You are on page 1of 8

Article

Cite This: J. Am. Chem. Soc. 2017, 139, 15522-15529 pubs.acs.org/JACS

Factors Controlling the Reactivity and Chemoselectivity of


Resonance Destabilized Amides in Ni-Catalyzed Decarbonylative and
Nondecarbonylative Suzuki-Miyaura Coupling
Chong-Lei Ji and Xin Hong*
Department of Chemistry, Zhejiang University, Hangzhou, 310027, China
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: N-Glutarimide amides have recently emerged as


an exceptional group of compounds with unusually high
reactivity in amide C−N bond activation. To understand the
Downloaded via INDIAN INST OF SCIENCE on June 22, 2022 at 06:16:00 (UTC).

key factors that control the remarkable reactivity of these


resonance destabilized amides, we explored the Ni-catalyzed
decarbonylative and nondecarbonylative Suzuki-Miyaura cou-
pling with N-glutarimide amides through density functional
theory calculations. Two leading effects are responsible for the
C−N cleavage activity of N-glutarimide amides, the coordinat-
ing N-substituents and the geometric twisting. The carbonyl
substituent of the N-glutarimide amides provides crucial
nickel−oxygen interaction, which essentially acts as a directing
group to facilitate the formation of the reactive intermediate
for the amide C−N bond cleavage. The geometric twisting weakens the resonance stability by removing the acyl-nitrogen
conjugation, which lowers the energy penalty for the C−N bond stretch during oxidative addition. For the chemoselectivity of
decarbonylation versus carbonyl retention, we found that the C−C reductive elimination for ketone formation is kinetically faster
than that for biaryl formation, while ketone is thermodynamically less stable with respect to the decarbonylated biaryls. The
computations also suggest that the nickel catalyst is able to promote the decarbonylation of biaryl ketones via an unexpected C−
C bond activation.

■ INTRODUCTION
Ni-catalyzed C−N bond activation of amides has recently
C−N cleavage of amides. Understanding these two fundamen-
tal mechanistic questions with N-glutarimide amides is critical
emerged as a distinctive strategy to utilize amides in organic toward the rational design of amide C−N bond activation and
synthesis.1 A number of exciting transformations, contributed functionalization.
by Garg and co-workers,2 Szostak and co-workers,3,4 Shi and Regarding the mechanistic studies of Ni-catalyzed amide C−
co-workers,5 Zou and co-worker,6 Rueping and co-workers,7 N bond activation, Houk and Garg elucidated the mechanism
Maiti and co-workers,8 Stanley and co-workers,9 and Molander of Ni/NHC-catalyzed esterification of anilides with DFT
and co-workers,10 have been realized based on the Ni-mediated calculations, and the three-centered oxidative addition model
C−N cleavage of amides. Compared to the common planar successfully explains the reactivities of N-substituted aryl
amides, Szostak and co-workers discovered that N-glutarimide amides.2a In addition, the group of Zhao and Zhu,14 as well
amides exhibit significant geometric distortions and unusually as the group of Fu and Yu,15 independently studied the
high reactivities in C−N bond activations,3,4 even in metal-free mechanism of the Ni/NHC-catalyzed Suzuki-Miyaura coupling
conditions.11 The high reactivity of these N-glutarimide amides with Boc-activated amides, which elucidated the roles of K3PO4
are usually rationalized by the twisting geometric nature and water in facilitating the transmetalation step.
(Scheme 1).3a,12 To elucidate the origins of reactivity and chemoselectivity
Despite the vast development of synthetic transformations with N-glutarimide amides, here we report the first computa-
with N-glutarimide amides,3,4 the controlling factors that tional study on the C−N bond activation of this special group
differentiate the C−N bond activation reactivity of N- of amides, focusing on the Ni/PCy3-catalyzed Suzuki-Miyaura
glutarimide amides from those of other amides remain elusive. coupling with aryl boronates. Two essential factors were found
In addition to this key question of reactivity, the Suzuki- to contribute to the exceptional reactivities of N-glutarimide
Miyaura coupling with N-glutarimide amides produced the amides in the Ni-mediated C−N cleavage, the carbonyl
decarbonylated biaryl products,3a which is distinctive compared coordination, and the geometric twisting. The coordinating
with the nondecarbonylative Suzuki-Miyaura couplings involv-
ing other amides (Scheme 1).13 This chemoselectivity is Received: September 5, 2017
another general issue that exists in all transformations involving Published: October 10, 2017

© 2017 American Chemical Society 15522 DOI: 10.1021/jacs.7b09482


J. Am. Chem. Soc. 2017, 139, 15522−15529
Journal of the American Chemical Society Article

Scheme 1. Twisting Nature of N-Glutarimide Amides and reactions with acyl derivatives,2a,14,15,23 the proposed catalytic
Comparisons between the N-Glutarimide Amides and the cycles of the Ni-catalyzed Suzuki-Miyaura coupling between
Boc-Activated Amides in Ni-Catalyzed Suzuki-Miyaura amides and arylboronic acids are shown in Scheme 2. Starting
Coupling2b,3a
Scheme 2. Proposed Catalytic Cycles of Ni-Catalyzed
Suzuki-Miyaura Coupling between Amides and Arylboronic
Acids

glutarimidyl moiety acts as a directing group which facilitates


the generation of the reactive intermediate for C−N bond
cleavage. In addition, the geometric twisting alleviates the
energy penalty for the C−N bond stretch, lowering the intrinsic
barrier for oxidative addition. These synergistic effects also
differentiate the target C−N bond from the adjacent C−N
bonds of the glutarimide substituent, preventing side reactions with the in situ generated Ni(0) active catalyst A, the oxidative
involving undesired C−N bond cleavage. Thus, the inert nature addition breaks the amide C−N bond and generates the
of the glutarimide group is an additional important factor that LnNi(acyl)(amino) intermediate B. B undergoes the trans-
contributes to the high reactivity of the N-glutarimide amides metalation with boronic acid to form the LnNi(acyl)Ar2
toward the designed C−N bond activation. For the chemo- intermediate C, and subsequent decarbonylation generates
selectivity between ketone and decarbonylated biaryl, the the LnNiAr1Ar2 intermediate D. The C−C reductive elimi-
formation of ketone product is kinetically favorable due to the nation of D produces the biaryl product and regenerates the
faster C−C reductive elimination, while ketone is thermody- nickel(0) active catalyst A. Alternatively, C can directly undergo
namically less stable than the decarbonylated biaryl products. the C(acyl)−C(aryl) reductive elimination to produce the
These mechanistic insights will shed light on the future design ketone product.
of amide C−N bond activations and related synthetic We first studied the proposed catalytic cycle using the
applications.


experimental substrates, 1-benzoylpiperidine-2,6-dione and 2-
naphthaleneboronic acid. The free energy changes of the most
COMPUTATIONAL METHODS
favorable pathway leading to the biaryl product are shown in
All density functional theory (DFT) calculations were performed by Figure 1. From the most stable nickel−amide complex 1, an
Gaussian 09 program.16 The geometry optimizations were conducted
isomerization occurs to generate the preoxidative addition
using the B3LYP functional,17 with LANL2DZ basis set18 for nickel
and 6-31G(d) basis set for the other atoms. To confirm whether each intermediate 2. Subsequent oxidative addition via a five-
optimized stationary point is an energy minimum or a transition state, centered transition-state TS3 requires a barrier of 9.6 kcal·
as well as evaluate the zero-point vibrational energy and thermal mol−1 as compared to the intermediate 1, leading to the
corrections at 298 K, the vibrational frequencies were computed at the PCy3Ni(acyl)(amino) intermediate 4. This exceptionally low
same level of theory as for the geometry optimizations. On the basis of barrier for C−N cleavage corroborates the high reactivity of the
the gas-phase optimized structures, the single-point energies and N-glutarimide amides. Notably, the N-glutarimide amides have
solvent effects were evaluated with the M06 functional,19 SDD basis very low nN−π*(acyl) resonance, which correlates with the
set20 for nickel, and 6-311+G(d,p) basis set for the other atoms. The amide reactivity in cross-coupling reactions according to
solvation energies were calculated using the self-consistent reaction
Szostak’s recent study12c and is one of the factors that
field with the CPCM implicit solvent model.21 Fragment distortion
and interaction energies were calculated with the M06 functional, SDD facilitates the oxidative addition via C−N cleavage (vide infra).
basis set for nickel, and 6-311+G(d,p) basis set for the other atoms, From 4, the coordination with sodium carbonate is quite
without the inclusion of solvation energy corrections. The details of exergonic, and the signals responsible for the nickel−carbonate
distortion/interaction analysis are provided in the Supporting complex without sodium cations are detected in the
Information. The 3D diagrams of computed species were generated stoichiometric electrospray ionization mass spectrometry
using CYLView.22 experiments by Szostak and co-worker.3a 5 further reacts with

■ RESULTS AND DISCUSSION


Proposed Catalytic Cycle. On the basis of previous
2-naphthaleneboronic acid to generate the stable intermediate
6, and the sodium carbonate in 6 acts as an intramolecular base
to facilitate the transmetalation via TS7. The PCy3Ni(acyl)-
mechanistic studies on the nickel-catalyzed cross-coupling (naphthyl) intermediate 9 then undergoes the decarbonylation
15523 DOI: 10.1021/jacs.7b09482
J. Am. Chem. Soc. 2017, 139, 15522−15529
Journal of the American Chemical Society Article

Figure 1. DFT-computed Gibbs free energy changes of the most favorable pathway of Ni/PCy3-catalyzed decarbonylative Suzuki-Miyaura coupling
between the N-glutarimide amide 14 and 2-naphthaleneboronic acid.

via TS10, and the subsequent C(sp2)−C(sp2) reductive oxidative addition with 14 is significantly more exergonic than
elimination via TS13 eventually produces the biaryl product the other two amides. Both the kinetics and thermodynamics
15 and regenerate intermediate 1 for the next catalytic cycle. highlight the distinctive reactivities of the N-glutarimide amides.
On the basis of the calculated free energy changes of the whole We also compared the C−N bond activation of the N-
catalytic cycle, the resting state is the pretransmetalation glutarimide amide 14 to that of the corresponding N-
intermediate 6, and the rate-limiting step is the C(sp2)−C(sp2) succinimide amide, and this N-succinimide amide has a C−N
reductive elimination with an overall barrier of 27.2 kcal·mol−1 bond activation barrier of 13.0 kcal/mol (Figure S3). These
(6 to TS13). This significant barrier explains the requirement computed barriers of amide C−N bond activation agree well
for high temperature in experiment.3a To further validate this with the observed reactivities of amides in Ni/PCy3-catalyzed
mechanistic model, we calculated the free energy barriers of a Suzuki-Miyaura coupling under the same conditions (Scheme
number of substituted substrates and compared the computed 3).3a
barriers with the competition experiments from Szostak’s The thermodynamics of the oxidative addition depends
work3a (Figure S1). The satisfying consistency between the highly on the coordinating N-substituents of amide. The
computational and experimental results provides additional carbonyl group of 14 provides an additional strong Ni−O
support for the proposed mechanism. Our calculations also interaction in the postoxidative addition intermediate 4, which
showed that the aliphatic N-glutarimide amide has a higher stabilizes this intermediate and makes the corresponding C−N
overall barrier for the same Suzuki-Miyaura coupling (Figure cleavages exergonic. For the Boc-activated amide 16, similar
S2). Ni−O interaction exists in the intermediate 20, leading to the
Origins of the High Reactivity of N-Glutarimide exergonic oxidative addition. While for the anilide 21, no such
Amides in the Ni-Mediated C−N Bond Activation. To Ni−O interaction exists in tricoordinated nickel intermediate
understand the origins of the exceptional reactivity of N- 20, and thus the oxidative addition is endergonic by 5.7 kcal·
glutarimide amides in the Ni-mediated C−N bond cleavage, we mol−1.
compared the N-glutarimide amide 14 to the N-Bn-N-Boc To elucidate the effects on the kinetic barriers of C−N bond
benzamide 16 and the N-Me-N-Ph benzamide 21, which activation, we decomposed the overall barrier to two parts: the
represents three typical types of amides that are generally free energy of isomerization from the most stable nickel−amide
employed in the Ni-catalyzed cross couplings.1 The free energy complex to the preoxidative addition intermediate (e.g., 1 to 2,
changes of the Ni/PCy3-mediated C−N bond activation of the labeled in green, Figure 3), and the intrinsic barrier of oxidative
three amides are included in Figure 2. The overall C−N bond addition from the preoxidative addition intermediate to the C−
activation barrier of 14 is only 9.6 kcal·mol−1, while those of 16 N cleavage transition state (e.g., 2 to TS3, labeled in yellow,
and 21 are more than 6.0 kcal·mol−1 higher. In addition, the Figure 3). These comparisons reveal two crucial factors that
15524 DOI: 10.1021/jacs.7b09482
J. Am. Chem. Soc. 2017, 139, 15522−15529
Journal of the American Chemical Society Article

Figure 2. DFT-computed free energy changes of Ni/PCy3-mediated C−N activation of three typical types of amides. Free energies are in kcal·mol−1.

Scheme 3. Comparisons of the Amide Reactivities in Ni/ contribute to the exceptional C−N bond activation reactivities
PCy3-Catalyzed Suzuki-Miyaura Coupling3a of N-glutarimide amides: the twisting C−N bond and the
coordinating N-substituents.
The comparisons between the N-glutarimide amide 14 and
the Boc-activated amide 16 suggest that the geometric twisting
can dramatically lower the intrinsic barrier of C−N cleavage.
Both 14 and 16 have a similar free energy of isomerization (1.9
and 3.8 kcal·mol−1, Figure 3) because both the corresponding
preoxidative addition intermediates (2 and 18) have similar
stabilizing nickel−oxygen (carbonyl) interactions. These
carbonyl substituents, glutarimidyl group of 14 and Boc
group of 16, essentially act as directing groups to facilitate
the generation of the reactive intermediates 2 and 18. While
both the amides 14 and 16 contain the directing carbonyl
groups, the cleaving C−N bond of N-glutarimide amide 14 is
intrinsically weaker than that of N-Boc amide 16. The
geometric twisting of 14 removes the nN−π*(acyl) resonance
of the amide C−N bond, making it a single C(acyl)−N bond.
This mitigates the energy penalty to distort the corresponding
C−N bond in TS19, leading to the significantly lower intrinsic
C−N cleavage barrier (7.7 kcal·mol−1 of 14 vs 13.1 kcal·mol−1
of 16). The results of distortion/interaction analysis24,25 on
TS3 and TS19 are included in the Supporting Information
(Figures S4 and S5).
In addition, the comparison between 14 and 21 highlights
the importance of coordinating leaving group for the formation
of preoxidative addition intermediates. The glutarimidyl group
of N-glutarimide amides provides the nickel−oxygen inter-
Figure 3. Components of the overall Ni/PCy3-mediated C−N bond action that stabilizes the preoxidative addition intermediate 2.
activation barriers of the three amides. In contrast, such directing effect is not present in the N-methyl-
N-phenylbenzamide 21, and the intermediate 23 has the amide
C−N bond coordinating to nickel. This weak coordination
15525 DOI: 10.1021/jacs.7b09482
J. Am. Chem. Soc. 2017, 139, 15522−15529
Journal of the American Chemical Society Article

Figure 4. DFT-computed structures, free energies, and distortion/interaction analysis of the competing transition states of the Ni/PCy3-mediated
bond activations of the N-glutarimide amide 14. Energies are in kcal·mol−1.

Figure 5. DFT-computed free energy changes of the competing pathways for the ketone formation and biaryl formation from resting state 6.

destabilizes 23 and leads to the high isomerization energy from activation transition states are shown in Figure 4. The cleavage
22 to 23, eventually increasing the overall C−N bond activation of the twisted amide C−N bond via TS3 requires a barrier of
barrier. only 9.6 kcal·mol−1, while the C−N bond of the glutarimidyl
Origins of Chemoselectivity of Bond Activations. We group requires a 17.2 kcal·mol−1 barrier for cleavage via TS26,
next explored the origins of chemoselectivity of the Ni/PCy3- and the barrier of the C(aryl)−C(carbonyl) bond activation via
mediated bond activations of N-glutarimide amide 14. The TS28 is 23.0 kcal·mol−1. Therefore, there is a strong
optimized structures and free energies of the competing bond chemoselectivity toward the cleavage of the desired amide
15526 DOI: 10.1021/jacs.7b09482
J. Am. Chem. Soc. 2017, 139, 15522−15529
Journal of the American Chemical Society Article

C−N bond, which is consistent with the experimental Scheme 4. (a) Experimental Results That Support the
observations that only this amide C−N bond was cleaved.3a Formation of Ketone Product in Ni-Catalyzed Cross
In addition, the high cleavage barriers for the C−C and C−N Coupling with N-Glutarimide Amides;3b (b) Ni-Mediated
bonds of the glutarimide group (via TS26 and TS28) highlight Decarbonylation of Aryl Ketone27
the inert nature of this substituent, and the stability of N-
substituents is critical in achieving the high reactivity of the
desired C−N bond activation.
To understand the origins of this chemoselectivity, the
distortion/interaction analysis was applied on the competing
transition states (Figure 4). Comparing the two C−N bond
activation transition states, TS3 and TS26, the difference of
ΔEint is the leading cause for the chemoselectivity. ΔEint(TS3)
is −44.6 kcal·mol−1 while the ΔEint(TS26) is −18.6 kcal·mol−1.
This is due to the lack of nickel−oxygen interaction in TS26,
which again highlights the importance of the directing effect of
the glutarimidyl group. Comparing the C−N and C−C bond
activation transition states, TS3 and TS28, both the ΔEdist and
ΔEint disfavor the C−C bond activation transition-state TS28.
The ΔEint favors TS3 because of the favorable nickel−oxygen predictions are validated by Chatani’s recent breakthroughs on
interaction in this transition state. In addition, the C−N bond the Ni-catalyzed decarbonylation of biaryl ketones27 (Scheme
of N-glutarimide amide 14 is essentially a single C(acyl)−N 4b). A Ni-mediated C(aryl)−C(acyl) bond activation
bond due to the geometric twisting,26 which is much weaker as occurred,28 and subsequent decarbonylation and reductive
compared to the C(aryl)−C(carbonyl) bond. The stretch of elimination produced the decarbonylated biaryl products. The
the strong C−C bond leads to the large ΔEdist for TS28, additional calculations with the experimental NHC ligand,
contributing to the high barrier of corresponding C−C bond IMesMe, also showed a surmountable barrier for the decarbon-
activation. ylation of phenyl naphthyl ketone (27.3 kcal/mol, Figure S6),
Origins of the Chemoselectivity of Decarbonylation which provides direct comparisons to the reported exper-
versus Carbonyl Retention. To understand the control of imental results.27
the competition between decarbonylation and carbonyl In addition, Szostak’s experimental studies showed that the
retention, we studied the free energy changes of the formations Ni/PCy3-catalyzed decarbonylative Suzuki-Miyaura coupling
of ketone and biaryl products from the resting state 6 (Figure with N-glutarimide amides can tolerate acetyl group.3a These
5). The formation of the biaryl product (blue pathway) has results suggest that the alkyl aryl ketones are significantly less
been discussed above, and the reductive elimination via TS13 is reactive toward the proposed Ni-catalyzed decarbonylation as
the rate-limiting step of this pathway. Alternatively, the compared to the biaryl ketones. To further validate the
PCy3Ni(acyl)(aryl) intermediate 9 can undergo a direct hypothesized in situ decarbonylation, we studied the Ni/PCy3-
C(aryl)−C(acyl) reductive elimination via TS30 to generate mediated decarbonylation of methyl naphthyl ketone, and
the ketone product 31 (red pathway). The formation of ketone compared the results with those of phenyl naphthyl ketone
is kinetically favorable because of the faster reductive (Scheme 5). The overall barrier with phenyl naphthyl ketone is
elimination (TS30 vs TS13), while the ketone product is 5.8 kcal·mol−1 lower than that with methyl naphthyl ketone
thermodynamically less stable as compared to the biaryl (26.6 vs 32.4 kcal·mol−1), which is consistent with the fact that
product (31 vs 15). In addition, it is noteworthy that the acetyl group is tolerated in this reaction.3a
ketone product 31 is less stable than the resting state of the The increment of overall decarbonylation barrier with methyl
catalytic cycle 6. Therefore, lowering the reaction temperature naphthyl ketone is mainly contributed by the substituent effects
cannot allow the substantial formation of ketone product on the C−C bond activation and reductive elimination. The
because of the equilibrium between 31 and 6, which is C−C bond activation is 13.8 kcal·mol−1 endergonic with phenyl
consistent with the experimental results.3a Only when the naphthyl ketone (31 to 9), while that with methyl naphthyl
reaction condition is sufficient enough to overcome the high ketone is endergonic by 17.1 kcal·mol−1 (32 to 34). This is due
kinetic barrier for the biaryl formation (27.2 kcal·mol−1, 6 to to the change of dNi-π*acyl interaction in the LNi(acyl)(aryl)
TS13), the irreversible C(aryl)−C(aryl) does reductive intermediates (9 and 34); the benzoyl group of 9 is a stronger π
elimination occur to produce the observed biaryl product. acceptor than the acetyl group of 34, which stabilizes the
This mechanistic rationale is supported by several related intermediate 9 through the favorable dNi-π*acyl interaction.23c In
experimental observations.3b,27 Since ketone is the kinetic addition, the reaction with methyl naphthyl ketone requires the
product, if a Ni-catalyzed cross coupling with N-glutarimide C(sp3)−C(sp2) reductive elimination via TS37, which is
amides can occur under mild conditions, the ketone formation intrinsically more difficult compared with the C(sp2)−C(sp2)
should be observed instead of the biaryl formation. Indeed, reductive elimination via TS13.29 Therefore, the decarbon-
Szostak and co-workers showed that the Ni-catalyzed Negishi ylation reactivity of alkyl aryl ketones is significantly lower than
that of biaryl ketones.


coupling with the same N-glutarimide amides can occur under
mild conditions, and this reaction produces the ketone
products3b (Scheme 4a). More importantly, our computations CONCLUSIONS
essentially reveal a background reaction in which the nickel In summary, the reactivities and selectivities of N-glutarimide
catalyst promotes the decarbonylation of biaryl ketones to amides in Ni-catalyzed Suzuki-Miyaura coupling have been
generate biaryls (from 31 to 15) through an unexpected elucidated with DFT calculations. Two critical factors lead to
C(aryl)−C(acyl) bond activation of ketone (via TS30). These the unusually high reactivity of N-glutarimide amides in Ni-
15527 DOI: 10.1021/jacs.7b09482
J. Am. Chem. Soc. 2017, 139, 15522−15529
Journal of the American Chemical Society Article

Scheme 5. DFT-Computed Free Energy Changes of the Ni/ transformation was confirmed by Chatani’s recent experimental
PCy3-Mediated Decarbonylation of Phenyl Naphthyl Ketone studies.27
(a) and Methyl Naphthyl Ketone (b)

*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/jacs.7b09482.
Additional computational results; coordinates and
energies of DFT-computed stationary points (PDF)

■ AUTHOR INFORMATION
Corresponding Author
*hxchem@zju.edu.cn (X.H.)
ORCID
Xin Hong: 0000-0003-4717-2814
Notes
The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
Financial support from Zhejiang University, the Chinese
“Thousand Youth Talents Plan”, and the “Fundamental
Research Funds for the Central Universities” is gratefully
acknowledged. Calculations were performed on the National
Supercomputing Center in Shenzhen and the high-performance
computing system at the Department of Chemistry, Zhejiang
University.

■ REFERENCES
(1) For reviews, see: (a) Meng, G.; Szostak, M. Org. Biomol. Chem.
2016, 14, 5690. (b) Meng, G.; Shi, S.; Szostak, M. Synlett 2016, 27,
2530. (c) Liu, C.; Szostak, M. Chem. - Eur. J. 2017, 23, 7157.
(d) Dander, J. E.; Garg, N. K. ACS Catal. 2017, 7, 1413. (e) Gao, Y.; Ji,
C.-L.; Hong, X. Sci. China: Chem. 2017, DOI: 10.1007/s11426-017-
9025-1.
(2) (a) Hie, L.; Nathel, N. F. F.; Shah, T. K.; Baker, E. L.; Hong, X.;
Yang, Y.-F.; Liu, P.; Houk, K. N.; Garg, N. K. Nature 2015, 524, 79.
(b) Weires, N. A.; Baker, E. L.; Garg, N. K. Nat. Chem. 2016, 8, 75.
(c) Baker, E. L.; Yamano, M. M.; Zhou, Y.; Anthony, S. M.; Garg, N.
K. Nat. Commun. 2016, 7, 11554. (d) Hie, L.; Baker, E. L.; Anthony, S.
M.; Desrosiers, J.-N.; Senanayake, C.; Garg, N. K. Angew. Chem., Int.
Ed. 2016, 55, 15129. (e) Simmons, B. J.; Weires, N. A.; Dander, J. E.;
mediated C−N bond activation, the coordinating N-substitu- Garg, N. K. ACS Catal. 2016, 6, 3176. (f) Simmons, B. J.; Hoffmann,
ents, and the geometric twisting. The coordinating glutarimidyl M.; Hwang, J.; Jackl, M. K.; Garg, N. K. Org. Lett. 2017, 19, 1910.
substituent acts as a directing group, which facilitates the (g) Dander, J. E.; Baker, E. L.; Garg, N. K. Chem. Sci. 2017, 8, 6433.
formation of the reactive intermediate for the C−N bond (3) For nickel-catalyzed transformations involving N-glutarimide
cleavage. This nickel−oxygen(carbonyl) interaction also allows amides, see: (a) Shi, S.; Meng, G.; Szostak, M. Angew. Chem., Int. Ed.
2016, 55, 6959. (b) Shi, S.; Szostak, M. Chem. - Eur. J. 2016, 22,
the formation of the tetra-coordinated nickel intermediate after 10420. (c) Shi, S.; Szostak, M. Org. Lett. 2016, 18, 5872. (d) Shi, S.;
the oxidative addition, resulting in the exergonic C−N bond Szostak, M. Synthesis 2017, 49, 3602.
cleavage. In addition, the geometric twisting removes the planar (4) For other related transformations involving N-glutarimide
acyl−nitrogen conjugation of amide, and the twisting C−N amides, see: (a) Meng, G.; Szostak, M. Angew. Chem., Int. Ed. 2015,
bond of N-glutarimide amides is essentially a single N−C(acyl) 54, 14518. (b) Meng, G.; Szostak, M. Org. Lett. 2015, 17, 4364.
bond.26 This lowers the energy required for the C−N bond (c) Hu, F.; Lalancette, R.; Szostak, M. Angew. Chem., Int. Ed. 2016, 55,
stretch during the oxidative addition; thus, the intrinsic barrier 5062. (d) Liu, C.; Meng, G.; Szostak, M. J. Org. Chem. 2016, 81,
of C−N bond cleavage with N-glutarimide amides is low. 12023. (e) Meng, G.; Szostak, M. Org. Lett. 2016, 18, 796. (f) Liu, C.;
For the chemoselectivity of decarbonylation, a thermody- Achtenhagen, M.; Szostak, M. Org. Lett. 2016, 18, 2375. (g) Liu, C.;
namic versus kinetic control of product formation was Meng, G.; Liu, Y.; Liu, R.; Lalancette, R.; Szostak, R.; Szostak, M. Org.
Lett. 2016, 18, 4194. (h) Meng, G.; Shi, S.; Szostak, M. ACS Catal.
discovered. The biaryl ketones have a lower barrier of 2016, 6, 7335. (i) Liu, C.; Liu, Y.; Liu, R.; Lalancette, R.; Szostak, R.;
formation than the decarbonylated biaryl products, while the Szostak, M. Org. Lett. 2017, 19, 1434. (j) Liu, Y.; Shi, S.; Achtenhagen,
formation of biaryl ketones is thermodynamically less favorable. M.; Liu, R.; Szostak, M. Org. Lett. 2017, 19, 1614. (k) Meng, G.; Lei,
The calculations suggest that the nickel catalyst is able to P.; Szostak, M. Org. Lett. 2017, 19, 2158. (l) Lei, P.; Meng, G.; Shi, S.;
promote the decarbonylation of biaryl ketones through a Ling, Y.; An, J.; Szostak, R.; Szostak, M. Chem. Sci. 2017, 8, 6525.
unexpected C−C bond activation pathway. This proposed (m) Liu, C.; Szostak, M. Angew. Chem., Int. Ed. 2017, 56, 12718.

15528 DOI: 10.1021/jacs.7b09482


J. Am. Chem. Soc. 2017, 139, 15522−15529
Journal of the American Chemical Society Article

(5) (a) Hu, J.; Zhao, Y.; Liu, J.; Zhang, Y.; Shi, Z. Angew. Chem., Int. (23) (a) Li, Z.; Zhang, S.-L.; Fu, Y.; Guo, Q.-X.; Liu, L. J. Am. Chem.
Ed. 2016, 55, 8718. (b) Hu, J.; Wang, M.; Pu, X.; Shi, Z. Nat. Commun. Soc. 2009, 131, 8815. (b) Yu, H.; Fu, Y. Chem. - Eur. J. 2012, 18,
2017, 8, 14993. 16765. (c) Hong, X.; Liang, Y.; Houk, K. N. J. Am. Chem. Soc. 2014,
(6) Li, X.; Zou, G. Chem. Commun. 2015, 51, 5089. 136, 2017. (d) Lu, Q.; Yu, H.; Fu, Y. J. Am. Chem. Soc. 2014, 136,
(7) (a) Yue, H.; Guo, L.; Lee, S.-C.; Liu, X.; Rueping, M. Angew. 8252. (e) Xu, H.; Muto, K.; Yamaguchi, J.; Zhao, C.; Itami, K.;
Chem., Int. Ed. 2017, 56, 3972. (b) Yue, H.; Guo, L.; Liao, H.-H.; Cai, Musaev, D. G. J. Am. Chem. Soc. 2014, 136, 14834. (f) Muto, K.;
Y.; Zhu, C.; Rueping, M. Angew. Chem., Int. Ed. 2017, 56, 4282. Yamaguchi, J.; Musaev, D. G.; Itami, K. Nat. Commun. 2015, 6, 7508.
(c) Srimontree, W.; Chatupheeraphat, A.; Liao, H.-H.; Rueping, M. (g) Li, Z.; Liu, L. Chin. J. Catal. 2015, 36, 3.
Org. Lett. 2017, 19, 3091. (24) For reviews of distortion/interaction analysis, see: (a) van
(8) Dey, A.; Sasmal, S.; Seth, K.; Lahiri, G. K.; Maiti, D. ACS Catal. Leeuwen, P. W. N. M.; Kamer, P. C. J.; Reek, J. N. H.; Dierkes, P.
2017, 7, 433. Chem. Rev. 2000, 100, 2741. (b) Van Zeist, W.-J.; Bickelhaupt, F. M.
(9) Walker, J. A., Jr; Vickerman, K. L.; Humke, J. N.; Stanley, L. M. J. Org. Biomol. Chem. 2010, 8, 3118. (c) Fernandez, I.; Bickelhaupt, F. M.
Am. Chem. Soc. 2017, 139, 10228. Chem. Soc. Rev. 2014, 43, 4953. (d) Bickelhaupt, F. M.; Houk, K. N.
(10) Amani, J.; Alam, R.; Badir, S.; Molander, G. A. Org. Lett. 2017, Angew. Chem., Int. Ed. 2017, 56, 10070.
19, 2426. (25) For selected examples of distortion/interaction analysis in
(11) The enolization of the glutarimide group is proposed to be metal-mediated C−X bond activation: (a) Legault, C. Y.; Garcia, Y.;
involved in the metal-free Friedel−Crafts acylation of the N- Merlic, C. A.; Houk, K. N. J. Am. Chem. Soc. 2007, 129, 12664. (b) de
glutarimide amides, see: Liu, Y.; Meng, G.; Liu, R.; Szostak, M. Jong, G. T.; Bickelhaupt, F. M. ChemPhysChem 2007, 8, 1170. (c) van
Chem. Commun. 2016, 52, 6841. Zeist, W.-J.; Visser, R.; Bickelhaupt, F. M. Chem. - Eur. J. 2009, 15,
(12) For additional studies on amide bond properties that are 6112. (d) Shang, R.; Yang, Z.-W.; Wang, Y.; Zhang, S.-L.; Liu, L. J. Am.
relevant to the cross couplings, see: (a) Pace, V.; Holzer, W.; Meng, Chem. Soc. 2010, 132, 14391. (e) Yang, Y.-F.; Cheng, G.-J.; Liu, P.;
G.; Shi, S.; Lalancette, R.; Szostak, R.; Szostak, M. Chem. - Eur. J. 2016, Leow, D.; Sun, T.-Y.; Chen, P.; Zhang, X.; Yu, J.-Q.; Wu, Y.-D.; Houk,
22, 14494. (b) Szostak, R.; Shi, S.; Meng, G.; Lalancette, R.; Szostak, K. N. J. Am. Chem. Soc. 2014, 136, 344. (f) Cheng, G.-J.; Yang, Y.-F.;
M. J. Org. Chem. 2016, 81, 8091. (c) Szostak, R.; Meng, G.; Szostak, Liu, P.; Chen, P.; Sun, T.-Y.; Li, G.; Zhang, X.; Houk, K. N.; Yu, J.-Q.;
M. J. Org. Chem. 2017, 82, 6373. Wu, Y.-D. J. Am. Chem. Soc. 2014, 136, 894. (g) Green, A. G.; Liu, P.;
(13) For nickel-catalyzed nondecarbonylative Suzuki-Miyaura-type Merlic, C. A.; Houk, K. N. J. Am. Chem. Soc. 2014, 136, 4575.
coupling with amides, see refs 2b, 6, and 9. For palladium-catalyzed (h) Cannon, J. S.; Zou, L.; Liu, P.; Lan, Y.; O’Leary, D. J.; Houk, K. N.;
nondecarbonylative cross couplings with amides, see: (a) Lei, P.; Grubbs, R. H. J. Am. Chem. Soc. 2014, 136, 6733. (i) Hong, X.; Wang,
Meng, G.; Szostak, M. ACS Catal. 2017, 7, 1960. (b) Lei, P.; Meng, G.; J.; Yang, Y.-F.; He, L.; Ho, C.-Y.; Houk, K. N. ACS Catal. 2015, 5,
Ling, Y.; An, J.; Szostak, M. J. Org. Chem. 2017, 82, 6638. (c) Meng, 5545. (j) Wolters, L. P.; Koekkoek, R.; Bickelhaupt, F. M. ACS Catal.
G.; Szostak, R.; Szostak, M. Org. Lett. 2017, 19, 3596. (d) Meng, G.; 2015, 5, 5766.
Lalancette, R.; Szostak, R.; Szostak, M. Org. Lett. 2017, 19, 4656 and (26) Szostak, M.; Aubé, J. Chem. Rev. 2013, 113, 5701.
refs 4b, g−i, k, and l. (27) Morioka, T.; Nishizawa, A.; Furukawa, T.; Tobisu, M.; Chatani,
(14) Liu, L.; Chen, P.; Sun, Y.; Wu, Y.; Chen, S.; Zhu, J.; Zhao, Y.-F. N. J. Am. Chem. Soc. 2017, 139, 1416.
J. Org. Chem. 2016, 81, 11686. (28) For a recent review on transition-metal-mediated C−C bond
(15) Xu, Z.; Yu, H.-Z.; Fu, Y. Chem. - Asian J. 2017, 12, 1765. activation, see: Fumagalli, G.; Stanton, S.; Bower, J. F. Chem. Rev.
(16) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; 2017, 117, 9404 , and references therein..
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, (29) (a) Mann, G.; Baranano, D.; Hartwig, J. F.; Rheingold, A. L.;
B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. Guzei, I. A. J. Am. Chem. Soc. 1998, 120, 9205. (b) Cohen, R.; Milstein,
P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; D.; Martin, J. M. L. Organometallics 2004, 23, 2336. (c) Culkin, D. A.;
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, Hartwig, J. F. Organometallics 2004, 23, 3398. (d) Pérez-Rodríguez,
T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; M.; Braga, A. A. C.; Garcia-Melchor, M.; Pérez-Temprano, M. H.;
Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, Casares, J. A.; Ujaque, G.; de Lera, A. R.; Á lvarez, R.; Maseras, F.;
K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Espinet, P. J. Am. Chem. Soc. 2009, 131, 3650. (e) Racowski, J. M.;
Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, Dick, A. R.; Sanford, M. S. J. Am. Chem. Soc. 2009, 131, 10974.
N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; (f) Pérez-Rodríguez, M.; Braga, A. A. C.; de Lera, A. R.; Maseras, F.;
Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Á lvarez, R.; Espinet, P. Organometallics 2010, 29, 4983.
Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.;
Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.;
Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.;
Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.; Gaussian 09,
revision C.01; Gaussian Inc.: Wallingford, CT, 2010.
(17) (a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (b) Lee, C.;
Yang, W.; Parr, R. G. Phys. Rev. B: Condens. Matter Mater. Phys. 1988,
37, 785.
(18) (a) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270.
(b) Wadt, W. R.; Hay, P. J. J. Chem. Phys. 1985, 82, 284. (c) Hay, P. J.;
Wadt, W. R. J. Chem. Phys. 1985, 82, 299.
(19) (a) Zhao, Y.; Truhlar, D. G. Theor. Chem. Acc. 2008, 120, 215.
(b) Zhao, Y.; Truhlar, D. G. Acc. Chem. Res. 2008, 41, 157.
(20) (a) von Szentpaly, L.; Fuentealba, P.; Preuss, H.; Stoll, H. Chem.
Phys. Lett. 1982, 93, 555. (b) Dolg, M.; Wedig, U.; Stoll, H.; Preuss, H.
J. Chem. Phys. 1987, 86, 866. (c) Schwerdtfeger, P.; Dolg, M.; Schwarz,
W. H. E.; Bowmaker, G. A.; Boyd, P. D. W. J. Chem. Phys. 1989, 91,
1762.
(21) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. J. Comput. Chem.
2003, 24, 669.
(22) CYLview, 1.0b; Legault, C. Y. Université de Sherbrooke, 2009
(http://www.cylview.org).

15529 DOI: 10.1021/jacs.7b09482


J. Am. Chem. Soc. 2017, 139, 15522−15529

You might also like