substituted pyridines, it is important to per-
heterobiaryl synthesis occurs in minutes at room metal-catalyzed reactions. Substituents can be form the salt-forming sequence in the correct
temperature, with trace amounts of C–O coupling present at the 2- or 3-positions of pyridines, and order (Fig. 3). Taking heterobiaryl 4b as a repre-
also observed. Substantial amounts of products example 4e shows that a 2-position substituent sentative example, if heteroaryl phosphine 2b′ is
resulting from protiodephosphination are formed is not a requirement (see below). A fluorinated used instead in stage B, then salt 3b is not formed.
under these conditions, making this protocol less 2,4′-quinoline-pyridine was also synthesized by We believe that a biased Tf-salt equilibrium rapid-
practical than that under acidic conditions. phosphorus ligand coupling (4g) (39). Examples ly develops, and pyridinium-phosphine [2b′-Tf]+
of 2,2′-systems, 4h and 4i, showcase an alter- is favored on steric grounds; the 2-substituted
Substrate scope exploration native to Suzuki couplings, where 2-pyridyl and pyridine is then not activated for nucleophilic
We next selected a set of pyridines and diazines to quinolyl boronic acids often decompose during addition, and the desired salt is not formed (fig.
examine which substitution patterns and func- metal-catalyzed reactions (16). Pyrimidine- and S25). Instead, the 2-substituted pyridine should
tional groups could be tolerated in the ligand pyrazine-containing heterobiaryls 4j and 4k be converted into the corresponding phosphine

Downloaded from http://science.sciencemag.org/ on November 15, 2018


A X X X
B

(all energies
in kcal/mol)
P N P NH P NH

Y O Y O Y O
Me Me Me

X Y (N or CH) X Y (N or CH) X Y (NH+ or CH)


N

Intrinsic Reaction Coordinate (IRC)

C H
N
= Ph
Apical
= 1.99Å H
N
Equatorial

P NH
= 1.86Å

O P
Me
O N
apicophilic effect Me H

asynchronous coupling

D Ph N
N R krel E N Ph
N Me

O H
OTf OMe 0.16
P OTf
Me 0.37
R P
N H (3a) 1.00 N N
N N Ph
N Ph Cl 1.89
R 4a, 46% <1%

Fig. 2. Computational and experimental investigation of phosphorus each oxygen lone pair. (C) Optimized structures for [Int-III·2H]2+,
ligand coupling. (A) Biaryl-coupling activation barriers [coupled cluster/ [TS-I·2H]2+, and [Int-IV·2H]2+ show stepwise apical-equatorial ligand
density functional theory SMD-DLPNO-CCSD(T)/cc-pV(DT)Z//wB97XD/ coupling. (D) A kinetic study indicates that alcohol addition is rate-limiting.
6-31+G(d), kcal/mol] decrease upon protonation and consistently favor TfOH was used in place of HCl because of poor solubility of aryl
heterobiaryl formation. Py, pyridine. (B) Computed bond orders show derivatives. Yields after complete consumption of the phosphonium salts
a single (apical) P–C(py) bond breaking along the reaction coordinate, were approximately the same in each case (89 to 94%). (E) Room-
with little involvement of oxygen lone pairs. LPOx, number of electrons in temperature coupling using ethoxide as a nucleophile.

Hilton et al., Science 362, 799–804 (2018) 16 November 2018 3 of 6


R ES E A RC H | R E S EA R C H A R T I C LE

R2
H PPh2 H N
R1 A
A B X
X
R1 R1 R2 B
N
N N N C
C
Heteroaryl
Pyridine Azaarene Heterobiaryl, 4
Phosphine, 2

N N Cl N Cl N

n-Bu N
PPh2
N Cl N CO2Et N OMe N F Cl
B Cl
Ph-4-Br Me OTf
Br-4-Ph
4b A 61% 4c A 61% 4d A 65% 4e A 91% Ph P Ph-4-Br
N N Ph-4-Br
Ph
B 82% C 64% B 74% C 76% B 74% C 65% B 54% C 65% N
3b, 0%
F3C
N N CO2Et
Tf-salt equilibrium biased towards inactive heteroaryl phosphines
Me
n-Bu PPh2
N Br N F EtO
N Cl

Downloaded from http://science.sciencemag.org/ on November 15, 2018


N N Ph-4-Br
4f O F 4g 4h A 82% N N Ph-4-Br
OTf Tf
+ Tf OTf
A 84% B & C 29% A 83% B 58% C 57% B 61% C 47%
F
n-Pr
S
N Me
A B
Me O
N N
EtO N
Me N Me PPh2
N N N Cl
N B Cl
OTf
Cl P Ph-4-Br
4i 4j 4k N Ph-4-Br N Ph
Ph
A 83% B 53% C 67% A 83% B 72% C 37% A 82% B 90% C 23% N
2b 3b, 82%

Fig. 3. Azaarene scope and guidelines for phosphonium salt formation. Yields of isolated products after each stage are shown. n-Bu, normal butyl
group; n-Pr, normal propyl group. Reaction guidelines are shown for phosphonium salt formation involving ortho and non-ortho substituted pyridines as
partners. Further details of challenges and limitations are highlighted in fig. S25.

and used as a nucleophile with the 3-substituted couplings of pyridine-containing fragments were The use of previously synthesized heteroaryl
pyridine in stage B. Second, problematic sub- first examined; these molecules are representative phosphines (Fig. 4) shows that heteroarylation is
strates for heteroaryl phosphine and salt for- of drug leads, which are promising candidates for possible in these complex systems with complete
mation include pyridines with 2-trifluoromethyl a therapeutic target but have suboptimal phar- control of regioselectivity and site selectivity.
groups, 4-alkyl or aryl substituents, and 2,6- macokinetic and pharmacodynamic properties Chlorphenamine, a common antihistamine, and
disubstituted pyridines. In general, pyridines and (40). A convergent coupling strategy would en- loratadine, an allergy medicine, are competent
diazines with more than two electron-withdrawing able rapid access to complex heterobiaryls from substrates for this protocol, with the resulting
groups or electron-donating groups can result in compounds common in pharmaceutical libraries heterobiaryls isolated in good overall yields (4p
low yields or no phosphonium salt formation. (41, 42). Four examples in Fig. 4 are shown where and 4q) that again highlight how halides can
During ligand coupling, we have observed that the corresponding halide precursors are not com- be tolerated during the coupling procedure.
pyridines substituted with bromides and io- mercially available or would be challenging to Vismodegib was converted into a 2,4′-quinoline-
dides can be dehalogenated, that 2-chloro- or prepare (4l–4o). Heterobiaryl bonds are formed pyridine system in moderate yield (4r). A widely
2-fluoropyridines are not successful, and that with precise regioselectivity, and the presence of applied fungicide, quinoxyfen, was also compa-
2-methoxypyridines proceed with slower rates. additional saturated and unsaturated nitrogen tible with the reaction protocol (4s). Etoricoxib
For pyridines containing electron-withdrawing heterocycles is tolerated in this approach. Three and imatinib are challenging examples because
groups, using EtOH and HCl can result in eth- or four equivalents of acid are used in the cou- they contain multiple reactive sites (34). The
oxylation. Changing the acid to TfOH avoids this pling step in these cases to ensure adequate reac- structural features in etoricoxib enable selective
problem and leads us to believe that ethoxylation tion rates. transformation of the 2,5-disubstituted pyridine
results from chlorination followed by ethoxylation Next, we investigated whether the ligand- (4t), and heteroarylation of the pyridine occurs
via nucleophilic aromatic substitution. Trifluor- coupling strategy could be applied to advanced selectively over the pyrimidine in imatinib to
oethanol is preferred when molecules contain intermediates in drug development. Success in form 4u.
functional groups such as amides and esters this endeavor would offer distinct strategies to
that are susceptible to ethanolysis. In general, introduce heterobiaryls into complex molecules Outlook
one equivalent of acid per basic nitrogen is and alleviate concerns over metal contamina- This phosphorus ligand coupling method over-
optimal (see below). tion in subsequent biological testing. To demon- comes major limitations of metal-catalyzed
strate the feasibility of this approach, we chose approaches by virtue of its compatibility with
Application to complex intermediates a set of existing drug molecules with diverse polar functionalities found in drug-like mol-
Our attention then turned to ligand couplings structures, substitution patterns, and functional ecules and its circumvention of preformed het-
involving complex azaarenes (Fig. 4). Convergent groups (43). eroaryl halides and boronic acids. As well as

Hilton et al., Science 362, 799–804 (2018) 16 November 2018 4 of 6


R ES E A RC H | R E S EA R C H A R T I C LE

R2
H PPh2 H N A
R1
A B
B
R1 R1 R2
N
N N N C C
Heteroaryl
Pyridine Azaarene Heterobiaryl, 4
Phosphine, 2

Convergent coupling of pyridine- and quinoline-containing fragments


Me
N N

N O
F
O Me
Me N
N Cl Cl
F N N O
F N
N N N O Me
O O
Me N F3C
F N
F3C N N
CO2Et CO2Et
N
4l 4m 4n A 65% 4o

A 78% B 70% C 60% A 78% B 63% C 65% Me B 74% C 46% A 65% B 66% C 49%

Downloaded from http://science.sciencemag.org/ on November 15, 2018


Advanced-stage coupling demonstrated by complex bioactive molecules as partners A O O
N S
Ph-4-Br N Me
N Ph H
4p, from Chlorphenamine N N

B 63% C 84% O Cl
Cl Cl
Me 4q, from Loratadine Me
N
N B 81% C 87%
N Me 4r, from Vismodegib B 55% C 52%

O O Me Me
F N N
S
CO2Et Me N N
N
Cl O Cl Cl

HN N
N N Me
N N
N Cl Me
Me
O N
N H
OEt 4s, from Quinoxyfen 4t, from Etoricoxib 4u, from Imatinib
B 49% C 58% B 69% C 80% B 61% C 41%

Fig. 4. Heterobiaryl synthesis in complex molecules. Yields of isolated products after each stage are shown. Further examples of advanced stage
couplings are shown in fig. S26.

transforming building block compounds, con- 9. L.-C. Campeau, K. Fagnou, Chem. Soc. Rev. 36, 1058–1068 (2007). 27. Y. Uchida, K. Onoue, N. Tada, F. Nagao, S. Oae, Tetrahedron
vergent coupling of drug fragments and hetero- 10. D. Zhao, J. You, C. Hu, Chem. Eur. J. 17, 5466–5492 (2011). Lett. 30, 567–570 (1989).
11. K. L. Billingsley, K. W. Anderson, S. L. Buchwald, Angew. Chem. 28. Y. Uchida, H. Kozawa, S. Oae, Tetrahedron Lett. 30,
arylation of complex pharmaceuticals were Int. Ed. 45, 3484–3488 (2006). 6365–6368 (1989).
demonstrated. The protocol uses readily avail- 12. N. Kudo, M. Perseghini, G. C. Fu, Angew. Chem. Int. Ed. 45, 29. Y. Uchida, N. Echikawa, S. Oae, Heteroatom Chem. 5, 409–413
able reagents under simple conditions and is 1282–1284 (2006). (1994).
immediately applicable in medicinal chemistry. 13. A. S. Guram et al., J. Org. Chem. 72, 5104–5112 (2007). 30. F. Alonso, Y. Moglie, G. Radivoy, M. Yus, Green Chem. 14,
14. U. Kiehne, J. Bunzen, A. Lützen, Synthesis 1061–1069 (2007). 2699–2702 (2012).
15. T. Markovic, B. N. Rocke, D. C. Blakemore, V. Mascitti, 31. M. C. Hilton, R. D. Dolewski, A. McNally, J. Am. Chem. Soc. 138,
M. C. Willis, Chem. Sci. 8, 4437–4442 (2017). 13806–13809 (2016).
RE FE RENCES AND N OT ES 16. P. A. Cox et al., J. Am. Chem. Soc. 139, 13156–13165 (2017). 32. X. Zhang, A. McNally, Angew. Chem. Int. Ed. 56, 9833–9836
1. D. G. Brown, J. Boström, J. Med. Chem. 59, 4443–4458 17. D. C. Blakemore et al., Nat. Chem. 10, 383–394 (2018). (2017).
(2016). 18. J. A. Joule, K. Mills, Heterocyclic Chemistry (Wiley-Blackwell, 33. J. L. Koniarczyk, D. Hesk, A. Overgard, I. W. Davies, A. McNally,
2. S. D. Roughley, A. M. Jordan, J. Med. Chem. 54, 3451–3479 ed. 5, 2013). J. Am. Chem. Soc. 140, 1990–1993 (2018).
(2011). 19. M. R. Grimmett, Adv. Heterocycl. Chem. 58, 271–345 (1993). 34. R. D. Dolewski, P. J. Fricke, A. McNally, J. Am. Chem. Soc. 140,
3. R. Capdeville, E. Buchdunger, J. Zimmermann, A. Matter, Nat. 20. M. A. Larsen, J. F. Hartwig, J. Am. Chem. Soc. 136, 4287–4299 8020–8026 (2018).
Rev. Drug Discov. 1, 493–502 (2002). (2014). 35. R. G. Anderson, B. M. Jett, A. McNally, Angew. Chem. Int. Ed.
4. A. J. Roecker et al., ChemMedChem 9, 311–322 (2014). 21. H.-Q. Do, O. Daugulis, J. Am. Chem. Soc. 133, 13577–13586 (2011). 57, 12514–12518 (2018).
5. S. D. Martina, K. S. Vesta, T. L. Ripley, Ann. Pharmacother. 39, 22. J.-P. Finer, in Ligand Coupling Reactions with Heteroaromatic 36. E. Anders, F. Markus, Tetrahedron Lett. 28, 2675–2676
854–862 (2005). Compounds, Vol. 18 (Pergamon, 1998), chap. 4. (1987).
6. M. L. Crawley, B. M. Trost, Applications of Transition Metal 23. K. D. Reichl, A. T. Radosevich, Chem. Commun. 50, 9302–9305 37. P. S. Fier, J. Am. Chem. Soc. 139, 9499–9502 (2017).
Catalysts in Drug Discovery and Development: An Industrial (2014). 38. Both DLPNO-CCSD(T)/cc-pV(DT)Z and wB97XD/def2-QZVPP
Perspective (Wiley, 2012), pp. 25–96. 24. R. Hoffmann, J. M. Howell, E. L. Muetterties, J. Am. Chem. Soc. results are in close-agreement. Full details in the supplementary
7. A. de Meijere, F. Diederich, Metal-Catalyzed Cross-Coupling 94, 3047–3058 (1972). materials.
Reactions (Wiley-VCH, ed. 2, 2004). 25. F. G. Mann, J. Watson, J. Org. Chem. 13, 502–531 (1948). 39. O. Afzal et al., Eur. J. Med. Chem. 97, 871–910 (2015).
8. J. Hassan, M. Sévignon, C. Gozzi, E. Schulz, M. Lemaire, 26. G. R. Newkome, D. C. Hager, J. Am. Chem. Soc. 100, 40. R. B. Silverman, M. W. Holladay, in The Organic Chemistry of Drug
Chem. Rev. 102, 1359–1470 (2002). 5567–5568 (1978). Design and Drug Action (Academic Press, ed. 3, 2014), chap. 2.

Hilton et al., Science 362, 799–804 (2018) 16 November 2018 5 of 6


R ES E A RC H | R E S EA R C H A R T I C LE

41. D. A. Erlanson, S. W. Fesik, R. E. Hubbard, W. Jahnke, H. Jhoti, University, and the Extreme Science and Engineering Discovery SUPPLEMENTARY MATERIALS
Nat. Rev. Drug Discov. 15, 605–619 (2016). Environment (XSEDE) through allocations TG-CHE180006 and www.sciencemag.org/content/362/6416/799/suppl/DC1
42. C. W. Murray, D. C. Rees, Nat. Chem. 1, 187–192 (2009). TG-CHE180056. XSEDE is supported by NSF grant ACI-1548562. Materials and Methods
43. T. Cernak, K. D. Dykstra, S. Tyagarajan, P. Vachal, S. W. Krska, Author contributions: M.C.H., X.Z., and B.T.B. performed the Figs. S1 to S26
Chem. Soc. Rev. 45, 546–576 (2016). experimental work; the computational studies were performed by Tables S1 to S22
J.V.A.-R. and R.S.P.; all authors contributed to the design of NMR Spectra
ACKN OW LEDG MEN TS the experimental and computational work and to data analysis, References (44–97)
Funding: Supported by startup funds from Colorado State discussed the results, and commented on the manuscript; and Movies S1 and S2
University and by NIH award R01 GM124094-01 (A.M., M.C.H., A.M. and R.S.P. wrote the manuscript. Competing interests: The
X.Z., and B.T.B.). We acknowledge the RMACC Summit authors declare no competing interests; Data and materials 8 January 2018; resubmitted 3 July 2018
supercomputer, supported by NSF grants ACI-1532235 and availability: All data are available in the main text or the Accepted 25 September 2018
ACI-1532236, the University of Colorado Boulder, Colorado State supplementary materials. 10.1126/science.aas8961

Downloaded from http://science.sciencemag.org/ on November 15, 2018

Hilton et al., Science 362, 799–804 (2018) 16 November 2018 6 of 6


Heterobiaryl synthesis by contractive C−C coupling via P(V) intermediates
Michael C. Hilton, Xuan Zhang, Benjamin T. Boyle, Juan V. Alegre-Requena, Robert S. Paton and Andrew McNally

Science 362 (6416), 799-804.


DOI: 10.1126/science.aas8961

Heterocycles meet and marry on phosphorus


Metals such as palladium are routinely used to link together carbon rings in pharmaceutical synthesis. However,
the presence of nitrogen in both rings can trip up this process. Hilton et al. report a versatile alternative process in which
phosphorus takes the place of the metal. The phosphorus binds successively to both rings at the sites opposite the

Downloaded from http://science.sciencemag.org/ on November 15, 2018


nitrogen, and treatment with acidic ethanol then pushes them off, bound to each other. Theory implicates a
five-coordinate phosphorus intermediate that kinetically favors coupling of the two nitrogen-bearing rings over reactions
of the other all-carbon substituents.
Science, this issue p. 799

ARTICLE TOOLS http://science.sciencemag.org/content/362/6416/799

SUPPLEMENTARY http://science.sciencemag.org/content/suppl/2018/11/14/362.6416.799.DC1
MATERIALS

REFERENCES This article cites 81 articles, 0 of which you can access for free
http://science.sciencemag.org/content/362/6416/799#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.

Вам также может понравиться