You are on page 1of 23

Three cycles in the MoO22+-catalyzed reduction

of nitrobenzenes to anilines with pinacol. A


computational study towards the valorization of
biomass subproducts.

Sofia Kiriakidi, Carlos Silva López, and Olalla Nieto Faza∗

Departamento de Quı́mica Orgánica Fac. Quı́mica, Universidade de Vigo Campus


Lagoas-Marcosende 36310 Vigo (Spain)

E-mail: faza@uvigo.es
Phone: +34 988 368 888

Abstract

In this work, we use density functional theory to unravel the mechanism of the ni-

trobenzene to aniline reduction, catalyzed by dioxomolybdenum (VI) dichloride. The

use of pinacol as an oxoaccepting reagent and the production of only acetone and water

as byproducts, signals a novel and environmentally friendly way to add value to the

oxygen-rich biomass-derived polyols. The reaction proceeds through three consecutive

cycles, each one responsible for one of the three reductive steps needed to yield ni-

troaniline from nitrobenzene, with nitrosobenzene and benzylnitrene as intermediates.

Each cycle regenerates the Mo(VI) catalyst and releases two acetone molecules. The

mechanism involves singlet/triplet state crossings, a feature that has been found to be

key in related polyoxomolibdate catalyzed processes. The role of the Mo-coordinated

water, product of the reduction of pinacol, as the provider of the mysterious protons

1
needed to reduce the nitro group, was revealed. The disclosure of this challenging

mechanism and its rate limiting step can contribute to the design of more effective

Mo(VI) catalysts.

Introduction

The need for a green and sustainable economy has been a major concern both in the European
Union (EU) and worldwide for more than over a decade now, resulting in the launch of
Bioeconomy as a global strategy and to the investment of trillions of euros in designing
policies of waste valorisation. 1
One large contribution to the circular bioeconomy concept can be the conversion of agri-
cultural waste into added-value products, 2 prior or in parallel to its use as an energy resource,
through biodiesel or bioethanol technologies,, 3 thus achieving maximal sustainability.
Waste produced from the agricultural sector is rich in organic material such as sugars and
polyols, the oxygen content of which is usually too high for its incorporation to the current
chemical industry production lines. 4 Oxygen atom transfer reactions (OAT) catalysed by
dioxomolybdenum complexes have been shown to be a valuable tool in reducing the oxygen
content of biomass subproducts. They can be used to transform the highly oxygenated
biomass-derived polyols into compounds similar to those obtained from fossil fuels, providing
a renewable source of feedstocks for the chemical industry. 5–8
The catalytic behaviour of various dioxomolybdenum (VI) complexes, capable of mim-
icking the action of Mo-containing ofoxotransferases that catalyze metabolism-related oxy-
gen transfer reactions in biological systems, 9 has been widely studied, due to their ability
to serve as a relatively non-toxic and inexpensive alternative compared to precious met-
als or to the need for hazardous/toxic oxo-accepting reagents. 10–13 The use of Mo shows
also a clear advantage in terms of economy and geopolitics, as compared to the efficient and
widely applied Re-catalysed deoxygenation, due to the reduced cost and geographically wider
availability of the former metal. 14 Dioxomolybdenum(VI) dichloride complexes have been

2
proved to successfully catalyze the reduction of sulfoxides and nitroaromatics using pinacol
as an oxo-accepting reagent. These reductions with pinacol show excellent chemoselectivity
against other potentially reducible groups, while only water and acetone are generated as
by-products. 11 In this context, pinacol is expected to be a simple model system for the longer
polyol chains found in biomass, such as sugars, or the very large quantities of glycerol that
are generated as a subproduct of the biodiesel industry.
This chemoselective reduction of both sulfoxides and nitroarenes also works on challenging
substrates containing C=C, C≡C, C=O, C≡N, halides, and −OH groups. Moreover, the
methodology proposed by Garcı́a et al. involves clean reactions that lead to a facile access
of sulfide and aniline scaffolds with high isolated yields (reaching up to 99%) and easy
purification, using a simple extraction to remove the catalyst, the remaining pinacol and the
water and acetone byproducts. In the case of nitroarene to aniline reduction, the reaction
evolves easily under moderate conditions (130◦ C) despite the large change in oxidation state
from the −N O2 to the −N H2 moieties.
These advantages in safety, selectivity, accessibility and reaction conditions compared
to other methods for accesing anilines 15 along with the possibility to extend this scheme
to more complicated diols, led us to further investigate this complex reaction manifold and
shed light on its intriguing underlying mechanism.
In this work we use Density Functional Theory (DFT) calculations to study the mech-
anism of the nitrobenzene reduction to aniline, with the use of pinacol as the oxoaccepting
reagent for the OAT reduction, catalyzed by dioxomolybdenum(VI) dichloride. This reac-
tion, shown in Scheme 1, serves as a simplified model for the nitroreductions studied by
Garcı́a et al. in ref. 11. Previous mechanistic investigations of Mo-catalyzed reactions in-
clude the Mo-mediated Cadogan reaction presented in Ref. 6, the OAT from sulfoxides to
phosphines, 16 olefin epoxidation, 8,17,18 hydrosilylation 19 or silane- and hydorgen-mediated
reductions. 20,21 The need for a robust theoretical investigation of the underlying catalytic
mechanisms emerges from the complexity of the process, which includes several spin-state

3
changes and non-adiabatic transitions. Based on our experience in similar chemistry 6,16,22
that evolves through multisurface pathways, we took special care in the determination of the
spin-state of each minimum and transition structure along the analyzed reaction paths. In
case of spin-crossing events along the reaction pathway, we calculated the Minimum Energy
Crossing Point (MECP) 23 on the seam formed between the two surfaces and compared it
with the closest adiabatic Transition State (TS).
The complexity of this reaction is even more pronounced when one looks in more detail
at the large change in oxidation-state involved in the transformation of a nitro to an amino
moiety. Where are the hydrogens of the N H2 − moiety coming from, since the reaction
takes place in non-aqueous media? And how are six electrons exchanged in this red-ox
reaction? The intricate underlying mechanism makes it impossible to draw a simple curly-
arrow proposal. Balancing the chemical equation, however, can give some hints that will
guide our computational study.

experimental conditions:

NO2 HO OH NH2
Mo2Cl2(dmf)2 (5%mol)
+ + (CH3)2CHOH + H2O
o
o-xylene, 130 C, 8h

(4 equiv)

balanced equation:

C6H5NO2 + 3C6H14O2 C6H7N + 6C3H6O + 2H2O

Scheme 1: Conversion of a general nitrobenzyl scaffold into its aniline derivative via a
[M o2 O2 ]2+ reduction.

4
Computational Details

All geometry optimizations have been carried out using the Gaussian 16 code, 24 at the
UB3PW91/ Def2SVP level of theory, 25–28 where the inner electrons of Mo are described by
the SDD(28, MWB) effective core potential. 29 The SCF convergence was conducted using
tight criteria and an ultrafine grid. A superfine grid was used for the most challenging
structures. The nature of the stationary points as minima or TSs was characterized by
analysis of the normal modes via diagonalization of the Hessian matrix. In challenging
cases, IRC calculations were performed in order to connect without ambiguity the minima
that are associated with a TS. The presence of species with a strong diradical character
and the associated mixing of spin states, makes necessary the examination of the stability
of the wavefunction at every stationary point. Thus, we have checked that all calculated
wavefunctions are minima in the configuration space of the basis set being used. 30
Solvation effects have been taken into account using the PCM 31 implicit solvation model
with the solvent o-xylene, in order to model the experimental conditions described in Ref.
11. Bond orders and atomic charges were calculated with the natural bond orbital (NBO)
method. 32 The MECP between singlet and triplet surfaces were computed using the MECP
algorithm 33 implemented in the ORCA 5.0.3 code, 34 using the same level of theory. A
conductor-like polarizable continuum (CPCM) solvent model employing toluene instead of
o-xylene has been used due to the lack of parameters for the latter in the ORCA code.
In order to be able to compare the MECP with the corresponding stationary points (6-t
and 7-s) and the TS connecting them (ts-67-t) we performed geometry optimizations and
frequency calculations for these structures in ORCA as well, employing the same level of
theory and solvent as for the MECP. Nucleus Independent Chemical Shifts (NICS) 35 have
been calculated on ts-45-s in order to investigate whether this metallacycle can be deemed
aromatic, as a means to determine whether this step proceeds through a concerted pericyclic
or a radical mechanism.

5
Results and Discussion

The reaction mechanism that we propose can be conceptualized as three sequential cycles,
each involving the formal transfer of two electrons from a pinacol molecule onto the nitrogen
atom through a Mo(VI)/Mo(IV) catalytic cycle. All three cycles, which we have identified
with three different colors in Figure 1 start with the oxidative cleavage of a pinacol molecule,
which reduces MoO2 Cl2 to MoOCl2 (top half of the scheme). Then, in the first cycle (cycle
A, in blue) nitrobenzene (shaded in blue) is reduced to nitrosobenzene (shaded in green); in
the second (cycle B in green ) the nitroso group (shaded in green) is reduced to a nitrene
(shaded in purple) ; and in the third (cycle C, in purple) the nitrene (shaded in purple) is
transformed into aniline (shaded in orange), while the molybdenum(VI) original catalyst is
regenerated.
In our computational study, we chose dioxomolybdenum(VI) dichloride as the starting
catalytic species, in order to model the M o2 Cl2 (DM F )2 catalyst used in the experiment
by Garcı́a et al. Taking into account the lability of the dimethylformamide ligands and our
previous experience in the effect of the number of coordinating ligands on Mo(VI)/Mo(IV))
mechanisms, 36 we assume that our model catalyst, the active naked dioxomolybdenum core,
is a good approximation, which has the advantage of avoiding the complex conformational
space of the DMF ligands. The latter are expected to be easily cleaved from the metal center
under the reaction conditions (130◦ C). Previous studies 12 have demonstrated that the oxo-
ligands are very important for the stabilization of the high oxidation state of Mo(VI), serving
as both π− and σ−donors that engage in double bonds through the occupied p orbitals of
the electron deficient metal center.

Cycle A: Nitro- to nitrosobenzene reduction

The first steps in this catalytic cycle (illustrated with blue color in Figure 1) are common
to the two subsequent cycles. It starts by the formation of a van der Waals complex (2-s)

6
ts-34-s
Cl 16.6
Cl Cl
ts-23-s O
O Mo O O
32.9 Mo
Cl O
O H O
O
H
H H
4.4 15.5
H 7.9
3-s 4-s ts-45-s
O
O VI O O
Mo Cl 2x
Cl Cl O
HO O
Mo
2.5 Cl O
2-s O
pinacol H
-3.1
H

5-s

Cl Cl
O VI O
OH 0.0 -2.8 O Mo IV
NH Mo 1-s 1-red-t
1st reduction Cl Cl O
10-s
NO ts-71-s cycle A H H
-67.1 12.2
+ water
Cl
O Cl NO2
O
-31.3 Mo IV O
Ar N Cl
ts-910-s O
O Mo
Cl
v=1492 O
-19.4 H H
N O
Ar O H H
-10.3
2nd reduction 7-s -13.7
-5.5
ts-89-s ts-67-t
6-t
cycle B
-24.4
water + N
NO

O N Ar
9 (nitrene + water)
Cl
Mo O
-54.2
Cl

cycle C
O
H H

8-s
ts-to-aniline-s OH
-70.3 -63.0
3rd reduction NH

+ water Ar NH
NH2
O H
Cl
O Mo
-134.2 Cl
O
H H

11-s
-72.8

C6H5NO2 + 3C6H14O2 C6H7N + 6C3H6O + 2H2O

Figure 1: The suggested catalytic mechanism of molybdenum(VI) dichloride dioxide con-


verting nitrobenzene to aniline and pinacol to acetone and water. The catalytic cycle is
conceptualized as three sequential subcycles depicted in blue (nitro- to nitrosobenzene re-
duction), green (nitroso to nitrene reduction) and purple (nitrene to aniline reduction). The
final reaction products are highlighted with orange shaded boxes. Spin surface crossings are
indicated with a grey funnel. It is worth noticing that a water molecule is generated at every
run of the cycle, but the one produced along the second reduction needs to enter the third
cycle.

7
between 1-s and one pinacol molecule. The reaction then proceeds through a two-step con-
densation of the diol to form the molybdodioxolane 4-s. The first step is a nucleophilic attack
by one of the two hydroxylic oxygens on pinacol to one of the Mo=O π ∗ orbitals, concerted
with a proton transfer to the associated molybdenum oxo group. The enhancement of this
anti-bonding contribution leads to a reduced double character of the Mo-O bond (going from
1.69 Å in 1-s to 1.92 Å in 3-s), which is supported by the formation of the new O–H bond.
This leads to the formation of a hydroxyl group bound to the metal center. The activation
energy for ts-23-s lies at 30.4 kcal/mol, which is in excellent agreement with the reaction
times and temperatures reported in Ref. 11 ( 8h at 130◦ C). Previous mechanistic studies
on the reactivity of dioxomolybdenum(VI) complexes with phosphines 37,38 also highlight the
initial nucleophilic attack as the rate determining step. The departure of one oxygen atom
from the Mo center, due to the effect of the reducing agent (be it pinacol, phosphine or other
species), that in this case takes place through the transformation of the C=O bond to a
loosely coordinated hydroxyl, has been shown to be responsible for the high temperatures
associated to a wide range of reactions catalyzed by the MoO2+
2 core, making this step a

likely target for any attempt to improve their efficiency.


Once this energy-demanding step is completed, water is formed through a proton transfer
from the pinacolic to the metal-coordinated hydroxyl, which is concerted with a second nu-
cleophilic attack of the former onto the molybdenum center, yielding the pinacol-coordinated
complex 4-s.
The next step involves the oxidative cleavage of the C–C bond in dioxolane 4-s leading
to a product (5-s) that has two ketones loosely coordinated to the metal center (with an
average bond distance of 2.03 Å).
Describing the mechanism of this step, starting from the dioxolane C–C bond breaking
and evolving to two ketones weakly bound to a MoIV OCl2 complex is not a trivial task. If we
just look at the first part of this sequence of events, we could classify this transformation as
a [2+2+2] retrocycloaddition/cheletropic reaction where the two bonds being broken from

8
the molybdenum center and the breaking C–C bond combine with the two newly generated
C–C π bonds to generate a cyclic array of orbitals that facilitate an aromatic transition
state. Aromaticity of the transition state 39,40 has been widely used as a measurement of the
pericyclic character of a mechanism. 22,41–43 As a proxy for aromaticity we have calculated the
NICS 44 at points along a straight line perpendicular to the C-C-O-Mo-O ring which crosses
its geometrical center. Large, negative NICS values would be associated to the shielding
generated by the magnetic field from an aromatic ring current. Maximum NICS values
about -16 ppm, such as those found for ts-45-s (see the Supporting Information), together
with the bell shape of the curve when values further from the plane are calculated, indicate
the existence of an induced ring current. This value is larger than that reported for benzene
(7 ppm), 35 although is also significantly lower than what has been obtained for the in-plane
aromatic transition structures of other [2+2+2] cycloadditions. 45 Inspection of the IRC and
the changes in bonding along it (see Supporting Information) highlight the fact that this
reaction has an early transition state, and that changes in key bonding parameters start
very close to the geometry of the reactant. The cleavage of the C–C bond begins early and
is immediately followed by the planarization of the Me2 C=O moieties, the formation of a π
bond between carbon and oxygen and the elongation of the Mo–O bonds. The synchronicity
of the changes in bonding could result in an orbital contribution to the mechanism’s driving
force and the calculated NICS. The fact that the three fragments are not completely coplanar
(it can be attributed to the steric repulsion between the pinacol methyl groups) could be
partially responsible of this decrease in conjugation/aromaticity. However, the non-negligible
spin density associated to the molybdenum atom (-0.34 au) and the two carbons whose bond
is being cleaved (0.12 au), together with the closeness of the triplet surface in this region
of the PES, makes the contribution of a radical mechanism the most likely responsible of
this decreased aromatic character. As a result, we propose this reaction as a novel pericyclic
process, but well in the frontier of application of the Woodward-Hoffmann rules. More
detailed studies are being carried out on this system to evaluate the accuracy of this claim.

9
After intermediate 5-s we observe the first spin state change of the mechanism; the reac-
tion evolves on the triplet surface through the release of the ketones, which yields the reduced
molybdenum(IV) dichloride monoxide (1-red-t). The reduced catalyst is 7.2 kcal/mol more
stable on the triplet surface than on the singlet, denoting undoubtedly that the reduction
from Mo(VI) to Mo(IV) occurs via a spin surface crossing.
This part of the first catalytic cycle, reducing a Mo(VI) center on the singlet surface to
Mo(IV) in a triplet state will be common to the other two cycles, as denoted by the green
and purple bold arced arrows above this part of the mechanism illustrated in Figure 1.
In this point, nitrobenzene enters cycle A and attacks the metal center through one of
the nitro- oxygens, forming a distorted square pyramidal complex in an exergonic process
(∆G=-13.7 kcal/mol), with an undoubtful coordination of the water molecule (Mo-water
bonds changes from 2.22Å in 1-red-t to 2.07Å in 6-t). A second spin-state surface crossing
takes place near ts-67-t (the RMSD between the TS and CP structures is only 0.6 Å) which
leads to the coordination of the second nitro oxygen on the Mo-centre. The MECP was
located and the calculated electronic energy differences show that CP67 lies at 3 kcal/mol
lower than ts-67-t and 4.1 kcal/mol higher than the singlet product 7-s, which is a slightly
distorted octahedral complex. The last step of the first catalytic cycle releases water and
nitrosobenzyl in a highly exergonic fashion (∆G=-31.3 kcal/mol) and regenerates the initial
molybdenum(VI) dichloride dioxide catalyst 1-s.

Cycle B: Nitrosobenzene to nitrene reduction

Cycle B begins following the same steps as Cycle A with the attack of a second pinacol
molecule up to the generation of the reduced catalyst 1-red-t. Then, the nitrosobenzene that
was produced by the end of Cycle A enters the cycle through the attack of the π-electrons of
the N=O bond to the empty d-orbital on the molybdenum in the reduced catalytic complex
1-red-t. After a new spin-state surface crossing, the distorted square pyramidal singlet
complex 8-s is formed, where the N=O bond is slightly elongated, going from 1.21 to 1.25

10
TS67-t
5.6

CP67
2.6

6-t
7-s
triplet surface 0.0 singlet surface
-1.5

Figure 2: Schematic representation of the spin surface crossing event between compounds
6-t and 7-s. Electronic energy differences are presented in kcal/mol, using the reactant
as a reference. The carbons and chlorines of the superimposed structures of the transition
state on the triplet surface (ts-67-t) and the MECP between the triplet and singlet surfaces
in this region (CP67 ) are depicted along the path between 6-t and 7-s in grey and cyan,
respectively. Carbon is depicted in grey, chlorine in green, molybdenum in pink and oxygens
in red for the non-superimposed structures.

11
Å. The reaction further proceeds by the breaking of the N-O bond and the forming of a
Mo-O bond in ts-89-s, with an associated barrier of 38.6 kcal/mol. However, this relatively
large barrier is more associated to the stability of 8-s than to the transition state itself
and as a consequence ts-89-s is not destabilized with respect to the reactants. An IRC
calculation starting from this transition structure showed that a three-member ring between
N, O and Mo is formed first, followed by the cleavage of the N-O bond, which leaves a loosely
coordinated nitrene, with a N-Mo bond length of 2.1 Åon the TS structure. Finally, nitrene
and water are released in an exergonic way (∆G = −54.2 kcal/mol) and the initial catalyst
1-s is regenerated, ready to take part in a new cycle.
At the beginning of the reaction, the N–O bond can be described as occupying a single
coordination site in the equatorial position of a trigonal bipyramid with apical chlorines, with
a O–N–Mo–Cl dihedral of 163 (or -44.8) degrees. Along the IRC path, this pseudo trigonal
bipyramid evolves to a octahedron, where the new oxo ligand on the molybdenum center,
together with the nitrogen of the nitrene share the same plane as the other two oxygens
(see Figure 3). A long part of the reaction path involves the rotation of the N–O bond
towards this octahedral arrangement. This conformational change is responsible of most
of the reaction barrier with an energetic cost of about 35 kcal/mol. It is remarkable that
other changes in the geometry of the complex are minimal up until this rotation has been
completed. After this point, we start seeing the strengthening of the Mo–O bond and the
cleavage of the N–O bond, which runs in parallel to the loosening of the interaction between
molybdenum and nitrogen. Thus, the energy cost associated to the O–N bond cleavage at
this stage of the reaction is partially offset by the formation of a new π Mo–O bond, so that
the contribution of this step to the overall barrier is just about 5 kcal/mol.1 The analysis of
the reaction path associated to this costly step, suggests that a judicious choice of ligands
favoring this equatorial orientation of the N–O bond might lead to a lowering of the barrier
for this second cycle in the mechanism.
1
In the Supporting Information a more complete representation of the geometry changes along the reaction
path can be found for this step, were only small variations in distances and angles are found.

12

U12  
5HODWLYH6&)HQHUJ\ NFDOPRO

U0R2
 U0R1 

2 1 0R &OGLKHGUDO


%RQGGLVWDQFH Å

 

 

 

      
,5&VWHS
Figure 3: Representation of the minimum energy path between reactants and products (IRC)
associated to ts-89-s. The variation of energy along the reaction path is indicated with a
dashed black line which has the left axis as reference. One of the O–N–Mo–Cl dihedrals is
plotted with reference to the right axis. A second left axis, with ticks and labels in violet is
used to represent bond distances. In this graph, it is clearly shown that the reaction starts
with a rotation of the O–N bond on the molybdenum complex, and that it is only after this
conformational change is finished that the cleavage of the N–O bond starts.

13
Cycle C: Nitrene to aniline reduction

The final cycle starts by the reaction of a third pinacol molecule with the oxidized MoV I
catalyst and follows the same steps as cycles A and B up to the generation of the reduced
MoIV catalyst. In parallel, the nitrene and water molecules that were released in the end
of cycle B react with each other in order to form the hydroxilamine 10-s, through another
energy demanding step with an activation energy of 34.8 kcal/mol. Nevertheless, the high
energy ts-910 corresponds to a proton transfer from water to the nitrogen atom, which
could proceed through tunnelling, due to the small mass of hydrogen and the high oscillation
frequency of 1442 cm− 1. An experiment conducted with deuterated pinacol could confirm
this explanation and also highlight the role of pinacol as the proton providing agent, through
the transfer of two protons in one of the Mo-coordinated complex oxygens, in order to form
one water molecule. This water molecule stays loosely coordinated in the metal centre
throughout all the subcycles. The IRC performed on ts-910-s shows that after the proton
transfer from the water to the nitrogen, the remaining hydroxyl attacks the nitrogen atom
in a downhill process, to form the occurying hydroxylamine, which is stabilized by 12.9
kcal/mol relatively to nitrene and water. Cycle C proceeds when the latter reacts with
the third molecule of the reduced catalyst 1-red-t and forms a distorted square pyramidal
complex 11-s, after the attack of the hydroxylaminic oxygen on the metal center, resulting
in another spin-crossing event.
Finally, a proton transfer occurs from the loosely Mo-coordinated hydroxyl to nitrogen,
with an activation energy of only 2.5 kcal/mol. In this step, the oxygen is transferred to the
molybdenum center, oxidizing it back to its VI form. The catalytic cycle ends by releasing
the newly formed aniline, water and the regenerated catalyst 1-s, releasing -134.2 kcal/mol.

14
Conclusions

In conclusion, in this work we provided with a detailed mechanism for the reduction of ni-
trobenzyl to aniline, using the molybdenum(VI) dichloride dioxide complex as a catalyst and
pinacol as a reducing agent, releasing only water and acetone as byproducts. The reaction
starts with the typical enzymatic mechanism 46 for OAT reactions, where the first step is
the reduction of the catalyst which generates a coordination vacancy into the metal center.
However, instead of substrate coordination that takes place in the enzymatic procedures, in
this case the coordination vacancy is occupied by the oxo-accepting reagent. The oxidative
cleavage of the reagent releases two acetones and a reduced catalyst, which is responsible for
the three consecutive reductions of the substrate, from −N O2 to −N H2 . A water molecule
is generated by the proton transfer from pinacol to one of the Mo-coordinated oxygens and
it stays loosely coordinated throughout the whole catalytic scheme. Then, the substrate
occupies the coordination vacancy left by the acetone departure. The first pinacol oxidation
is used for the reduction of the catalyst which is later re-oxidized in order to provoke the first
reduction from nitro to nitroso and regenerate. The latter takes place during cycle A; cy-
cles B and C start with the same pinacol reduction procedure in order to produce two more
equivalents of the key-species 1-red-t that consequently reduces nitroso to nitrene during
cycle B and finally, nitrene to aniline during cycle C. The role of the water molecule that is
formed from the pinacol’s protons and the catalyst’s oxygen is quite important, serving as a
proton source in each one of the three subsequent reductions. In contrast with the traditional
Cadogan reaction, 47 where the oxoaccepting phosphine directly eliminates the oxygens from
the nitroso compound, in this case the reducing agent pinacol acts as a proton supplier by
forming the necessary water molecule that stays coordinated in the catalyst and provides
with the necessary protons when needed. However, the formation of a nitrene intermediate
is a common and necessary step for both mechanisms. The rate determining step of the
proposed mechanism is the initial nucleophilic attack that leads to reagent coordination, as
it is also reported previously. 37,38

15
Acknowledgement

The authors thank the Centro de Supercomputacion de Galicia (CESGA) for the alloca-
tion of computational resources. This work has been funded by the Ministerio de Cien-
cia e Innovación, project: CATALISIS HOMOGENEA COMPUTACIONAL: PROCESOS
DE TRANSFERENCIA DE OXIGENO Y ACTIVACION DE ENLACES MULTIPLES
(PID2020-115789GB-C22).

Supporting Information Available

The relevant minima and TS that take part in the three catalytic cycles are uploaded in the
ioChem-BD repository (https://www.iochem-bd.org/) hosted at the Barcelona Supercom-
puting Center (https://www.bsc.es/), where cartesian coordinates, energies and frequencies
are made publicly available with the following doi: https://www.iochem-bd.org//handle/10/309208.
Supporting information including alternative steps and spin states is included.

References

(1) European Commission, Directorate-General for Research and Innovation, Innovat-


ing for sustainable growth : a bioeconomy for Europe, Publications Office, 2012,
https://data.europa.eu/doi/10.2777/6462.

(2) Stegmann, P.; Londo, M.; Junginger, M. The circular bioeconomy: Its elements and
role in European bioeconomy clusters. Resources, Conservation & Recycling: X 2020,
6, 100029.

(3) (a) Chen, J.; Zhang, B.; Luo, L.; Zhang, F.; Yi, Y.; Shan, Y.; Liu, B.; Zhou, Y.;
Wang, X.; Lü, X. A review on recycling techniques for bioethanol production from
lignocellulosic biomass. Renewable and Sustainable Energy Reviews 2021, 149, 111370;

16
(b) Mathew, G. M.; Raina, D.; Narisetty, V.; Kumar, V.; Saran, S.; Pugazhendi, A.;
Sindhu, R.; Pandey, A.; Binod, P. Recent advances in biodiesel production: Challenges
and solutions. Science of The Total Environment 2021, 794, 148751.

(4) Bozell, J. J.; Petersen, G. R. Technology development for the production of biobased
products from biorefinery carbohydrates—the US Department of Energy’s “Top 10”
revisited. Green Chemistry 2010, 12, 539–554.

(5) Lupp, D.; Christensen, N. J.; Dethlefsen, J. R.; Fristrup, P. DFT Study of the
Molybdenum-Catalyzed Deoxydehydration of Vicinal Diols. Chemistry – A European
Journal 2015, 21, 3435–3442.

(6) Castiñeira Reis, M.; Marı́n-Luna, M.; Silva López, C.; Faza, O. N. Mechanism of the
Molybdenum-Mediated Cadogan Reaction. ACS Omega 2018, 3, 7019–7026.

(7) Asako, S.; Sakae, T.; Murai, M.; Takai, K. Molybdenum-Catalyzed Stereospecific De-
oxygenation of Epoxides to Alkenes. Advanced Synthesis and Catalysis 2016, 358, 3966–
3970.

(8) Costa, P. J.; Calhorda, M. J.; Kühn, F. E. Olefin epoxidation catalyzed by η5-
cyclopentadienyl molybdenum compounds: A computational study. Organometallics
2010, 29, 303–311.

(9) Mendel, R. R. Molybdenum: biological activity and metabolism. Dalton Transactions


2005, 3404–3409.

(10) Rubio-Presa, R.; Fernández-Rodrı́guez, M. A.; Pedrosa, M. R.; Arnáiz, F. J.; Sanz, R.
Molybdenum-Catalyzed Deoxygenation of Heteroaromatic N-Oxides and Hydroxides
using Pinacol as Reducing Agent. Advanced Synthesis & Catalysis 2017, 359, 1752–
1757.

17
(11) Garcı́a, N.; Garcı́a-Garcı́a, P.; Fernández-Rodrı́guez, M. A.; Rubio, R.; Pedrosa, M. R.;
Arnáiz, F. J.; Sanz, R. Pinacol as a New Green Reducing Agent: Molybdenum- Cat-
alyzed Chemoselective Reduction of Sulfoxides and Nitroaromatics. Advanced Synthesis
& Catalysis 2012, 354, 321–327.

(12) Sanz, R.; Pedrosa, M. R. Applications of Dioxomolybdenum(VI) Complexes to Organic


Synthesis. Current Organic Synthesis 2009, 6, 239–263.

(13) Jeyakumar, K.; Chand, D. K. Application of molybdenum(VI) dichloride dioxide (MoO


2 Cl 2 ) in organic transformations. J. Chem. Sci 2009, 121, 111–123.

(14) Vesborg, P. C.; Jaramillo, T. F. Addressing the terawatt challenge: scalability in the
supply of chemical elements for renewable energy. RSC Advances 2012, 2, 7933–7947.

(15) (a) Blaser, H. U.; Steiner, H.; Studer, M. Selective Catalytic Hydrogenation of Function-
alized Nitroarenes: An Update. ChemCatChem 2009, 1, 210–221; (b) Liu, Y.; Lu, Y.;
Prashad, M.; Repič, O.; Blacklock, T. J. A Practical and Chemoselective Reduction of
Nitroarenes to Anilines Using Activated Iron. Advanced Synthesis & Catalysis 2005,
347, 217–219.

(16) Castiñeira Reis, M.; Marı́n-Luna, M.; Silva López, C.; Faza, O. N. [MoO2]2+-Mediated
Oxygen Atom Transfer via an Unusual Lewis Acid Mechanism. Inorganic Chemistry
2017, 56, 10570–10575.

(17) Herbert, M.; Montilla, F.; Álvarez, E.; Galindo, A. New insights into the mechanism
of oxodiperoxomolybdenum catalysed olefin epoxidation and the crystal structures of
several oxo–peroxo molybdenum complexes. Dalton Transactions 2012, 41, 6942–6956.

(18) Jose Calhorda, M.; Jorge Costa, P. Unveiling the Mechanisms of Catalytic Oxida-
tion Reactions Mediated by Oxo-Molybdenum Complexes: A Computational Overview.
Current Organic Chemistry 2012, 16, 65–72.

18
(19) Costa, P. J.; Romão, C. C.; Fernandes, A. C.; Royo, B.; Reis, P. M.; Calhorda, M. J.
Catalyzing Aldehyde Hydrosilylation with a Molybdenum(VI) Complex: A Density
Functional Theory Study. Chemistry – A European Journal 2007, 13, 3934–3941.

(20) Wang, Y.; Gu, P.; Wang, W.; Wei, H. Heterolytic cleavage of Si–H bonds: reduction
of imines using silane/high-valent oxo-molybdenum MoO2Cl2 as a catalyst. Catalysis
Science & Technology 2013, 4, 43–46.

(21) Reis, P. M.; Costa, P. J.; Romão, C. C.; Fernandes, J. A.; Calhorda, M. J.; Royo, B.
Hydrogen activation by high-valent oxo-molybdenum(VI) and -rhenium(VII) and -(V)
compounds. Dalton Transactions 2008, 13, 1727–1733.

(22) De Vicente Poutás, L. C.; Castiñeira Reis, M.; Sanz, R.; López, C. S.; Faza, O. N. A
Radical Mechanism for the Vanadium-Catalyzed Deoxydehydration of Glycols. Inor-
ganic Chemistry 2016, 55, 11372–11382.

(23) Chachiyo, T.; Rodriguez, J. H. A direct method for locating minimum-energy crossing
points (MECPs) in spin-forbidden transitions and nonadiabatic reactions. The Journal
of Chemical Physics 2005, 123, 094711.

(24) Frisch, M. J. et al. Gaussian 16 Revision C.01. 2016.

(25) Perdew, J. P. Electronic Structure of Solids. Akademie Verlag. Berlin, 1991; pp 11–20.

(26) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. The
Journal of Chemical Physics 1993, 98, 5648–5652.

(27) Rappoport, D.; Furche, F. Property-optimized Gaussian basis sets for molecular re-
sponse calculations. The Journal of Chemical Physics 2010, 133, 134105.

(28) 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. Physical
Chemistry Chemical Physics 2005, 7, 3297–3305.

19
(29) Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjustedab initio
pseudopotentials for the second and third row transition elements. Theoretica chimica
acta 1990, 77, 123–141.

(30) Bauernschmitt, R.; Ahlrichs, R. Stability analysis for solutions of the closed shell Kohn-
Sham equation. The Journal of Chemical Physics 1996, 104, 9047–9052.

(31) Mennucci, B.; Tomasi, J.; Cammi, R.; Cheeseman, J. R.; Frisch, M. J.; Devlin, F. J.;
Gabriel, S.; Stephens, P. J. Polarizable Continuum Model (PCM) Calculations of Sol-
vent Effects on Optical Rotations of Chiral Molecules. Journal of Physical Chemistry
A 2002, 106, 6102–6113.

(32) Reed, A. E.; Curtiss, L. A.; Weinhold, F. Intermolecular Interactions from a Natural
Bond Orbital, Donor—Acceptor Viewpoint. Chemical Reviews 1988, 88, 899–926.

(33) Poli, R.; Harvey, J. N. Spin forbidden chemical reactions of transition metal compounds.
New ideas and new computational challenges. Chemical Society Reviews 2003, 32, 1–8.

(34) Neese, F. The ORCA program system. Wiley Interdisciplinary Reviews: Computational
Molecular Science 2012, 2, 73–78.

(35) Schleyer, P. V. R.; Maerker, C.; Dransfeld, A.; Jiao, H.; Van Eikema Hommes, N. J.
Nucleus-independent chemical shifts: A simple and efficient aromaticity probe. Journal
of the American Chemical Society 1996, 118, 6317–6318.

(36) Castiñeira Reis, M.; Marı́n-Luna, M.; Silva López, C.; Faza, O. N. Mechanism of the
Molybdenum-Mediated Cadogan Reaction. ACS Omega 2018, 3, 7019–7026.

(37) Ambroziak, K.; Pelech, R.; Milchert, E.; Dziembowska, T.; Rozwadowski, Z. New diox-
omolybdenum(VI) complexes of tetradentate Schiff base as catalysts for epoxidation of
olefins. Journal of Molecular Catalysis A: Chemical 2004, 211, 9–16.

20
(38) Most, K.; Hoßbach, J.; Vidović, D.; Magull, J.; Mösch-Zanetti, N. C. Oxygen-transfer
reactions of molybdenum- and tungstendioxo complexes containing η2-pyrazolate lig-
ands. Advanced Synthesis and Catalysis 2005, 347, 463–472.

(39) Zimmerman, H. E. Molecular Orbital Correlation Diagrams, Mobius Systems, and Fac-
tors Controlling Ground- and Excited-State Reactions. II. Journal of the American
Chemical Society 1966, 88, 1566–1567.

(40) Zimmerman, H. E. Möbius and hückel systems in the SCF and CI approximations.
Tetrahedron 1982, 38, 753–758.

(41) Schleyer, P. V. R.; Wu, J. I.; Cossı́o, F. P.; Fernández, I. Aromaticity in transition
structures. Chemical Society Reviews 2014, 43, 4909–4921.

(42) Lopez, C. S.; Faza, O. N.; Souto, J. A.; Alvarez, R.; De Lera, A. R. Pseudopericyclic
design drives antara-antara [1,5] methylene sigmatropic shifts from a stepwise to a
concerted mechanism. Journal of Computational Chemistry 2007, 28, 1411–1416.

(43) Silva López, C.; Nieto Faza, O. Overview of the computational methods to assess aro-
maticity. Aromaticity: modern computational methods and applications, 2021, ISBN
978-0-12-822723-7, págs. 41-70 2021, 41–70.

(44) Chen, Z.; Wannere, C. S.; Corminboeuf, C.; Puchta, R.; Von, P.; Schleyer, R. Nucleus-
Independent Chemical Shifts (NICS) as an Aromaticity Criterion. 2005,

(45) Villar López, R.; Nieto Faza, O.; Matito, E.; López, C. S. Cycloreversion of the CO2
trimer: a paradigmatic pseudopericyclic [2 + 2 + 2] cycloaddition reaction. Organic &
Biomolecular Chemistry 2017, 15, 435–441.

(46) Millar, A. J.; Doonan, C. J.; Smith, P. D.; Nemykin, V. N.; Basu, P.; Young, C. G.
Oxygen atom transfer in models for molybdenum enzymes: isolation and structural,
spectroscopic, and computational studies of intermediates in oxygen atom transfer from

21
molybdenum(VI) to phosphorus(III). Chemistry (Weinheim an der Bergstrasse, Ger-
many) 2005, 11, 3255–3267.

(47) Cadogan, G.; Todd, M. J. Reduction of Nitro-and Nitroso-compounds by Terva-


lent Phosphorus Reagents. Part IVJ Mechanistic Aspects of the Reduction of 2,4,6-
trimethyl-2’-nitrobiphenyl, , 2-nitrobiphenyl, and nitrobenzene. Journal of the Chemi-
cal Society C: Organic 1969, 2808–2813.

22
TOC Graphic

23

You might also like