ArticlePDF Available

The Origin of Stereoselectivity in the Hydrogenation of Oximes Catalyzed by Iridium Complexes: A DFT Mechanistic Study

Authors:

Abstract and Figures

Herein the reaction mechanism and the origin of stereoselectivity of asymmetric hydrogenation of oximes to hydroxylamines catalyzed by the cyclometalated iridium (III) complexes with chiral substituted single cyclopentadienyl ligands (Ir catalysts A1 and B1) under acidic condition were unveiled using DFT calculations. The catalytic cycle for this reaction consists of the dihydrogen activation step and the hydride transfer step. The calculated results indicate that the hydride transfer step is the chirality-determining step and the involvement of methanesulfonate anion (MsO−) in this reaction is of importance in the asymmetric hydrogenation of oximes catalyzed by A1 and B1. The calculated energy barriers for the hydride transfer steps without an MsO− anion are higher than those with an MsO− anion. The differences in Gibbs free energies between TSA5−1fR/TSA5−1fS and TSB5−1fR/TSB5−1fS are 13.8/13.2 (ΔΔG‡ = 0.6 kcal/mol) and 7.5/5.6 (ΔΔG‡ = 1.9 kcal/mol) kcal/mol for the hydride transfer step of substrate protonated oximes with E configuration (E−2a−H+) with MsO− anion to chiral hydroxylamines product R−3a/S−3a catalyzed by A1 and B1, respectively. According to the Curtin–Hammet principle, the major products are hydroxylamines S−3a for the reaction catalyzed by A1 and B1, which agrees well with the experimental results. This is due to the non-covalent interactions among the protonated substrate, MsO− anion and catalytic species. The hydrogen bond could not only stabilize the catalytic species, but also change the preference of stereoselectivity of this reaction.
Content may be subject to copyright.
Citation: Ali, Q.; Chen, Y.; Zhang, R.;
Li, Z.; Tang, Y.; Pu, M.; Lei, M. The
Origin of Stereoselectivity in the
Hydrogenation of Oximes Catalyzed
by Iridium Complexes: A DFT
Mechanistic Study. Molecules 2022,27,
8349. https://doi.org/10.3390/
molecules27238349
Academic Editors: Daoshan Yang
and Zhanhui Yang
Received: 27 September 2022
Accepted: 25 November 2022
Published: 30 November 2022
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
molecules
Article
The Origin of Stereoselectivity in the Hydrogenation of Oximes
Catalyzed by Iridium Complexes: A DFT Mechanistic Study
Qaim Ali 1, Yongyong Chen 1, Ruixue Zhang 1, Zhewei Li 1, Yanhui Tang 1,2, Min Pu 1and Ming Lei 1, *
1State Key Laboratory of Chemical Resource Engineering, Institute of Computational Chemistry,
College of Chemistry, Beijing University of Chemical Technology, Beijing 100029, China
2School of Materials Design and Engineering, Beijing Institute of Fashion Technology, Beijing 100029, China
*Correspondence: leim@mail.buct.edu.cn; Tel.: +86-10-6444-6598
Abstract:
Herein the reaction mechanism and the origin of stereoselectivity of asymmetric hydro-
genation of oximes to hydroxylamines catalyzed by the cyclometalated iridium (III) complexes
with chiral substituted single cyclopentadienyl ligands (Ir catalysts
A1
and
B1
) under acidic con-
dition were unveiled using DFT calculations. The catalytic cycle for this reaction consists of the
dihydrogen activation step and the hydride transfer step. The calculated results indicate that the
hydride transfer step is the chirality-determining step and the involvement of methanesulfonate
anion (MsO
) in this reaction is of importance in the asymmetric hydrogenation of oximes cat-
alyzed by
A1
and
B1
. The calculated energy barriers for the hydride transfer steps without an
MsO
anion are higher than those with an MsO
anion. The differences in Gibbs free energies be-
tween
TSA51fR
/
TSA51fS
and
TSB51fR
/
TSB51fS
are 13.8/13.2 (
∆∆
G
= 0.6 kcal/mol) and
7.5/5.6 (
∆∆
G
= 1.9 kcal/mol) kcal/mol for the hydride transfer step of substrate protonated oximes
with Econfiguration (
E2aH+
) with MsO
anion to chiral hydroxylamines product
R3a
/
S3a
catalyzed by
A1
and
B1
, respectively. According to the Curtin–Hammet principle, the major products
are hydroxylamines
S3a
for the reaction catalyzed by
A1
and
B1
, which agrees well with the
experimental results. This is due to the non-covalent interactions among the protonated substrate,
MsO
anion and catalytic species. The hydrogen bond could not only stabilize the catalytic species,
but also change the preference of stereoselectivity of this reaction.
Keywords: DFT; non-covalent interactions; stereoselectivity; asymmetric hydrogenation; oxime
1. Introduction
The chiral hydroxylamines are one of the important organic intermediates in pharma-
ceutical and agricultural industries, which could be attributed to the reactive N–O bond
in the structural motif. The molecules owning the N-alkoxy amine group are very com-
mon in a wide range of bioactive and pharmaceutical products [
1
]. Among the syntheses
of chiral amines, the asymmetric hydrogenation of the C=N double bond catalyzed by
transition-metal complexes is one of the most efficient methods to prepare enantiomers at
an industrial scale [
2
]. However, the selective reduction of oximes to the corresponding
chiral hydroxylamine derivatives remains a challenge because the undesired reductive
cleavage of the labile N–O bond leads to undesired primary amines (see Scheme 1a) [
3
].
Chiral hydroxylamine could be obtained by partial oxidation of chiral amines, but this
method could either involve multiple steps or be prone to overreaction [
4
7
]. In 2014,
Oestreich et al. reported the reductive hydrogenation of oxime ether catalyzed by B(C
6
F
5
)
3
at room temperature [
8
]. Recently, Zhang et al. realized Ni-catalyzed asymmetric hydro-
genation of oxides to chiral hydroxyl amines via weak attractive interactions between
the catalyst and substrate (Scheme 1b) [
9
]. In 2020, Cramer et al. reported this reaction
catalyzed by cyclometallated iridium (III) complexes with chiral substituted single cy-
clopentadienyl ligands and C,N-bidentate aryl imine ligand (see Scheme 1c) [
10
]. Despite
Molecules 2022,27, 8349. https://doi.org/10.3390/molecules27238349 https://www.mdpi.com/journal/molecules
Molecules 2022,27, 8349 2 of 9
the experimental reports above, the effective strategy of asymmetric hydrogenation of
oxime to hydroxylamine is still relatively scarce. The reaction by Cramer et al. proceeded
at room temperature with 98% ee and the turnover number (TON) reached 4000 before
becoming inactive. This transformation was comprised of the quantitative protonation of
the substrate and subsequent hydride transfer. The hydride could be transferred to the
substrate either directly or with the help of the anion mediation by means of non-covalent
contacts with the transition-metal catalytic species and the protonated substrates. Although
it is obvious that the cyclometallated iridium complexes such as
A1
and
B1
owning chiral
cyclopentadienyl ligand with binaphthyl as backbone and an achiral C,N-bidentate aryl
imine ligand are of importance in the asymmetric hydrogenation of oximes, the reaction
mechanism and the origin of stereoselectivity for this reaction are still unclear. Therefore,
herein a DFT mechanistic study was performed in order to unveil the nature of this reaction
catalyzed by catalysts A1 and B1.
Molecules 2022, 27, x FOR PEER REVIEW 2 of 9
at room temperature [8]. Recently, Zhang et al. realized Ni-catalyzed asymmetric hydro-
genation of oxides to chiral hydroxyl amines via weak attractive interactions between the
catalyst and substrate (Scheme 1b) [9]. In 2020, Cramer et al. reported this reaction cata-
lyzed by cyclometallated iridium (III) complexes with chiral substituted single cyclopen-
tadienyl ligands and C,N-bidentate aryl imine ligand (see Scheme 1c) [10]. Despite the
experimental reports above, the effective strategy of asymmetric hydrogenation of oxime
to hydroxylamine is still relatively scarce. The reaction by Cramer et al. proceeded at room
temperature with 98% ee and the turnover number (TON) reached 4000 before becoming
inactive. This transformation was comprised of the quantitative protonation of the sub-
strate and subsequent hydride transfer. The hydride could be transferred to the substrate
either directly or with the help of the anion mediation by means of non-covalent contacts
with the transition-metal catalytic species and the protonated substrates. Although it is
obvious that the cyclometallated iridium complexes such as A1 and B1 owning chiral cy-
clopentadienyl ligand with binaphthyl as backbone and an achiral C,N-bidentate aryl
imine ligand are of importance in the asymmetric hydrogenation of oximes, the reaction
mechanism and the origin of stereoselectivity for this reaction are still unclear. Therefore,
herein a DFT mechanistic study was performed in order to unveil the nature of this reac-
tion catalyzed by catalysts A1 and B1.
S
O
OH3C
O
Ir N
O
O
O
MeO
OMe
S
O
OH3C
O
Ir N
O
MeO
OMe
O
NOH
RH2,AcOH
ClCl
Cl
Me
NOtBu
Ar Me
NOtBuH
E2a H+
Ar Me
HN OtBu
H
Ir, H2
23 oC, 20 h
MsOH
B1
>99 %, 89:11 er
A1
41 %, 70:30 er
(b) The Nickel-catalysed asymmetric hydrogenation of oximes
S3a
E1a
(c) The acid-assisted stereoselective hydrogenation of oximes catalyzed by chiral Ir catalysts
M
NOR3
R2
R1
HN
R2
R1
NH2
R2
R1H
H2
N-O
reduction
H
OR3
(a) The transition-metal-catalyzed hydrogenation of oximes deliver undesired amines
over reduction
lack of reactivity
Ni(OAc)2
PhPBE(L*) HN OH
R
Scheme 1. (a) The transition-metal-catalyzed hydrogenation of oximes deliver undesired amines [3],
(b) the Nickel-catalyzed asymmetric hydrogenation of oximes [9] and (c) the acid-assisted stereose-
lective hydrogenation of oximes catalyzed by chiral Ir catalysts [10].
As shown in Scheme 2, the catalytic cycle of this reaction consists of the dihydrogen
activation (DA) step and the hydride transfer (HT) step. In the dihydrogen activation step,
Scheme 1.
(
a
) The transition-metal-catalyzed hydrogenation of oximes deliver undesired amines [
3
],
(
b
) the Nickel-catalyzed asymmetric hydrogenation of oximes [
9
] and (
c
) the acid-assisted stereoselec-
tive hydrogenation of oximes catalyzed by chiral Ir catalysts [10].
As shown in Scheme 2, the catalytic cycle of this reaction consists of the dihydrogen
activation (DA) step and the hydride transfer (HT) step. In the dihydrogen activation step,
one molecular hydrogen interacts with
1
by replacing the weakly coordinated methanesul-
fonate anion (MsO
) ligand to afford intermediate
2
, then the dihydrogen is heterolytically
split with the assistance of the MsO
anion via the transition state
TS23
to form in-
termediate
3.
Before the addition of oxime substates in the catalytic cycle, the original
oxime substrate
E1a
is protonated by MsOH to achieve protonated adduct
E2aH+
Molecules 2022,27, 8349 3 of 9
without MsO
anion or
E2a
with MsO
anion. In the subsequent hydride transfer step,
the methanesulfonic acid (MsOH) is released to produce intermediate
4
; the protonated
substrates (
E2aH+
or
E2a
) could interact with
4
to achieve the chiral hydroxylamine
product
3a
and regenerate
1
completing the catalytic cycle. In this work, the reaction mech-
anism of the asymmetric hydrogenation of oximes to hydroxylamines was discussed, the
origin of the stereoselectivity was unveiled, and the important role of the MsO
anion in
this reaction was described. This might shed lights on the development of transition-metal
catalysts based on the mechanism of the asymmetric hydrogenation of oximes in the future.
Molecules 2022, 27, x FOR PEER REVIEW 3 of 9
one molecular hydrogen interacts with 1 by replacing the weakly coordinated me-
thanesulfonate anion (MsO) ligand to afford intermediate 2, then the dihydrogen is het-
erolytically split with the assistance of the MsO anion via the transition state TS23 to
form intermediate 3. Before the addition of oxime substates in the catalytic cycle, the orig-
inal oxime substrate E1a is protonated by MsOH to achieve protonated adduct E2aH+
without MsO anion or E2a with MsO anion. In the subsequent hydride transfer step,
the methanesulfonic acid (MsOH) is released to produce intermediate 4; the protonated
substrates (E2aH+ or E2a) could interact with 4 to achieve the chiral hydroxylamine
product 3a and regenerate 1 completing the catalytic cycle. In this work, the reaction
mechanism of the asymmetric hydrogenation of oximes to hydroxylamines was dis-
cussed, the origin of the stereoselectivity was unveiled, and the important role of the MsO
anion in this reaction was described. This might shed lights on the development of tran-
sition-metal catalysts based on the mechanism of the asymmetric hydrogenation of ox-
imes in the future.
H2
Direct hydride
transfer
MsO
ClCl
Cl
NOtBu
Me
Ar
N
OtBu
Me
H
N-protonation
E1a
MsO
Ar
N
OtBu
Me
H
E2a
MsOH
E2a H+
Ir
CN
H
4
MsO
Ir
CN
H
Ar
N
OtBu
Me
H
5f
Ir
CN
H
Ar
N
OtBu
Me
H
5u
Ir
CN
6
IrMsO
CN
1
Ir
CN
H
H
MsO
2
3
Ir
CN
H
H
MsO
3a
Ar
N
OtBu
Me
H
H
TS5 1f
TS2 3
MsOH
3
a
TS5 1u
MsO
Anion assisted
hydride transfer
Dihydrogen activation
Scheme 2. Proposed reaction mechanism for the asymmetric hydrogenation of oximes to hydroxyl-
amines catalyzed by iridium complexes.
2. Results and Discussion
The catalytic cycle of the asymmetric hydrogenation of oxime to hydroxyamine cat-
alyzed by Ir catalysts (A1 and B1) includes the dihydrogen activation step and the hydride
transfer step. In this work, the (S)-Ir catalysts bearing a chiral binaphthyl-derived cyclo-
pentadienyl ligand was used as the catalytic species for this reaction, which was reported
by experiments [10]. Figure 1 presents the Gibbs free energy profiles for the asymmetric
hydrogenation of E-oxime (E1a) catalyzed by A1; the energies of the stationary points
are relative to corresponding starting point A1 (see that of hydrogenation of Z-oxime in
Figure S1 of SM). E-oxime was used because it is more stable than Z-oxime. In the dihy-
drogen activation step, one molecule hydrogen coordinates with Ir center by replacing the
weakly coordinated MsO anion to form intermediate A2, which is endergonic by 14.7
Scheme 2.
Proposed reaction mechanism for the asymmetric hydrogenation of oximes to hydroxy-
lamines catalyzed by iridium complexes.
2. Results and Discussion
The catalytic cycle of the asymmetric hydrogenation of oxime to hydroxyamine cat-
alyzed by Ir catalysts (
A1
and
B1
) includes the dihydrogen activation step and the hydride
transfer step. In this work, the (S)-Ir catalysts bearing a chiral binaphthyl-derived cyclopen-
tadienyl ligand was used as the catalytic species for this reaction, which was reported
by experiments [
10
]. Figure 1presents the Gibbs free energy profiles for the asymmetric
hydrogenation of E-oxime (
E1a
) catalyzed by
A1
; the energies of the stationary points
are relative to corresponding starting point
A1
(see that of hydrogenation of Z-oxime in
Figure S1 of SM). E-oxime was used because it is more stable than Z-oxime. In the dihy-
drogen activation step, one molecule hydrogen coordinates with Ir center by replacing
the weakly coordinated MsO
anion to form intermediate
A2
, which is endergonic by
14.7 kcal/mol. Then the coordinated dihydrogen causes a heterolytic splitting with the
assistance of the MsO
anion. The free energy barrier from
A1
to
A3
is 14.5 kcal/mol via
TSA23
.
A3
forms iridium hydride intermediate
A4
with the release of one molecule of
MsOH, which is exergonic by 4.9 kcal/mol. The following hydride transfer step could
be the direct hydride transfer without the involvement of the MsO
anion or the MsO
-
Molecules 2022,27, 8349 4 of 9
assisted hydride transfer. In the direct hydride transfer step, the substrate tethering with
one molecule MsOH,
E2a
, could produce the protonated oxime
E2aH+
with the
removal of one MsO
anion. Then
E2aH+
approaches the Ir center of
A4
along the
si-face direction to form intermediate
A5uR
leading to the hydroxylamine product
R3a
.
If it approaches the Ir center of
A4
along the re-face direction, the intermediate
A5uS
is
generated leading to
S3a
hydroxylamine. The “u” in the names of stationary points such
as
A5uS
represents stationary points along an unfavorable pathway, the “f” is used for
those along a favorable pathway, and “S” or “R” denotes the reaction pathways leading to
S or R products, respectively. In the direct hydride transfer step from
A5uR
/
A5uS
to
A6
,
the hydride is transferred from an Ir center to the carbon atom of pronated oxime moiety of
A5uR
/
A5uS
directly via
TSA56uR
/
TSA56uS
. The free energy barriers for this step
are 5.2/8.1 kcal/mol from
A5uR
/
A5uS
to
A6uR
/
A6uS
and hydroxylamines, respectively.
Molecules 2022, 27, x FOR PEER REVIEW 4 of 9
kcal/mol. Then the coordinated dihydrogen causes a heterolytic splitting with the assis-
tance of the MsO anion. The free energy barrier from A1 to A3 is 14.5 kcal/mol via
TSA23. A3 forms iridium hydride intermediate A4 with the release of one molecule of
MsOH, which is exergonic by 4.9 kcal/mol. The following hydride transfer step could be
the direct hydride transfer without the involvement of the MsO anion or the MsO-as-
sisted hydride transfer. In the direct hydride transfer step, the substrate tethering with
one molecule MsOH, E2a, could produce the protonated oxime E2aH+ with the re-
moval of one MsO anion. Then E2aH+ approaches the Ir center of A4 along the si-face
direction to form intermediate A5uR leading to the hydroxylamine product R3a. If it
approaches the Ir center of A4 along the re-face direction, the intermediate A5uS is gener-
ated leading to S3a hydroxylamine. The “u” in the names of stationary points such as
A5uS represents stationary points along an unfavorable pathway, the “f” is used for those
along a favorable pathway, and “S” or “R” denotes the reaction pathways leading to S or
R products, respectively. In the direct hydride transfer step from A5uR/A5uS to A6, the
hydride is transferred from an Ir center to the carbon atom of pronated oxime moiety of
A5uR/A5uS directly via TSA56uR/TSA56uS. The free energy barriers for this step are
5.2/8.1 kcal/mol from A5uR/A5uS to A6uR/A6uS and hydroxylamines, respectively.
Ir
0.0
10.0
-10.0
-20.0
20.0
30.0
0.0
14.7 14.5
12.1
7.2
A5uR 21.8
A5uS 21.1
TSA5 6uS 29.2
TSA5 6uR 27.0
9.3
IrMsO
Ir
H
H
MsO
IrH
OMs
H
IrH
OMs
H
IrH
TSA2 3
H2
MsOH
3a
Cp*
C
NH
C
NH
Cp*
Ir
C
NH
C
NH
Dihydrogen activation Hydride transfer
Cp*
Ir
C
N
MsO
E2a
C
N
Me
H
BuO
H
C
N
Me
H
BuO
S
O
O
CH3
O
E2a 3a
A1
A2 A3
A4
Ar
Ar
t
t
A5fR 2.9
A5fS 2.2
TSA5 1fR 13.8
TSA5 1fS 13.2
MsO
Cp*
Ir
C
NH
C
NH
S
O
O
CH3
O
Cp*
Ir
C
NH
C
NH
S
O
O
CH3
O
without anion assistance
with anion assistance
A1
A6
24.3
Figure 1. The Gibbs free energy profiles of asymmetric hydrogenation of oximes to hydroxylamines
catalyzed by Ir complex A1 (unit: kcal/mol). * represents chirality and ‡ means this is a transition
state.
Finally, A6uR/A6uS combine with the MsO anion to regenerate the active catalytic
species A1 and the catalytic cycle is completed. It should be noted that the energetic span
[11] of generating R3a/S3a products are 27.0/29.2 kcal/mol
(A1TSA56uR/TSA56uS), respectively. That means that R3a is the dominant product
instead of S3a, and the energetic span is high, which is contrary with the conditions for
the experiment to be carried out at room temperature.
Therefore, the direct hydride transfer mode cannot explain well the experimental ob-
servation. The non-covalent interactions have been proposed to be of importance in the
asymmetric reactions, which might stabilize the transition states, accelerate chemical re-
action and regulate the stereoselectivity [9,12,13]. Note that the Gibbs free energy of the
oxime substrate tethering with one molecule MsOH (E2a) is 17.3 kcal/mol lower than
that of protonated oxime E2aH+, which could be more stable to form E2a for E2aH+
Figure 1.
The Gibbs free energy profiles of asymmetric hydrogenation of oximes to hydroxylamines
catalyzed by Ir complex
A1
(unit: kcal/mol). * represents chirality and
means this is a transition state.
Finally,
A6uR
/
A6uS
combine with the MsO
anion to regenerate the active catalytic
species
A1
and the catalytic cycle is completed. It should be noted that the energetic
span [
11
] of generating
R3a
/
S3a
products are 27.0/29.2 kcal/mol (
A1TSA56uR
/
TSA56uS
), respectively. That means that
R3a
is the dominant product instead of
S3a
,
and the energetic span is high, which is contrary with the conditions for the experiment to
be carried out at room temperature.
Therefore, the direct hydride transfer mode cannot explain well the experimental
observation. The non-covalent interactions have been proposed to be of importance in the
asymmetric reactions, which might stabilize the transition states, accelerate chemical reac-
tion and regulate the stereoselectivity [
9
,
12
,
13
]. Note that the Gibbs free energy of the oxime
substrate tethering with one molecule MsOH (
E2a
) is 17.3 kcal/mol lower than that of pro-
tonated oxime
E2aH+
, which could be more stable to form
E2a
for
E2aH+
com-
bining with the conjugated base MsO
anion. There are non-covalent interactions including
hydrogen bonds formed by
E2aH+
, MsO
and Ir catalytic species after
E2a
interacts
with
A4
. It is exergonic by 4.3 and 5.0 kcal/mol to generate
A5fR
and
A5fS
from the
interaction of
E2a
and
A4
. In the subsequent MsO
-assisted hydride transfer step from
A5fR
/
A5fS
to
A1
+
R3a
/
A1
+
S3a
, the hydride is transferred from the Ir center to the car-
bon atom of oxime moiety. The free energy barriers for this step are 10.9 and
11.0 kcal/mol
for the formation of
R3a
and
S3a
via
TSA51fR
/
TSA51fS
, respectively. Based on
Molecules 2022,27, 8349 5 of 9
the energetic span model and Curtin–Hammet principle [
14
], the energetic spans of the
reaction pathways to achieve
R-3a
and
S-3a
products adopting the
MsO-assisted
hydride
transfer mode are 13.8 (
A1TSA51fR
) and 13.2 (
A1TSA51fS
) kcal/mol, respec-
tively. It is obvious that the reaction pathways with the involvement of the MsO
anion is
much favorable than those without the MsO
anion. The difference of the energetic spans
for the reactions leading to R and S products are 0.6 kcal/mol (
∆∆
G
= 0.6 kcal/mol). The
calculated results show that
S3a
is the dominant product. The predicted enantioselectiv-
ity is 73:27 e.r., which agrees well with that obtained in the experiment (70:30 e.r.) [
10
]. This
indicates that the non-covalent interactions including hydrogen bonds among
E2aH+
,
MsO
and Ir catalytic species are very important, as they could not only stabilize the
catalyst but also turn over the stereoselectivity of the reaction.
Meanwhile, the asymmetric hydrogenation of oxime
E1a
to hydroxylamine cat-
alyzed by Ir catalyst
B1
was also investigated using DFT method in this work. The free
energy profiles for this reaction catalyzed by
B1
are shown in Figure 2. Following the simi-
lar reaction pathways as discussed for catalyst
A1
above, in the dihydrogen activation step
from
B1
to
B4
molecular hydrogen replaces weakly coordinated the MsO
anion of
B1
to
afford
B2
at first, which is endergonic by 9.3 kcal/mol. Then
B2
proceeds with a heterolytic
dihydrogen splitting to achieve
B3
via
TSB23
with a free energy barrier of 3.0 kcal/mol.
Subsequently,
B3
releases MsOH to form iridium hydride
intermediate B4
. This step is
exergonic by 4.6 kcal/mol. In the following hydride transfer step (the chirality-determining
step), the hydride of metal center of B4 could be stereoselectively transferred to the proto-
nated substrate with or without MsO
anion assistance. Similarly, MsOH tethered substrate
E2a releases the MsOanion to afford protonated substrate E2aH+before the inter-
action with the iridium hydride intermediate
B4
. In the situation without the participation
of the MsO
anion,
E2aH+
interacts with
B4
along its Si/Re face to form
B5uR
/
B5uS
intermediates, which could achieve the final chiral hydrogenated products. Instead, the
intermediate
B5fR
/
B5fS
could be formed by the combination of
B4
and
E2a
with the
participation of the MsO
anion. Similarly, the intermediates
B5fR
/
B5fS
are more stable
than
B5uR
/
B5uS
due to the formation of hydrogen bonds among
E2aH+
and MsO
anion in the former ones. The calculated free energy barriers of the MsO
assisted hydride
transfer step of this reaction to achieve
R-3a
and
S-3a
products are 6.6 (
B5fRTSB5f1R
)
and 9.4 (
B5fSTSB5f1S
) kcal/mol, respectively. Compared to the MsO
-assisted hy-
dride transfer, those adopting the direct hydride transfer mode without the involvement
of the MsO
anion are 18.7 (
B1TSB51uR
) and 22.8 (
B1TSB51uS
) kcal/mol, re-
spectively. It is obvious that the MsO
-assisted hydride transfer with MsO
anion is
much more favorable than the direct hydride transfer without MsO
anion. In addition,
according to the Curtin–Hammet principle, the difference in relative free energies between
TSB51fS
and
TSB51fR
is 1.9 kcal/mol (
∆∆
G
=
1.9 kcal/mol
). This implies that the hy-
droxylamine
S3a
is the favorable product and the predicted enantioselectivity is 96:4 e.r.,
this also agrees well with the result reported by experiment (89:11 e.r.) [10].
In general, the calculated results above for the asymmetric hydrogenation of oximes
to hydroxylamines catalyzed by iridium complexes
A1
and
B1
could predict that the final
favorable product would be
S3a
, which is consistent with experimental results. The
stereoselectivity of
B1
is better than that of
A1
. The hydride transfer step is the chirality-
determining step and the MsO
anion is proposed to be of importance in this asymmetric
hydrogenation of oximes. The MsO
-assisted hydride transfer mode is superior to the
direct hydride transfer mode.
In order to explore the origin of stereoselectivity of asymmetric hydrogenation of
oximes to hydroxylamines catalyzed by chiral Ir catalysts (
A1
and
B1
), the independent
gradient model based on Hirshfeld partition of molecular density (IGMH) analysis were
employed using Multiwfn software and were visualized using VMD software [
15
]. IGMH
analysis could characterize and visualize the non-covalent interactions of key structures
along reaction pathways for this reaction. IGMH analysis of key transition state structures
TSB51fR
/
TSB51fS
for the hydride transfer step in the asymmetric hydrogenation
Molecules 2022,27, 8349 6 of 9
of oximes to hydroxylamines catalyzed by
B1
are shown in Figure 3. It could be seen
that there existed a strong hydrogen bond interaction between the MsO
anion and the
substrate in the anion-assisted hydride transfer step. This hydrogen bond interaction
could stabilize the catalyst and make the MsO
anion-assisted hydride transfer mode
superior to the direct hydride transfer. In addition, we found that the distance between
the hydrogen atom of benzene ring moiety of catalyst
B1
and the hydrogen atom of
the substrate is close (marked by a red cycle in Figure 3). The distance between the
two hydrogen atoms in
TSB51fR
/
TSB51fS
is 1.904 Å/2.191 Å, respectively. IGMH
analysis visualized this van der Waals repulsion, and the distance in
TSB51fR
is closer
than that in
TSB51fS
, implying that the repulsion interaction in
TSB51fR
is greater
than that in
TSB51fS
. This might be the origins of stereoselectivity of this reaction and
the instability of
TSB51fR
compared to
TSB51fS
. However, this recognition effect is
weakened in the catalyst
A1
due to the smaller ligand and could not be visualized by
IGMH analysis (see
Figure S4 of SM
). The distances between the two hydrogen atoms of
TSA51fR
and
TSA51fS
are 2.221 Å and 2.291 Å, respectively, which are longer than
those of corresponding TSB51fR and TSB51fS.
Molecules 2022, 27, x FOR PEER REVIEW 6 of 9
Figure 2. The Gibbs free energy profiles of asymmetric hydrogenation of oximes to hydroxylamines
catalyzed by Ir complex B1 (unit: kcal/mol). * represents chirality and ‡ means this is a transition
state.
In order to explore the origin of stereoselectivity of asymmetric hydrogenation of ox-
imes to hydroxylamines catalyzed by chiral Ir catalysts (A1 and B1), the independent gra-
dient model based on Hirshfeld partition of molecular density (IGMH) analysis were em-
ployed using Multiwfn software and were visualized using VMD software [15]. IGMH
analysis could characterize and visualize the non-covalent interactions of key structures
along reaction pathways for this reaction. IGMH analysis of key transition state structures
TSB51fR/TSB51fS for the hydride transfer step in the asymmetric hydrogenation of
oximes to hydroxylamines catalyzed by B1 are shown in Figure 3. It could be seen that
there existed a strong hydrogen bond interaction between the MsO anion and the sub-
strate in the anion-assisted hydride transfer step. This hydrogen bond interaction could
stabilize the catalyst and make the MsO anion-assisted hydride transfer mode superior
to the direct hydride transfer. In addition, we found that the distance between the hydro-
gen atom of benzene ring moiety of catalyst B1 and the hydrogen atom of the substrate is
close (marked by a red cycle in Figure 3). The distance between the two hydrogen atoms
in TSB51fR/TSB51fS is 1.904 Å/2.191 Å, respectively. IGMH analysis visualized this
van der Waals repulsion, and the distance in TSB51fR is closer than that in TSB51fS,
implying that the repulsion interaction in TSB51fR is greater than that in TSB51fS. This
might be the origins of stereoselectivity of this reaction and the instability of TSB51fR
compared to TSB51fS. However, this recognition effect is weakened in the catalyst A1
due to the smaller ligand and could not be visualized by IGMH analysis (see Figure S4 of
SM). The distances between the two hydrogen atoms of TSA51fR and TSA51fS are
2.221 Å and 2.291 Å, respectively, which are longer than those of corresponding TSB51fR
and TSB51fS.
Figure 2.
The Gibbs free energy profiles of asymmetric hydrogenation of oximes to hydroxylamines
catalyzed by Ir complex
B1
(unit: kcal/mol). * represents chirality and
means this is a transition state.
Molecules 2022, 27, x FOR PEER REVIEW 7 of 9
Figure 3. The IGMH analysis of the transition states of the MsO-assisted hydride transfer step of
the reaction by catalyst B1 (δginter = 0.008 a.u.).
3. Computational Methods
Being consistent with our previous research on the mechanisms of hydrogenation
reaction catalyzed by transition-metal complexes [16–24], all the calculations were per-
formed using the ωB97X-D/BSI method employing the Gaussian 09 program [25,26]. BSI
indicates that the LANL2DZ basis set was used for Ir atom and 6-31G* basis set for all
other non-metal atoms [2729]. Additionally, single point energies were calculated with
ORCA package at the ωB97M-V/def2-TZVP level using optimized geometries at the
ωB97X-D/BSI level [3033]. The solution model based on the density (SMD) solvation
model using tert-amyl alcohol as the solvent (ε = 5.78) was employed in calculations [34].
The frequency analyses were performed to verify that all transition states have one and
only one imaginary frequency. The intrinsic reaction coordinate (IRC) calculations were
performed for key steps to confirm transition states connecting two desired minima [35].
The quasi-rigid-rotor harmonic oscillator was used to consider low-frequency contribu-
tions utilizing the Shermo code [36,37]. To correct the overestimations of entropy contri-
butions due to using the ideal gas phase model in the Gaussian program, we applied a
correction of (n-m)*1.9 for a process from m components to n components according to the
reference [38]. The non-covalent interactions were observed by mapping the independent
gradient model based on Hirshfeld partition (IGMH) surfaces using Multiwfn [39,40]. Un-
less otherwise stated, all energies of stationary points are the Gibbs free energies calcu-
lated at 298.15 K, 1.0 atm and compared to A1 or B1. Total energies, Cartesian coordinates
of all optimized structures and the calculation formula for enantioselectivity ratios (e.r.)
are given in the Supplementary Materials (SM).
4. Conclusions
In summary, the reaction mechanism and the origin of stereoselectivity of asymmet-
ric hydrogenation of oximes to hydroxylamines catalyzed by the cyclometalated iridium
(III) complexes (A1 and B1) with chiral substituted single cyclopentadienyl ligand were
investigated using DFT method. The hydride transfer step is proposed to be the chirality-
determining step. The reaction could proceed via the direct hydride transfer or the MsO-
assisted hydride transfer. The calculated results showed that the major product should be
the hydroxylamines S3a along the MsO-assisted hydride transfer pathway, and that the
involvement of the MsO anion be of importance in this asymmetric hydrogenation of
oximes catalyzed by A1 and B1. This agrees well with experimental results and implies
that the non-covalent interactions including hydrogen bonds formed by the protonated
Figure 3.
The IGMH analysis of the transition states of the MsO
-assisted hydride transfer step of
the reaction by catalyst B1 (δginter = 0.008 a.u.).
Molecules 2022,27, 8349 7 of 9
3. Computational Methods
Being consistent with our previous research on the mechanisms of hydrogenation
reaction catalyzed by transition-metal complexes [
16
24
], all the calculations were per-
formed using the
ω
B97X-D/BSI method employing the Gaussian 09 program [
25
,
26
]. BSI
indicates that the LANL2DZ basis set was used for Ir atom and 6-31G* basis set for all
other non-metal atoms [
27
29
]. Additionally, single point energies were calculated with
ORCA package at the
ω
B97M-V/def2-TZVP level using optimized geometries at the
ωB97X-D/BSI level
[
30
33
]. The solution model based on the density (SMD) solvation
model using tert-amyl alcohol as the solvent (
ε
= 5.78) was employed in calculations [
34
].
The frequency analyses were performed to verify that all transition states have one and
only one imaginary frequency. The intrinsic reaction coordinate (IRC) calculations were
performed for key steps to confirm transition states connecting two desired minima [
35
].
The quasi-rigid-rotor harmonic oscillator was used to consider low-frequency contributions
utilizing the Shermo code [
36
,
37
]. To correct the overestimations of entropy contributions
due to using the ideal gas phase model in the Gaussian program, we applied a correction of
(n-m)*1.9 for a process from mcomponents to ncomponents according to the reference [
38
].
The non-covalent interactions were observed by mapping the independent gradient model
based on Hirshfeld partition (IGMH) surfaces using Multiwfn [
39
,
40
]. Unless otherwise
stated, all energies of stationary points are the Gibbs free energies calculated at 298.15 K,
1.0 atm and compared to
A1
or
B1
. Total energies, Cartesian coordinates of all optimized
structures and the calculation formula for enantioselectivity ratios (e.r.) are given in the
Supplementary Materials (SM).
4. Conclusions
In summary, the reaction mechanism and the origin of stereoselectivity of asymmetric
hydrogenation of oximes to hydroxylamines catalyzed by the cyclometalated iridium (III)
complexes (A1 and B1) with chiral substituted single cyclopentadienyl ligand were inves-
tigated using DFT method. The hydride transfer step is proposed to be the chirality-
determining step. The reaction could proceed via the direct hydride transfer or the
MsO-assisted
hydride transfer. The calculated results showed that the major product
should be the hydroxylamines
S3a
along the MsO
-assisted hydride transfer pathway,
and that the involvement of the MsO
anion be of importance in this asymmetric hydro-
genation of oximes catalyzed by
A1
and
B1
. This agrees well with experimental results
and implies that the non-covalent interactions including hydrogen bonds formed by the
protonated substrate, MsO
anion and Ir catalytic species could not only stabilize the
catalyst but also turn over the stereoselectivity of the reaction. This work might provide
theoretical insights for the mechanism-based development of transition-metal catalysts for
the asymmetric hydrogenation of oximes in the future.
Supplementary Materials:
The following supporting information can be downloaded at: https:
//www.mdpi.com/article/10.3390/molecules27238349/s1. Figure S1: The Gibbs free energy profiles
of asymmetric hydrogenation of Z-oximes to hydroxylamines catalyzed by Ir complex
A1
(unit:
kcal/mol). * represents chirality and
means this is a transition state. Figure S2: The Gibbs
free energy profiles of asymmetric hydrogenation of Z-oximes to hydroxylamines catalyzed by Ir
complex B1
(unit: kcal/mol). * represents chirality and
means this is a transition state. Figure
S3: Several possible forms of
E2a
. Figure S4: The IGMH analysis of the transition states of the
MsO
-assisted hydride transfer step of the reaction by catalyst
A1
and
B1
(
δ
g
inter
= 0.008 a.u.). Figure
S5: The Gibbs free energy profiles for the hydride transfer step of the asymmetric hydrogenation of
oximes to hydroxylamines catalyzed by Ir complex
A1
using different functional (unit: kcal/mol).
* represents chirality and
means this is a transition state. Table S1: The calculated absolute electronic
energies (
E, in a.u.
), thermal free energies (G, in a.u.), and relative Gibbs energies (
G, in kcal/mol)
(Calculated at 298.15 K and 1 atm). Table S2: Calculated imaginary frequencies of transition states at
ω
B97X-D/BSI level. Table S3: Atomic cartesian coordinates of intermediates and transition states
(presented in Å). References [4143] are cited in the Supplementary Materials.
Molecules 2022,27, 8349 8 of 9
Author Contributions:
The manuscript was written through contributions of all authors. M.L.
designed this work; Q.A., Y.C., R.Z. and Z.L. performed the DFT calculations; Z.L., Y.T., M.P. and M.L.
co-wrote the manuscript; M.L. supervised the whole research. All authors have read and agreed to
the published version of the manuscript.
Funding:
This research was funded by National Natural Science Foundation of China (Grant number
22073005).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: All data is available in the main text or the Supplementary Materials.
Acknowledgments:
We thank the National Supercomputing Center in Tianjin and the High Per-
formance Computing (HPC) Platform at Beijing University of Chemical Technology (BUCT) for
providing part of the computational sources.
Conflicts of Interest: The authors declare no competing financial interests.
References
1.
Paudyal, M.P.; Adebesin, A.M.; Burt, S.R.; Ess, D.H.; Ma, Z.; Kürti, L.; Falck, J.R. Dirhodium-catalyzed c-h arene amination using
hydroxylamines. Science 2016,353, 1144–1147. [CrossRef] [PubMed]
2. Ager, D.J.; de Vries, A.H.; de Vries, J.G. Asymmetric homogeneous hydrogenations at scale. Chem. Soc. Rev. 2012,41, 3340–3380.
3.
Maj, A.M.; Suisse, I.; Agbossou-Niedercorn, F. Asymmetric hydrogenation of 2,3-dihydro-1H-inden-1-one oxime and derivatives.
Tetrahedron Asymmetry 2016,27, 268–273. [CrossRef]
4.
Li, B.; Liu, D.; Hu, Y.; Chen, J.; Zhang, Z.; Zhang, W. Nickel-catalyzed asymmetric hydrogenation of hydrazones.
Eur. J. Org. Chem.
2021,2021, 3421–3425. [CrossRef]
5.
Liu, D.; Li, B.; Chen, J.; Gridnev, I.D.; Yan, D.; Zhang, W. Ni-catalyzed asymmetric hydrogenation of n-aryl imino esters for the
efficient synthesis of chiral alpha-aryl glycines. Nat. Commun. 2020,11, 5935. [CrossRef] [PubMed]
6.
Li, B.; Chen, J.; Zhang, Z.; Gridnev, I.D.; Zhang, W. Nickel-catalyzed asymmetric hydrogenation of N-sulfonyl imines.
Angew. Chem. Int. Ed. 2019,58, 7329–7334. [CrossRef]
7.
Quan, M.; Wang, X.; Wu, L.; Gridnev, I.D.; Yang, G.; Zhang, W. Ni(ii)-catalyzed asymmetric alkenylations of ketimines.
Nat. Commun. 2018,9, 2258. [CrossRef]
8.
Mohr, J.; Oestreich, M. B(C
6
F
5
)
3
-catalyzed hydrogenation of oxime ethers without cleavage of the N-O bond.
Angew. Chem. Int. Ed.
2014,53, 13278–13281. [CrossRef]
9.
Li, B.; Chen, J.; Liu, D.; Gridnev, I.D.; Zhang, W. Nickel-catalysed asymmetric hydrogenation of oximes. Nat. Chem.
2022
,
14, 920–927. [CrossRef]
10.
Mas-Rosello, J.; Smejkal, T.; Cramer, N. Iridium-catalyzed acid-assisted asymmetric hydrogenation of oximes to hydroxylamines.
Science 2020,368, 1098–1102. [CrossRef]
11.
Kozuch, S.; Shaik, S. How to conceptualize catalytic cycles? The energetic span model. Acc. Chem. Res.
2011
,44, 101–110.
[CrossRef] [PubMed]
12.
Li, Z.-w.; Zhang, L.; Pu, M.; Lei, M. Mechanistic understanding of base-catalyzed aldimine/ketoamine condensations: An old
story and a new model. Asian J. Org. Chem. 2021,10, 634–641. [CrossRef]
13.
Liao, G.; Wu, Y.-J.; Shi, B.-F. Noncovalent interaction in transition metal-catalyzed selective c-h activation. Acta Chim. Sinica
2020
,
78, 289–298. [CrossRef]
14.
Jeffery, I.S. The Curtin-Hammett principle and the Winstein-Holness equation: New definition and recent extensions to classical
concepts. J. Chem. Educ. 1986,63, 42–48.
15. Humphrey, W.; Dalke, A.; Schulten, K. Vmd: Visual molecular dynamics. J. Mol. Graph. 1996,14, 33–38. [CrossRef] [PubMed]
16.
Zhao, Y.; Zhang, L.; Pu, M.; Lei, M. A phosphine-free Mn(i)-NNS catalyst for asymmetric transfer hydrogenation of acetophenone:
A theoretical prediction. Dalton Trans. 2021,50, 14738–14744. [CrossRef] [PubMed]
17.
Liu, Y.; Yue, X.; Li, L.; Li, Z.; Zhang, L.; Pu, M.; Yang, Z.; Wang, C.; Xiao, J.; Lei, M. Asymmetric induction with a chiral amine
catalyzed by a ru-pnp pincer complex: Insight from theoretical investigation. Inorg. Chem. 2020,59, 8404–8411. [CrossRef]
18.
Feng, R.; Xiao, A.; Zhang, X.; Tang, Y.; Lei, M. Origins of enantioselectivity in asymmetric ketone hydrogenation catalyzed by a
Ruh2(binap)(cydn) complex: Insights from a computational study. Dalton Trans. 2013,42, 2130–2145. [CrossRef]
19.
Li, L.; Pan, Y.; Lei, M. The enantioselectivity in asymmetric ketone hydrogenation catalyzed by Ruh
2
(diphosphine)(diamine)
complexes: Insights from a 3D-QSSR and DFT study. Catal. Sci. Technol. 2016,6, 4450–4457. [CrossRef]
20.
Xiao, M.; Yue, X.; Xu, R.; Tang, W.; Xue, D.; Li, C.; Lei, M.; Xiao, J.; Wang, C. Transition-metal-free hydrogen autotransfer:
Diastereoselective n-alkylation of amines with racemic alcohols. Angew. Chem. Int. Ed. 2019,58, 10528–10536. [CrossRef]
21.
Liu, Y.; Yue, X.; Luo, C.; Zhang, L.; Lei, M. Mechanisms of ketone/imine hydrogenation catalyzed by transition-metal complexes.
Energy Environ. Mater. 2019,2, 292–312. [CrossRef]
Molecules 2022,27, 8349 9 of 9
22.
Liu, C.; Zhang, L.; Li, L.; Lei, M. Theoretical design of a catalyst with both high activity and selectivity in C-H borylation.
J. Org. Chem. 2021,86, 16858–16866. [CrossRef] [PubMed]
23.
Zhou, Y.; Zhao, Y.; Shi, X.; Tang, Y.; Yang, Z.; Pu, M.; Lei, M. A theoretical study on the hydrogenation of CO
2
to methanol
catalyzed by ruthenium pincer complexes. Dalton Trans. 2022,51, 10020–10028. [CrossRef]
24.
Zhao, Y.; Zhang, L.; Tang, Y.; Pu, M.; Lei, M. A theoretical study of asymmetric ketone hydrogenation catalyzed by Mn complexes:
From the catalytic mechanism to the catalyst design. Phys. Chem. Chem. Phys. 2022,24, 13365–13375. [CrossRef] [PubMed]
25. Gaussian, version 09, revision D.01; Gaussian Inc.: Pittsburgh, PA, USA, 2010.
26.
Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections.
Phys. Chem. Chem. Phys. 2008,10, 6615–6620. [CrossRef]
27.
Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for K to Au including the outermost
core orbitals. J. Chem. Phys. 1985,82, 299–310. [CrossRef]
28.
Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-consistent molecular-orbital methods. IX. An extended gaussian-type basis for
molecular-orbital studies of organic molecules. J. Chem. Phys. 1971,54, 724–728. [CrossRef]
29.
Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self—Consistent molecular orbital methods. XII. Further extensions of gaussian—Type
basis sets for use in molecular orbital studies of organic molecules. J. Chem. Phys. 1972,56, 2257–2261. [CrossRef]
30.
Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn:
Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005,7, 3297–3305. [CrossRef]
31. Neese, F. Software update: The orca program system, version 4.0. WIREs Comput. Mol. Sci. 2018,8, e1327. [CrossRef]
32.
Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The orca quantum chemistry program package. J. Chem. Phys.
2020
,152, 224108.
[CrossRef] [PubMed]
33.
Mardirossian, N.; Head-Gordon, M.
ω
B97m-V: A combinatorially optimized, range-separated hybrid, meta-GGA density
functional with VV10 nonlocal correlation. J. Chem. Phys. 2016,144, 214110. [CrossRef] [PubMed]
34.
Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and on a continuum
model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B
2009
,113, 6378–6396.
[CrossRef] [PubMed]
35.
Hratchian, H.P.; Schlegel, H.B. Accurate reaction paths using a Hessian based predictor-corrector integrator. J. Chem. Phys.
2004
,
120, 9918–9924. [CrossRef]
36.
Grimme, S. Supramolecular binding thermodynamics by dispersion-corrected density functional theory. Chem. Eur. J.
2012
,
18, 9955–9964. [CrossRef]
37.
Lu, T.; Chen, Q. Shermo: A general code for calculating molecular thermochemistry properties. Comput. Theor. Chem.
2021
,
1200, 113249. [CrossRef]
38.
Bryantsev, V.S.; Diallo, M.S.; Goddard, W.A. Calculation of solvation free energies of charged solutes using mixed clus-
ter/continuum models. J. Phys. Chem. B 2008,112, 9709–9719. [CrossRef]
39. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012,33, 580–592. [CrossRef]
40.
Lu, T.; Chen, Q. Independent gradient model based on Hirshfeld partition: A new method for visual study of interactions in
chemical systems. J. Comput. Chem. 2022,43, 539–555. [CrossRef]
41.
Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab initio calculation of vibrational absorption and circular dichroism
spectra using density functional force fields. J. Phys. Chem. 1994,98, 11623–11627. [CrossRef]
42.
Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion
correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010,132, 154104. [CrossRef] [PubMed]
43.
Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model.
J. Chem. Phys.
1999,110, 6158–6170. [CrossRef]
Article
Full-text available
Chiral hydroxylamines are vital substances in bioscience and versatile subunits in the preparation of a variety of functional molecules. However, asymmetric and non-asymmetric synthetic approaches to these compounds are far from satisfactory. Although atom-economic metal-catalysed asymmetric hydrogenations have been studied for over 50 years, the asymmetric hydrogenation of oximes to the corresponding chiral hydroxylamines remains challenging because of the labile N–O bond and inert C=N bond. Here we report an environmentally friendly, earth-abundant, transition-metal nickel-catalysed asymmetric hydrogenation of oximes, affording the corresponding chiral hydroxylamines with up to 99% yield, 99% e.e. and with a substrate/catalyst ratio of 1,000. Computational results indicate that the weak interactions between the catalyst and substrate play crucial roles not only in the transition states, but also during the approach of the substrate to the catalyst, by selectively reducing the reaction barriers and thus improving the reaction efficiency and securing the generation of chirality. The asymmetric hydrogenation of oximes to chiral hydroxylamines is a long-standing challenge because of the labile N–O bond and inert C=N bond. Now, it has been shown that this reaction can be catalysed with a chiral nickel complex, and the weak interactions between catalyst and substrate are found to play a crucial role.
Article
Full-text available
An efficient nickel‐catalyzed asymmetric hydrogenation of hydrazones to chiral hydrazines has been realized with up to 99 % yield and 99.4 : 0.6 er. Deuterium labelling experiments indicated that the hydrazone substrates undergo imine‐enamine tautomerization in the mixed solvents. Studies on the effects of acids revealed that the required acid assistance promoted the dissociation of the active nickel catalyst in the catalytic cycle.
Article
Full-text available
Aldimine/ketoamine condensations are basic reactions in chemical synthesis. Herein a density functional theory (DFT) study was performed to characterize the mechanism of the ketoamine condensation of acetophenone with tert‐butanesulfinamide catalyzed by NaOH. The calculated results showed that the main N‐tert‐butanesulfinimines products own E configuration, which are due to both thermodynamic and kinetic effects. Meanwhile, the key roles of the substrate cage and the synergistic catalytic mode formed by metal cation and hydroxyl anion were pointed out, which imply the influence of noncovalent interactions on reaction activity. This theoretical study revealed the origin of E sulfinimines achieved in aldimine/ketoamine condensations catalyzed by base, which could provide new insights into the nature for the condensation reactions.
Article
Full-text available
Chiral α-aryl glycines play a key role in the preparation of some bioactive products, however, their catalytic asymmetric synthesis is far from being satisfactory. Herein, we report an efficient nickel-catalyzed asymmetric hydrogenation of N-aryl imino esters, affording chiral α-aryl glycines in high yields and enantioselectivities (up to 98% ee). The hydrogenation can be conducted on a gram scale with a substrate/catalyst ratio of up to 2000. The obtained chiral N-p-methoxyphenyl α-aryl glycine derivatives are not only directly useful chiral secondary amino acid esters but can also be easily deprotected by treatment with cerium ammonium nitrate for further transformations to several widely used molecules including drug intermediates and chiral ligands. Formation of a chiral Ni-H species in hydrogenation is detected by ¹H NMR. Computational results indicate that the stereo selection is determined during the approach of the substrate to the catalyst.
Article
Herein, a density functional theory (DFT) study was performed to investigate thoroughly the cascade reaction mechanism for the hydrogenation of carbon dioxide to methanol catalyzed by ruthenium pincer complex [RuH2(Me2PCH2SiMe2)2NH(CO)]. Three catalytic stages involving the hydrogenation of carbon dioxide (stage I), formic acid (stage II) and formaldehyde (stage III) were studied. The calculated results show that the dominant H2 activation strategy in the hydrogenation of CO2 to methanol may not be the methanol-assisted H2 activation, but the formate-assisted H2 activation. In this cascade reaction, all energy spans of stage I, II and III are 20.2 kcal mol-1 of the formate-assisted H2 activation. This implies that it could occur under mild conditions. Meanwhile, the catalyst is proposed to be efficient for the transfer hydrogenation using isopropanol as the hydrogen resource, and the ruthenium pincer complexes [RuH2(Me2PCH2SiMe2)2NH(CO)], [RuH2(Ph2PCH2SiMe2)2NH(CO)] and [RuH2(Me2PCH2SiMe2)2NH(CO)] exhibit similar catalytic activities for the hydrogenation of CO2 to methanol.
Article
Herein, a density functional theory (DFT) study was performed to investigate asymmetric ketone hydrogenation (AKH) catalyzed by Mn complexes, from the catalytic mechanism to the catalyst design. The calculated results indicated that the Mn(CO)2-PSiNSiP (A1, PSiNSiP = P(Ph)2Si(CH3)2NSi(CH3)2P(Ph)2) pincer complex has potential high catalytic activity for ketone hydrogenation. The Mn(CO)-LYB (B, LYB = P(Ph)2Si(CH3)2NSi(CH3)2P(Me)2) pincer complex was then designed to catalyze AKH with good stereoselectivity. The hydrogen transfer (HT) step is the chirality-determining step. To avoid the enantiomer of Mn(CO)2-LYB, which could eliminate the high stereoselectivity during AKH, novel Mn complexes with quadridentate ligands, such as Mn(CO)-LYC (C, LYC = P(CH3)2CH2Si(CH3)NSi(CH3)(Si(CH3)CH2P(CH3)2)CH2P(Ph)2) and Mn(CO)-LYD (D, LYD = P(CH3)2CH2Si(CH3)NSi(CH3)(Si(CH3)CH2P(CH3)2)CH2P(Cy)2), were designed to drive AKH with medium stereoselectivity. In order to increase the stereoselectivity of AKH, Mn(CO)-LYE (E, LYE = PH2CH2Si(CH3)NSi(CH3)(Si(CH3)CH2P(CH3)2)CH2P(Ph)2) and Mn(CO)-LYF (F, LYF = PH2CH2Si(CH3)NSi(CH3)(Si(CH3)CH2P(CH3)2)CH2P(Cy)2) were further designed and showed very good stereoselectivity, which is due to the lower deformation energy and stronger interactions between the ketone substrates and catalysts. This work may shed light on the design of cheap metal catalysts with a new ligand framework for the asymmetric hydrogenation (AH) of CX bonds (X = O, N).
Article
The powerful independent gradient model (IGM) method has been increasingly popular in visual analysis of intramolecular and intermolecular interactions in recent years. However, we frequently observed that there is an evident shortcoming of IGM map in graphically studying weak interactions, that is its isosurfaces are usually too bulgy; in these cases, not only the graphical effect is poor, but also the color on some areas on the isosurfaces is inappropriate and may lead to erroneous analysis conclusions. In addition, the IGM method was originally proposed based on promolecular density, which is quite crude and does not take actual electronic structure into account. In this article, we propose an improvement version of IGM, namely IGM based on Hirshfeld partition of molecular density (IGMH), which replaces the free-state atomic densities involved in the IGM method with the atomic densities derived by Hirshfeld partition of actual molecular electron density. This change makes IGM have more rigorous physical background. A large number of application examples in this article, including molecular and periodic systems, weak and chemical bond interactions, fully demonstrate the important value of IGMH in intuitively understanding interactions in chemical systems. Comparisons also showed that the IGMH usually has markedly better graphical effect than IGM and overcomes known problems in IGM. Currently IGMH analysis has been supported in our wavefunction analysis code Multiwfn (http://sobereva.com/multiwfn). We hope that IGMH will become a new useful method among chemists for exploring interactions in wide variety of chemical systems.
Article
The density functional theory (DFT) method was employed to investigate the reaction mechanism of the hydrogen activation and asymmetric transfer hydrogenation (ATH) of acetophenone catalyzed by a well-defined phosphine-free Mn(I)-NNS complex. The calculation results indicate that the Mn-NNS complex has potential high catalytic hydrogenation activity. Meanwhile, the hydrogen transfer step of this reaction is proposed to be a concerted but asynchronous process, and the hydride transfer precedes proton transfer. This work also pointed out that the stereoselectivity of ATH catalyzed by the Mn(I)-NNS complex mainly originates from the noncovalent interaction between the substrate and the catalyst. Additionally, the catalytic activities of Mn-NNS complexes with different X ligands (X = CO, Cl, H, OMe, NCMe, CCMe, and CHCHMe) were compared, and the calculated total reaction energy barriers were all viable, which indicates that these Mn-NNS complexes show higher CO bond hydrogenation activity under mild conditions. This theoretical study predicts that the reactions catalyzed by complexes with H and NCMe ligands exhibit high stereoselectivity with enantiomeric excess (ee) values of 97% and 93%, respectively.
Article
In this article, we present a stand-alone, versatile and flexible code named Shermo for calculating various thermochemistry data. This code is compatible with various mainstream quantum chemistry codes, and has many unique advantages: The output information is very clear and easy to read; thermodynamic quantities can be fully decomposed to contributions of various sources to gain a deeper insight; temperature and pressure can be conveniently scanned; two quasi-rigid-rotor harmonic oscillator (quasi-RRHO) models are supported to properly deal with contributions of low frequencies; different frequency scale factors can be simultaneously specified for calculating different thermodynamic quantities; conformation weighted thermodynamic data can be directly evaluated; the code can be easily run and embedded into shell script to automatically process large amount of systems. We hope the Shermo program will bring great convenience to quantum chemists in their daily research. The Shermo code can be freely obtained at http://sobereva.com/soft/shermo.