Anda di halaman 1dari 84

ARTICLE IN PRESS

Catalytic Oxidation of Alcohols:


Recent Advances
Maximilian N. Kopylovicha,*, Ana P.C. Ribeiroa,
Elisabete C.B.A. Alegriaa,b, Nuno M.R. Martinsa,
Luísa M.D.R.S. Martinsa,b, Armando J.L. Pombeiroa,*
a
Centro de Quı́mica Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, Lisbon,
Portugal
b
Chemical Engineering Department, ISEL, R. Conselheiro Emı́dio Navarro, Lisboa, Portugal
*Corresponding authors: e-mail address: maximilian.kopylovich@tecnico.ulisboa.pt; pombeiro@tecnico.
ulisboa.pt

Contents
1. Introduction 93
2. Aerobic and Peroxidative Oxidations 94
2.1 Metal Catalysts 94
2.2 Organocatalysts, Organic Radicals, and Other Additives 101
2.3 Prospective Substrates and Oxidation Agents 106
3. Acceptorless Dehydrogenative Oxidations 112
4. Oxidative Desymmetrizations 120
5. Cascade and Sequential Reactions 125
6. Conversion of Renewable Sources and Hydrogen Production 130
6.1 Transformation of Renewable Materials into Added-Value Compounds 130
6.2 Alcohol Oxidation for Hydrogen Storage and Production 132
7. Irradiation-Promoted Oxidations 134
7.1 Photocatalytic Oxidations 135
7.2 MW-Promoted Oxidations 138
7.3 Others 143
8. Catalysts Recyclization 144
8.1 Heterogeneous Solid Oxides, Alloys, and Related Materials 145
8.2 Supported Catalysts 148
8.3 Nano, Dispersed and Micellar Catalysts 150
8.4 ILs and Related Systems with Phase Division 152
8.5 Other Directions 155
Acknowledgments 157
References 157

Advances in Organometallic Chemistry # 2015 Elsevier Inc. 91


ISSN 0065-3055 All rights reserved.
http://dx.doi.org/10.1016/bs.adomc.2015.02.004
ARTICLE IN PRESS

92 Maximilian N. Kopylovich et al.

ABBREVIATIONS
[C2mim] 1-ethyl-3-methylimidazolium
[C4mim] 1-butyl-3-methylimidazolium
[C4py] 1-butyl-pyridine
[C6mim] 1-hexyl-3-methylimidazolium
[C8mim] 1-octyl-3-methylimidazolium
1-Me-AZADO 1-methyl-2-azaadamantane N-oxyl
2IBAcid 2-iodobenzoic acid
ABNO 9-azabiciclo[3.3.1]nonane N-oxyl
Aliquat N-methyl-N,N-dioctyloctan-1-ammonium chloride
AZADO 2-azaadamantane N-oxyl
Bmim 1-buthyl-3-methylimidazolium
BOX bis(oxazoline)
bpyO bis(oxazoline)α,α0 -bipyridonate
Cp cyclopentadienyl
Cp* pentamethylcyclopentadienyl
DESs deep eutective solvents
DIAD diisopropyl azodicarboxylate
DKR dynamic kinetic resolution
DPIO 4,7-bis(4-pyridyl)-1,1,3,3-tetramethylisoindolin-2-yloxyl
Emim 1-ethyl-3-methylimidazolium
FDCA 2,5-furandicarboxylic acid
HFCA 5-hydroxymethyl-2-furancarboxylic acid
HMB hexamethylbenzene
HMF 5-hydroxymethylfurfural
Hmim 3-methylimidazolium
IBA iodosobenzoic acid
IBX o-iodoxybenzoic acid
IBXF o-iodoxybenzoic acid with a fluorous tag
IL ionic liquid (room-temperature)
ketoABNO 2,2,6,6-tetramethylpiperidine-1-oxyl
LED light emitting diode
MOF metal-organic framework
MW microwave
NAD nicotinamide adenine dinucleotide
NBS N-bromosuccinimide
NHC N-heterocyclic carbene
NHPI N-hydroxyphthalimide
Nor-AZADO 2-azanoradamantane N-oxyl
NP nanoparticle
NT nanotube
NTf2 bis(trifluoromethylsulfonyl)imide
OKR oxidative kinetic resolution
Oxone potassium peroxomonosulfate KHSO5
PINO phthalimide-N-oxyl
SPB surface plasmon band
TBAB tetrabutylammonium bromide
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 93

TBHP tert-butyl hydroperoxide (tBuOOH)


TBN tert-butyl nitrite
TEMPO 2,2,6,6-tetramethylpiperidine-1-oxyl radical
TOF turnover frequency
TON turnover number

1. INTRODUCTION
The oxidation of alcohols to carbonyl-containing compounds1,2 or
their full oxidations3,4 are among the central reactions in organic chemis-
try5,6 and are of interest for the development of environmentally benign
processes,7,8 production of new materials9,10 and energy sources.11,12 Due
to their pivotal role in industrial fields and expected further
applications,13 these reactions continue to attract a great attention, disclosing
new catalysts,14,15 substrates, oxidants with peculiar features, and
applications.1
Concerning the oxidants, stoichiometric oxidations with transition-
metal compounds or sulfoxides are still in common use, despite the formation
of a large amount of undesirable products.1 The most used oxidants include
small organic molecule-based reagents, e.g., Dess-Martin periodinane,
Swern, Moffatt, Corey-Kim oxidants, SO3/pyridine, some of them being
moisture-sensitive and expensive (e.g., N,N0 -dicyclohexylcarbodiimide,
oxalyl chloride), or metal-based systems (such as Jones, Collins, Oppenauer
reagents, pyridinium chlorochromate (PCC), pyridinium dichromate, bar-
ium permanganate, manganese dioxide, ruthenium tetroxide, silver carbon-
ate).1 A recent environmental compatibility and sustainability approach leads
toward aerobic oxidations with transition-metal catalysts (based on Pd, Ru,
Fe, Cu, Pt, Au, Ir, Rh, etc.,) and dioxygen or hydrogen peroxide as
oxidants.1–14 The use of molecular oxygen as a stoichiometric reoxidant
in combination with a catalytic metal has practical advantages due to the
favorable economics associated with O2 and the formation of environmen-
tally benign by-products (water and hydrogen peroxide).
Advances on the development of new methodologies, oxidation agents,
catalysts and applications have been regularly surveyed,1–13 and the field
continues to be one of the most extensively and actively investigated areas
of current organic synthesis. Newly developed green oxidations of alcohols
usually involve active and selective recyclable catalysts that ideally should
ARTICLE IN PRESS

94 Maximilian N. Kopylovich et al.

work with dioxygen, air, or other cheap oxidants, not leaving aside toxic or
wasteful by-products.
However, despite some remarkable advances, only few of the known
methods are capable of offering an economic and practical oxidation toward
a particular industrially important transformation. Many of the found
catalytic systems suffer from high reagent cost, instability, employment of
hazardous metals or oxidants, harsh reaction conditions, operational com-
plexity, functional group incompatibility, or production of unprocessable
wastes.1 Thus, there is a continuing demand for new catalytic systems that
could overcome such challenges.
Moreover, other perspectives for alcohol oxidation have been tested,
including atom-efficient transformations (e.g., direct synthesis of esters),
hydrogen transfer and production, oxidation of natural substrates, such as
cellulose, cascade and sequential reactions, etc.
The achievements in the alcohol oxidations until 2010 have been cov-
ered in several books, book chapters, and reviews,1–15 and thus the current
work focuses on the most recent advances in the 2010–2014 period with
some notable examples back to 2005.

2. AEROBIC AND PEROXIDATIVE OXIDATIONS


Aerobic and peroxidative oxidations of alcohols, in particular of
benzylic alcohols, are typical model reactions due to their importance and
generality; inexpensive O2, H2O2 or tert-butyl hydroperoxide (TBHP) oxi-
dants, and simple procedures are usually involved.1–15 In this section, an
overview of some interesting catalytic systems, which were lately introduced
into the field of alcohol oxidation, is presented. This concerns mainly homo-
geneous systems, since recent advances on heterogeneous catalysts are
included in Section 8. Moreover, a glance at new substrates and oxidants
which could successfully be used in a near future and make a difference
in terms of efficiency, selectivity, economy and/or sustainability of the
processes, is also presented.

2.1 Metal Catalysts


Historically, Pt, Pd, Ru, Ir, and Rh complexes were among the most
effective catalysts for alcohol oxidation. The series was expanded to 3d
metals, e.g., V16 and Cu/TEMPO systems (TEMPO ¼ 2,2,6,6-
tetramethylpiperidine-1-oxyl radical) which have been reinvented and
developed,17,18 and now includes representatives of most of the groups
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 95

and subgroups of transition and even nontransition metals.19 In contrast to


those of noble metals, the newly introduced catalysts based on abundant 3d
and related metals typically operate by redox mechanisms that usually
involve one-electron (radical) processes.20 During the last decade, the search
for new and effective catalysts for alcohol oxidation has mainly concentrated
on finding cheaper and more effective metal–ligand combinations,21 achiev-
ing regio- and enantioselective reactions,22 and explaining mechanistic
details of action of known catalytic systems.23 In addition, recent reports24,25
showed that there is a continuous interest concerning the structural details of
the catalysts. For instance, the role of nuclearity in multinuclear copper(II)
complexes26,27 was recently discussed.28
When using dioxygen, a challenge to overcome concerns the fact that it
is a four-electron oxidant, while the aerobic oxidation of alcohols to car-
bonyl compounds involves two electrons. Apart from that, partially reduced
oxygen species are usually more reactive than O2 itself. Hence, the introduc-
tion of special “oxidation buffer agents” which can balance the specific ener-
getic requirements of the substrates with the possibilities of the oxidant is an
important task, and complex, metallorganic, or organocatalysts can play the
role of such agents. Generally, the effective catalytic systems contain an
organic component, as a ligand in a coordination compound or an additive,
but sometimes simple salts, such as Mn(II) acetate,29 can be efficient catalysts.
Organic ligands in complexes can play different roles: adjust electronic
and steric properties, provide the required solubility or arrangement of cen-
tral metal ions or protect them from overoxidation or reduction. Thus, a
systematic study of the catalytic activity of palladium complexes with com-
mercially available pyridine-containing ligands30 found the conditions
where precipitation of Pd black does not occur. Similarly, tertiary phosphine
oxides (O]PR3) can be used as ligands for Pd catalysts.31 Recently, the
importance of the trinuclear Pd3O2 intermediate [(LPdII)3(μ3–O)2]2+
(L ¼ 2,9-dimethylphenanthroline) in Pd-assisted catalysis was unveiled.32
This trinuclear compound is a product of oxygen activation by reduced pal-
ladium species and is an important intermediate in the aerobic oxidation of
alcohols.32
The introduction of new ligands is an important task in the development
of new catalytic systems. For instance, some polymers can be used as mac-
roligands to host metal ions. The polymer ligands can not only provide the
catalyst reutilization, but also stabilize the central metal ions and prevent
their aggregation (e.g., precipitation of palladium black, if a Pd(II) catalytic
system is applied).33 This approach is rather attractive and combines the
ARTICLE IN PRESS

96 Maximilian N. Kopylovich et al.

Scheme 1 Synthesis of Pd(II) complexes with poly(l-lactide) and poly(caprolactone)


macroligands.33

advantages of homogeneous and heterogeneous catalysts. It involves coor-


dination of Pd(II) by 4-pyridinemethylene-end-capped poly(l-lactide) and
poly(caprolactone) (Scheme 1). The polymer-anchored catalysts are soluble
under the applied catalytic conditions, but upon addition of n-pentane or
methanol the polymer-anchored catalyst precipitates, and thus can be easily
separated from the reaction mixture. These catalysts are effective in the oxi-
dation of several primary and secondary alcohols with O2.33
In a related work, a design of enzyme-inspired star block-copolymers
with branched topologies and protein-like tertiary or quaternary structures
was performed.34 These polymers incorporate hydrophilic, super-
hydrophobic, and polydentate metal-binding sites and self-assemble in
water, their mode of assembly being controlled by the composition of the
polymer. An important feature of the star block-copolymers is that they
incorporate perfluorocarbons and, due to that, their emulsions in water
can attract and preconcentrate O2 in the vicinity of the active metal site.
Addition of Cu(II) and TEMPO leads to an effective catalytic system for
oxidation of alcohols to aldehydes in water.34
A series of tetradentate pyridyl-imine terminated Schiff-bases,
bis(pyridyl-imine) terminated siloxane and other related polymers, can be
used as ligands to host copper(II) ions.35 These CuBr2/polyL/TEMPO cat-
alytic systems (polyL stands for polydimethylsiloxane derived pyridyl-imine
terminated ligand) are effective for aerobic oxidations of primary and sec-
ondary alcohols under aqueous conditions. Chiral N,O-ligands, e.g., inex-
pensive L-proline, can also be used to prepare copper catalysts that are
particularly effective for the oxidation of sterically hindered, allylic or
heterocyclic alcohols such as 1-(3-pyridyl)ethanol, 1-(2-furfuryl)ethanol,
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 97

2-thienyl, 2-furyl and 3-pyridyl methanol.36 Related mono- and


dicopper(II) aminopolyalcoholates were easily prepared by self-assembly
and also studied.37 The selectivity parameters for oxidative transformations
were measured and discussed, supporting free-radical mechanisms. Copper
complexes with hydrazone ligands have been extensively studied during the
last decade as oxidation catalysts38 and their family continues to grow. For
example, an easy to synthesize and to handle trinuclear dihydrazone
copper(II) complex [Cu3(L)(μ2–Cl)2(H2O)6] can be used as a reusable
(up to eight runs) catalyst for the selective oxidation of a wide variety of
alcohols, not being deactivated by N/S-heteroatom-containing substrates.39
Copper-containing metal-organic frameworks (MOFs) based on 5-(4-
pyridyl)tetrazole building blocks, easily prepared in situ by 1,3-dipolar cyclo-
addition between 4-cyanopyridine and azide in the presence of copper(II)
chloride, were successfully applied40 as precatalysts for the low power
(10 W) microwave (MW)-assisted peroxidative oxidation of secondary
alcohols leading to the corresponding ketones with yields up to 86% and
turnover frequencies (TOFs) up to 430 h1 after 1 h, in the absence
of any added solvent or additive. Zr(IV)-based robust MOFs with open
Fe- and Cr-monocatecholato metal sites on the structure of the organic
linkers were prepared by postsynthetic metal exchange,41 an approach that
allows good control over the number of metal-binding sites, and can be used
as a facile and efficient way to obtain MOFs that cannot be directly synthe-
sized under solvothermal conditions. The Cr-metalated MOFs are efficient,
versatile, and reusable heterogeneous catalysts for the oxidation of alcohols
to ketones with TBHP or H2O2 as oxidants.
Biomimetic Cu(II) and Fe(II) complexes with bis- and tris-pyridyl
amino and imino thioether ligands and vacant (or potentially so) coordina-
tion positions (Fig. 1)42 are active as catalyst precursors for the solvent- and
halogen-free MW-assisted oxidation of 1-phenylethanol by TBHP, in the
presence of pyridazine or other N-based additives. Maximum TOF of
5220 h1 (corresponding to 87% yield) was achieved just after 5 min of reac-
tion time under the low power MW irradiation. The same authors
reported43 the catalytic activity of related copper, iron, and vanadium sys-
tems with mixed-N,S pyridine thioether ligands. The Cu and Fe complexes
proved to be useful catalysts in various MW-assisted alcohol oxidations with
TBHP, at 80 °C. Thus, S-containing ligands can also be used to create effec-
tive catalyst precursors.
Another green and easy to prepare iron-based catalyst, [Fe(BPA)2]
(OTf )2, with the commercially available bis(picolyl)amine (BPA) ligand
ARTICLE IN PRESS

98 Maximilian N. Kopylovich et al.

Figure 1 Iron and copper complexes with bis- and tris-pyridyl amino and imino
thioether ligands.42,43

Figure 2 Bis(picolyl)amine (BPA), thymine-1-acetate (THA), and


8-hydroxyquinolinate (HQL).

(Fig. 2), chemoselectively oxidizes a variety of secondary alcohols in the


presence of primary ones into the corresponding hydroxy ketones within
15 min at room temperature with 3 mol% catalyst loading and H2O2 as oxi-
dant.44 The complex can also be generated in situ and operates similarly to
the preformed one. In situ generated iron chloride complexes with thymine-
acetate or 6-(N-phenylbenzimidazoyl)-2-pyridinecarboxylic acid ligands
(Fig. 2) have been also recently synthesized and proved to be selective
and convenient catalysts for the oxidation of benzylic and allylic alcohols.45
They can be applied to sensitive compounds like perillyl alcohol, geraniol, or
carveol, while diols can be oxidized in good yields without oxidative cleav-
age of products. Mechanistic investigations reveal that thymine-acetate pos-
sesses organocatalytic activity for the oxidation of alcohols.
Aerobic alcohol oxidations with vanadium catalysts continue to widen
their substrate scope and applications.16,46 The recently studied mechanism
of the intramolecular oxidation of benzyl alcoholate ligands in
8-hydroxyquinolinato(L) vanadium(V) complexes of the type [LV(O)
(OR)] resembles those proposed for certain metalloenzyme-catalyzed
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 99

Scheme 2 Key steps in base-assisted dehydrogenation for [LV(O)(OR)] complexes.20

oxidations and involves unusual ligand exchange and intermolecular


deprotonation at the benzylic position (Scheme 2).20 This biomimetic path-
way differs from the previously identified hydride-transfer and radical mech-
anisms for transition-metal-mediated alcohol oxidations. As a result, new
ways to enhance the activity and selectivity of vanadium catalysts were pro-
posed. They include the control of the outer coordination sphere and appli-
cation of ligands with appropriately positioned pendant bases to serve as
proton shuttles.
Related V complexes show activity toward the oxidative decom-
position of pinacol with CdC bond cleavage and aerobic oxidation
of 4-methoxybenzylalcohol and other lignin model compounds.47 Other
oxidovanadium(V) complexes with cis-2,6-bis-(methanolate)-piperidine
ligands of the type depicted on Scheme 3 were applied as catalysts
to convert prochiral alkenols into 2-(tetrahydrofuran-2-yl)-2-propanols,
2-(tetrahydropyran-2-yl)-2-propanols, oxepan-3-ols and epoxides, upon
oxidative alkenol cyclization with TBHP as oxidant (Scheme 3).48
These catalysts are rather stable and possess improved chemoselectivity,
e.g., epoxidation of geraniol occurs enantioselectively. It was ruled out
the vanadium(V) tert-butyl peroxy complex formation is a key step
to activate peroxides.
Silver N-heterocyclic carbene (NHC) catalysts (Scheme 4) can be
applied not only for the selective oxidation of alcohols to aldehydes or car-
boxylic acids but also for further tandem one-pot synthesis of imines
ARTICLE IN PRESS

100 Maximilian N. Kopylovich et al.

Scheme 3 Example of cis-2,6-bis-(methanolate)-piperidine ligands (A) and oxidation of


alkenols by TBHP, catalyzed by the piperidine-derived vanadium complexes (B).48

Scheme 4 Example of a Ag(NHC) catalyst (A) and oxidation of alcohols to aldehydes or


carboxylic acids and tandem one-pot synthesis of imines catalyzed by them (B).49

Figure 3 Example of a ruthenium(III) catalyst for the aerobic oxidative dehydrogenation


of benzyl alcohols.51

(see Section 5).49 Rhodium porphyrin complexes have been successfully


applied as catalysts for the selective oxidation of functionalized
alcohols, since they tolerate a variety of functional groups, such as methoxy,
C]C, and thiofuran moieties.50 The proposed catalyst is robust and does
not degrade under the studied conditions (1 atm O2, 80 °C, 7 h).
A porphyrin rhodium(III) methoxide complex was identified as a key
intermediate in the proposed mechanism. The cyclometalated complex
bearing phenylpyridine [RuCl(ppy)(tpy)][PF6] (ppy ¼ 2-phenylpyridine;
tpy ¼ 2,20 :60 ,200 -terpyridine) (Fig. 3) is an example of ruthenium(III) catalysts
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 101

Figure 4 Ionic catalyst containing a Ru(III)-complex cation and a α-Keggin-type


phosphotungstate anion.52

for the aerobic oxidative dehydrogenation of benzyl alcohols to benzalde-


hydes.51 The complex was also applied as catalyst for the one-pot synthesis
of benzonitriles from benzyl alcohol with ammonia.
Another approach to develop new catalysts is the combination of different
metals and even different types of complexes in one system, e.g., an ionic
compound containing a Ru(III)-complex cation and a α-Keggin-type
phosphotungstate anion (Fig. 4).52 This compound is robust because the
phosphotungstate [PW12O40]3 anion, although exhibiting a negligible con-
tribution to the activation of benzyl alcohol, significantly stabilizes the struc-
ture. On the other hand, due to the high polarity and the ionic nature of the
complex, ionic liquids (ILs) can be used as solvents. The catalyst was applied
as an efficient catalyst for aerobic oxidations of alcohols, free of base and
nitroxyl radical, and can be reused at least 5 without significant loss of activ-
ity.52 Other heteronuclear complexes are effective for the solvent-free
peroxidative (with H2O2) oxidation of primary and secondary alcohols,
e.g., a trinuclear complex with a dicopper(II)–monozinc(II) center.53
Finally, it is noteworthy that the alcohol oxidation reaction can be used
for the direct one-pot synthesis of coordination compounds. The applica-
tion of such a technique allows to generate in situ aldehydes, ketones, and
other carbonylic derivatives, which are not available commercially, are
unstable or cannot be prepared and isolated by conventional methods. As
a result, new interesting coordination compounds can be prepared with
ligands which are not attainable by usual synthetic methods.54

2.2 Organocatalysts, Organic Radicals, and Other Additives


In spite of their high activity, catalytic systems that employ transition metals
exhibit a number of disadvantages. For instance, substrates with a chelating
ability can bind to the metal and hamper the reaction. Moreover, the pres-
ence of the metal can show some environmental impact. Therefore, new
transition-metal-free systems have been searched for.7,13,15 One of such
systems mimics the Anelli–Montanari protocol (see below) and employs
an oxoammonium salt that carries out substrate oxidation while NO2
ARTICLE IN PRESS

102 Maximilian N. Kopylovich et al.

(generated in situ from nitric acid, nitrates, nitrites, or hydroxylamine)


regenerates the salt.55 It is possible to operate aerobic NOx systems under
halide free conditions; however, participation of halides (bromine, hyp-
obromous acid, nitrosyl bromide, or nitrosyl chloride) as active
co-oxidants can potentially widen the scope of such reactions. In addition,
an efficient transition-metal-free catalytic system mediated by
N-bromosuccinimide (NBS) for the aerobic oxidation of various aromatic
alcohols, under mild conditions, was recently reported.56 For instance, ben-
zyl alcohol is oxidized to benzaldehyde with 99% conversion (94.5% selec-
tivity) by the 2,3-dichloro-5,6-dicyano-1,4-benzoquinone–NaNO2–NBS
system under 0.3 MPa of O2 for 2 h at 90 °C.
The NH4NO3/TEMPO/H+ catalytic system was reported57 as efficient,
under mild aerobic conditions, for the chemoselective oxidation of a com-
prehensive range of alcohols, including those bearing oxidizable hetero-
atoms (S, N, O), alkyl-, cycloalkyl-, and allyl-type substituted substrates.
Very recently, tetra-n-butylammonium bromide was successfully applied58
as simple but efficient organocatalyst for the peroxidative (with TBHP) oxi-
dation of a variety of functionalized benzylic/allylic alcohols under mild
conditions. It shows excellent selectivity for secondary benzylic alcohols
over aliphatic alcohols. The analog tetra-n-butylammonium iodide was
employed for the alfa-oxyacylation of ketones by benzylic alcohols leading
to alfa-acyloxyketones, with TBHP, affording moderate to good yields.59
1,2-Di(1-naphthyl)-1,2-ethanediamine efficiently catalyzes the oxida-
tion of alcohols by using TBHP as oxidant. Secondary benzyl alcohols are
oxidized in almost quantitative yields, and the catalyst displays a high activity
toward hindered cycloaliphatic secondary alcohols.60 Quinine-derived urea
has been identified as a highly efficient organocatalyst for the enantioselective
oxidation of 1,2-diols using bromination reagents as the oxidants, at ambient
temperature, to yield a wide range of α-hydroxy ketones in good yield (up to
94%) and excellent enantioselectivity (up to 95% ee).61
Nitroxyl radicals (Fig. 5), such as 9-azabiciclo[3.3.1]nonane N-oxyl
(ABNO), 9-azabiciclo[3.3.1]non-3-one N-oxyl (ketoABNO),
2-azaadamantane N-oxyl (AZADO), 1-methyl-2-azaadamantane N-oxyl
(1-Me-AZADO) and 2-azanoradamantane N-oxyl (Nor-AZADO), and
especially TEMPO, are widely used promoters for the aerobic oxidation
of alcohols to the corresponding carbonyl compounds due to their high effi-
cacy and selectivity.1–23 TEMPO is applied62 in industrial processes using
aerobic or Anelli–Montanari1,63 type (neither aerobic nor peroxidative)
conditions (Scheme 5): the oxoammonium salt of TEMPO carries out
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 103

Figure 5 Structures of the nitroxyl radicals (from left to right) 2,2,6,6-


tetramethylpiperidine-1-oxyl (TEMPO), 2-azaadamantane N-oxyl (AZADO),
1-methyl-2-azaadamantane N-oxyl (1-Me-AZADO), 2-azanoradamantane N-oxyl (Nor-
AZADO), 9-azabiciclo[3.3.1]nonane N-oxyl (ABNO), and 9-azabiciclo[3.3.1]non-3-one
N-oxyl (ketoABNO).

Scheme 5 Anelli–Montanari's oxidation of alcohols.63

the alcohol oxidation in the presence of an excess of hypochlorite (which


may lead to undesirable chlorinated by-products). In contrast, few applica-
tions of nitroxyl radicals as catalysts for the peroxidative oxidation of alcohols
were reported, despite its enhancing effect on the alcohols conversion to the
respective ketones.38f,40,64
Compared to TEMPO, the less hindered AZADO and ABNO radicals
exhibit significantly enhanced reactivity toward a wide range of alcohols,
including structurally hindered secondary alcohols that TEMPO fails to
efficiently oxidize due to the steric congestion near its active center.65,66
ARTICLE IN PRESS

104 Maximilian N. Kopylovich et al.

O O

N OH N O

O O
NHPI PINO
Figure 6 N-hydroxyphtalimide (NHPI) and phthalimide-N-oxyl nitroxyl (PINO) radical.

Moreover, the sterically unhindered Nor-AZADO is more catalytically


active than AZADO, 1-Me-AZADO, ABNO, and TEMPO in the aerobic
oxidation of alcohols to their corresponding carbonyl compounds.65
Other related to TEMPO additives have been used, e.g.,
N-hydroxyphthalimide (NHPI) which generates the nitroxyl radical
phthalimide-N-oxyl (PINO, Fig. 6).66 The PINO radical and analogs have
been utilized for a range of aerobic oxidation reactions. PINO has been
shown to be more reactive than TEMPO, but it is not stable and is usually
formed from NHPI in situ by means of an initiator.
Usually the stable nitroxyl radicals alone cannot directly catalyze the oxi-
dation of alcohols with dioxygen or peroxide, so they rely on the assistance
of various cocatalysts that play an important role in activating the oxidation
agent. The most used cocatalysts are first row transition-metal complexes
where Cu compounds with various N-donor ligands account for the prime
ones.67 In many instances this combination serves as some kind of model to
compare catalytic properties of copper compounds. For example, the per-
formances of two asymmetric tetranuclear (with the {Cu4(μ–O)2(μ3–
O)2N4O4} core) and dinuclear (with the {Cu2(μ–O)2N2O2} core)
copper(II) complexes were compared in the catalytic TEMPO-mediated
aerobic oxidation of benzylic alcohols.28 In spite of their similarity, the com-
plexes perform differently: the tetranuclear copper(II) (R) complex is highly
active leading to yields up to 99% and TONs up to 770, while the (S,R)-2
dinuclear complex is not so efficient under the same conditions. However,
no solid explanation of the activity differences was proposed.
Nevertheless, almost all of the nitroxyl radicals are quite inefficient for the
oxidation of aliphatic and secondary alcohols, and require an additional base
for the oxidation to proceed. One of the reasons for the poor performance
toward secondary alcohols is believed to be the steric hindrance
(as secondary alcohols are bulkier) and the involvement of a bimolecular reac-
tion between the Cu-alkoxide and TEMPO on the final step.55 This can be
overcome by switching from TEMPO to the less sterically hindered nitroxyl
radicals ABNO68 or AZADO.69 Thus, Cu(I)/ABNO systems effectively
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 105

oxidize a wide range of secondary alcohols, including substrates with bulky


groups close to the alcohol group, and still be effective for primary alcohols.
With the AZADO catalyst system, substrates containing primary and second-
ary alcohols can also be oxidized in good to excellent yields.69
In contrast to the Cu/TEMPO combination, Fe(III)/TEMPO systems
readily catalyze the aerobic oxidation of secondary alcohols and do not
usually require any base or less sterically hindered ligands.70 The best activity
was observed with weakly coordinating solvents (e.g., dichloroethane)
unlike Cu/TEMPO systems, which normally are more active in acetoni-
trile. The performance and substrate scope (primary and secondary allylic,
benzylic, or inactivated aliphatic alcohols) of Fe–TEMPO catalyst systems
is improved by the presence of NaCl.70e
The use of cobalt and manganese cocatalysts with TEMPO and its deriv-
atives has been known for some time.71 The utilization of nitrate salts sug-
gests that the oxoammonium salt oxidizes the substrates. There are also
reports with heterogeneous Cu/Mn oxide cocatalysts; in these cases low
loadings of TEMPO were used while the heterogeneous catalyst can be
recycled. Cobalt can be used as a sole effective metallic component within
a Co(NO3)2/dimethylglyoxime/TEMPO system, with only 1 mol% cata-
lyst loadings.71
The polyoxometalate H5PV2Mo10O40 oxidizes TEMPO to form the
oxoammonium salt, which then electroxidizes alcohols to their
corresponding carbonyl compounds.72 Despite the fact that this system
was used industrially (by DSM),62 it has received little research interest.
Recently, functionalized TEMPO was immobilized with an IL on silica-
coated magnetic nanoparticles (NPs).71d They catalyze the aerobic oxidation
of a range of alcohols and the catalyst can be separated with an external mag-
net and recycled 10  without a significant loss of activity.
The first example of vanadium as the sole metallic component in
TEMPO-catalyzed aerobic alcohol oxidation in acetonitrile was recently
reported.73 However, the catalyst did not perform well with secondary ali-
phatic alcohols, even with extended reaction times.
The level of sophistication of stable radicals may increase in the
near future as it is becoming clear that the performance of catalysts can
be improved by tuning both steric and electronic effects of these
radicals.74 Another successful approach is the use of heterogenized radicals.
Thus, a free-radical porous coordination polymer [Cu(DPIO)2(SiF6)]
[DPIO ¼ 4,7-bis(4-pyridyl)-1,1,3,3-tetramethylisoindolin-2-yloxyl] (Fig. 7),
possesses one-dimensional channels with incorporated nitroxyl catalytic sites.
ARTICLE IN PRESS

106 Maximilian N. Kopylovich et al.

Figure 7 4,7-Bis(4-pyridyl)-1,1,3,3-tetramethylisoindolin-2-yloxyl (DPIO).

When dioxygen or air is used as oxidant, this polymer acts as an efficient, recy-
clable, and widely applicable catalyst for selective oxidation of various alcohols
to the corresponding aldehydes or ketones.75
Fullerene has been employed successfully to anchor TEMPO moieties
and used as an organocatalyst for the aerobic oxidation of primary and sec-
ondary alcohols in the presence of tert-butyl nitrite (TBN) as cocatalyst. The
reaction showed a general applicability to various alcohols, and the catalyst
was recovered easily and could be recycled for at least seven cycles with no
loss in catalytic activity.76 In a rare example of Fe/TEMPO catalyst systems,
Fe(NO3)3 or NaNO2 were not used; instead, the FeCl36H2O/TEMPO
combination was coupled with a silica support.77 In a related study, TEMPO
was covalently bound to silica and combined with FeCl3 6H2O/NaNO2.78
The TEMPO immobilization was performed via reductive amination of
4-oxo-TEMPO with amine-functionalized mesoporous silica SBA-15.79
A very efficient oxidation with a loading of 0.01 mol% of the expensive
TEMPO radical, with 8 mol% FeCl3 6H2O and 10 mol% NaNO2, in tol-
uene and O2 (1 atm) at 25 °C, was achieved. The iron salt and heteroge-
neous TEMPO could be recycled at least for five cycles. A continuous
flow approach employing a microreactor with a Fe–TEMPO system com-
prised of a heterogeneous iron oxide NP catalyst stabilized on a mesoporous
aluminosilicate support, was reported.80 A 42% yield of benzaldehyde is
achieved in a single pass of the reactor, but the reaction conditions were
rather harsh (120 °C and 35 bar of pure oxygen).

2.3 Prospective Substrates and Oxidation Agents


Finding or developing new starting materials for the aerobic and
peroxidative oxidation of alcohols is an important task, namely by widening
the range of substrates that are of significance in fine chemical synthesis. For
instance, since aldehydes are produced by the oxidation of alcohols, while
aldehydes by themselves can participate in further reactions, aminoxidation
of alcohols to nitriles and amide synthesis can be directly achieved using Mn
catalysts for both the transformations in one-pot (Scheme 6).81
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 107

Scheme 6 Proposed reaction pathway for the ammoxidation of alcohols over MnO2.81

Scheme 7 Oxidation of indole carbinols using the Fe(NO3)39H2O/TEMPO catalytic


system.82

Indole derivatives with carbonyl units (Scheme 7) have been found in


natural products, possess versatile bioactivity and are important intermedi-
ates in many organic syntheses. The introduction of the carbonyl moieties
to indoles significantly enhances their reactivity and can be achieved by oxi-
dation of indole carbinols with Fe(NO3)39H2O/TEMPO/NaCl at room
temperature and atmospheric pressure of dioxygen, using toluene as a sol-
vent.82 NaCl is an important accelerator for the oxidation, reducing the
reaction time from 15 to 2 h; however, the role of Cl is not quite clear.
Recently, there has been a growing interest in new processes to derive
chemicals and fuels from renewable carbon feedstocks, e.g., lignocellulosic
biomass (see below). Hence, the study of homogeneous oxovanadium and
copper catalysts toward aerobic oxidation of lignin model compounds is of
great potential.83 In this transformation, the vanadium catalyst affords pri-
marily ketone products, while the copper catalytic system leads to the prod-
ucts of CdC bond cleavage reactions, thus reflecting different mechanisms
of oxidation.
Glycerol is another substrate of importance in the biomass conversion,
and it can be transformed to the added-value dihydroxyacetone with a
supported palladium catalyst.84 A related approach is the added-value mod-
ification of nature-derived materials. For instance, pullulan, a polysaccharide
extracted from the fermentation medium of the Aureobasidium pullulans bac-
teria, can be functionalized by the introduction of various groups, such as
carboxylic ones, what significantly influences its hydrophilicity, etc.85
ARTICLE IN PRESS

108 Maximilian N. Kopylovich et al.

Vanadium was found to be active in the direct transformation of


4-pentenols to 3-acyloxy-γ-butyrolactones, using TBHP as oxidant
(Scheme 8).86 In this interesting oxidative conversion, stereocenters of
the tetrahydrofuran moiety retain their configuration. Allylic alcohols
can be also oxidized to stereodefined α,β-unsaturated aldehydes/
ketones with the retention of the C]C double-bond configuration, using
Fe(NO3)39H2O/TEMPO/NaCl, under atmospheric pressure of oxygen at
room temperature.87 The same catalyst was found to be effective in the con-
version of homopropargylic alcohols to the corresponding homopropargylic
ketones, which can be further isomerized to 1,2-allenic ketones.88 Iron cat-
alysts are also effective in the oxidative tandem assembly of 3-(2-oxoethyl)
indolin-2-ones from N-arylacrylamides and alcohols through the respective
1,2-difunctionalization of the C]C double bond in N-arylacrylamides
(Scheme 9).89
Until very recently, it was very difficult to prepare carbonyl compounds
from alcohols with strong electron-withdrawing groups adjacent to the
RCHOH moiety. For instance, 2,2,2-trifluoroethanol is often used as a sol-
vent in oxidation reactions since it has been considered inert to the oxida-
tion. However, the aerobic oxidation of “inert” perfluoro-substituted
alcohols to their corresponding carbonyl derivatives has been recently
achieved using Pt(II) complexes with dipyrido[3,2-a:20,30-c]-phenazine
ligands (Scheme 10).90 Thus, in the presence of H2SO4 and O2 oxidant,
trifluoroethanol was successfully oxidized to trifluoroethyl trifluoracetate
with >98% selectivity.

Scheme 8 Oxidation of 4-pentenols to 3-acyloxy-γ-butyrolactones.86

Scheme 9 Synthesis of 3-(2-oxoethyl)indolin-2-ones from N-arylacrylamides and


alcohols.89
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 109

Scheme 10 Aerobic oxidation of perfluoro-substituted alcohols.90

Scheme 11 Oxidation of imidazole diol to give imidazole-4,5-dicarbaldehyde.91

A bimetallic Pt/Bi/C catalyst was used for the oxidation of imidazole


diol to give imidazole-4,5-dicarbaldehyde, a key step in the synthesis of
pro-drugs of hepatitis C virus replicase inhibitors (Scheme 11).91
Introduction of new oxidative agents is also an important task since
“classical” oxidants, such as pyridinium chromates, are mostly not ecolog-
ical, produce high amounts of toxic agents and frequently are not enough
selective. The oxidations with eco-friendly air, oxygen, and peroxides, in
spite of being very attractive in theory and in small scale, encounter a num-
ber of difficulties when being scaled-up. In fact, oxidations in flammable
organic solvents will virtually eliminate such oxidants due to risk of inflam-
mation and other hazards related to high oxygen pressures. Apart from that,
in many instances these attractive oxidants do not operate or the reaction
goes through an undesired way, e.g., overoxidation or production of a num-
ber of by-products. Hence, the search for new effective oxidants is contin-
uously growing.
Rather selective and mild oxidizing agents, namely hypervalent halogen
compounds or their precursors (e.g., iodic acid, o-iodoxybenzoic acid
(IBX), NaBrO2, NaBrO3, etc.), have been widely introduced into the field
of alcohol oxidation during last few years.92 IBX, the less hazardous
2-iodosobenzoic acid (IBA) and even the commercially available
2-iodobenzoic acid (2IBAcid) (Fig. 8) in the presence of Oxone (potassium
peroxomonosulfate KHSO5), an environmentally acceptable reagent, as a
co-oxidant, were shown to be effective for the oxidation of many primary
and secondary alcohols in user- and eco-friendly solvent mixtures.93 In this
process, the reduced form of IBX, namely IBA is used in catalytic amounts
and is reoxidized with Oxone, in aqueous media, thus providing an attrac-
tive green protocol.
ARTICLE IN PRESS

110 Maximilian N. Kopylovich et al.

However, the application of IBX as oxidant can lead to some


unpredictable results, e.g., unexpected dehomologation of primary alcohols
to one-carbon shorter carboxylic acids (Scheme 12).94 In this case, the com-
bination of IBX and molecular iodine affords a different type of active
hypervalent iodine species, which was isolated and shown to be crucial
for the reaction. It should be noted that usually the dehomologation involves
complicated multistep procedures, whereas in the described oxidative strat-
egy it proceeds smoothly in one step. Another useful direct oxidation of sec-
ondary alcohols was performed using performic acid within just 15 min.95
The modification of IBX with a fluorous tag (IBXF, Fig. 9) allows its easy
recovery since insoluble fluorous IBA can be separated from the reaction
mixture by simple filtration, and be reused without significant loss of its
activity.96 Further, IBXF can be easily generated in situ from cheaper and
readily available Oxone. Other modifications of the oxidizing component
are being introduced. For instance, a combination of NH2OHHCl and
NaIO4 was recently proposed for the selective and mild oxidation of alco-
hols to the corresponding carbonyl compounds at room temperature.97
Concerning particular substrates, a wide range of β-hydroxyketones was

Figure 8 o-Iodoxybenzoic acid (IBX), 2-iodosobenzoic acid (IBA), and 2-iodobenzoic


acid (2IBAcid).

Scheme 12 Oxidative dehomologation of primary alcohols to one-carbon shorter car-


boxylic acids.94

Figure 9 Modified IBX with a fluorous substituent for its easy recovery.96
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 111

selectively oxidized in quantitative yields to β-diketones with IBX being a


superior oxidant for this transformation.98
The ability to perform oxidations without generating species harmful for
potential intermediates of further transformations is important to perform
multistep synthesis, such as the domino reactions described below
(Section 5). In this respect, the use of aryl halides as readily available, stable
and cheap oxidants (hydride acceptors) is a powerful option due to the pro-
duction of inert, dehalogenated aryl by-products in anaerobic conditions.
Commercially available Pd and Ni complexes with NHC ligands were
found to be active in a temperature-controlled domino oxidation/R-ketone
arylation with aryl halide.99
tert-Butylnitrite (t-BuONO) was recently introduced as a convenient
and easy-removable oxidant for an environmentally benign conversion of
primary and secondary benzylic alcohols to ketones and aldehydes, which
can be readily isolated by simple evaporation of the reaction mixtures since
t-BuONO decomposes giving only volatile side products.100 The oxidation
requires neither metal-based reagents nor organic catalysts and presumably
involves a nitrosyl exchange and a subsequent thermal decomposition of
benzylic nitrites. t-BuONO can potentially be recovered, since several alkyl
nitrites are known to be industrially produced from the corresponding alco-
hols and gaseous NO under an O2 atmosphere. A related approach was real-
ized for the simple, high-yield conversion of various achiral and chiral
alcohols to carbonyl compounds using TEMPO or AZADO in conjunction
with BF3OEt2 or LiBF4 as precatalysts and t-BuONO as oxidant.101
A NO+/NOpair was used for mild anaerobic nitroxide reoxidation, which
allowed the oxidation of enantiomerically pure substrates without
racemization.
K3[Fe(CN)6] was applied as a secondary oxidant in a chemoselective
osmium(VI)-catalyzed oxidation of benzylic, allylic, and propargylic alco-
hols.102 An uncommon oxidation agent, diisopropyl azodicarboxylate
(DIAD), can be used as an effective terminal oxidant103; in this case 1,2-diols
were oxidized to hydroxyl ketones or diketones depending on the amount
of DIAD used. Diaziridinone (Scheme 13) as oxidant allows the reactions to

Scheme 13 Diaziridinone as oxidant for acid- or base-sensitive substrates.104


ARTICLE IN PRESS

112 Maximilian N. Kopylovich et al.

be performed in neutral conditions and makes it compatible with acid- or


base-sensitive substrates.104 As a result, various acyclic and cyclic secondary
benzylic alcohols with alkenyl, alkynyl, thioether, silyl ether, amide, carba-
mate, ketal, ester, and heterocyclic moieties were effectively oxidized to
ketones in 73–99% yields, while no racemization of stereocenters occurred
during the oxidation.
A sure way of avoiding the toxic chemicals is the use of electric current as
oxidizing agent. Thus, an electrochemical process for selective oxidation of
1,2-diols to the corresponding α-hydroxyketones in water using
[Me2SnCl2] catalyst, KBr, and platinum electrodes has been introduced.105
The “Br+” ions, generated at the anode, are oxidants, while OH ions,
electro-generated at the cathode, play the role of a base. However, in this
synthetic strategy the toxicity of organotin catalysts is a drawback. This issue
was addressed in another study,106 where methylboronic acid [MeB(OH)2]
was used as a safe alternative. In this case, the selective oxidation of 1,2-diols
to the corresponding α-hydroxy ketones in aqueous medium most probably
occurs through the formation of boronate esters.

3. ACCEPTORLESS DEHYDROGENATIVE OXIDATIONS


The aerobic and peroxidative oxidations, described in the previous
sections, in spite of being very attractive in many aspects, can produce a con-
siderable amount of by-products. Other problems concern the over-
oxidation and risk of explosion due to the coexistence of oxygen and
organic reactants or solvents. In this respect, the oxidant- and acceptor-free
dehydrogenation (Scheme 14) is an alternative environmentally friendly
route for the conversion of alcohols into aldehydes or ketones, since gaseous
H2 can be easily separated from the reaction mixture. Moreover, this strategy
is much more attractive from the atom efficiency viewpoint and can provide
a promising route for H2 synthesis and storage.107
Thus, some advances in the oxidant- or acceptor-free oxidations based
on the catalytic dehydrogenation of alcohols accompanied by the evolution
of hydrogen gas will be described in this section. Some new attractive pro-
cesses were introduced recently in terms of efficiency and substrates, and
examples are described below.

Scheme 14 Oxidant- and acceptor-free dehydrogenation of alcohols.


ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 113

To date, the homogeneous catalytic systems for the dehydrogenative


oxidation of alcohols are mainly based on ruthenium,107,108 iridium,109
and, in much lesser extent, other transition-metal107 complexes and
metal-organic compounds. For instance, water-soluble Cp*Ir
(Cp* ¼ pentamethylcyclopentadienyl) complexes with α-hydro-
xypyridine109 or bipyridonate109 ligands have shown a high catalytic activity
in dehydrogenative oxidation of a wide variety of primary and secondary
alcohols and reversible dehydrogenation–hydrogenation between
2-propanol and acetone. The Ir catalysts are reusable109 and can accelerate
the selective dehydrogenation of biologically important and complex mol-
ecules, e.g., β-estradiol to give estrone (Scheme 15).109
Production of hydrogen gas can be also achieved with the Ir catalysts
from 2-propanol, thus providing a prototype for a hydrogen storage sys-
tem, based on the interconversion between 2-propanol and acetone
(Scheme 16).109 The reversible transformation with hydrogen storage
and evolution was repeated several times without loss of the catalytic activ-
ity. A ligand-promoted mechanism of dehydrogenation was proposed,

Scheme 15 Selective dehydrogenation of β-estradiol to estrone.109

Scheme 16 Interconversion between 2-propanol and acetone.109


ARTICLE IN PRESS

114 Maximilian N. Kopylovich et al.

acting the ligand as a proton acceptor in the activation step and as a


proton donor in the dehydrogenation step, thus playing a dual role
in cooperative catalysis.109
A detailed discussion of the mechanism of the pH-dependent alcohol
dehydrogenation in aqueous solution catalyzed by related [C,N] or [C,C]
cyclometalated Cp*Ir complexes was recently reported,110 generally
supporting the mechanistic suggestions and demonstrating a significant
dependence of the studied reaction system on pH.
Structurally related PCP-pincer Ir complexes (Fig. 10) were synthesized
by straightforward [4+2] cycloaddition and employed as catalysts in the
acceptorless dehydrogenation of alcohols.109 Such complexes can be easily
modified with a functional sidearm that is capable of interacting with the
catalytic site, thus making them suitable candidates for catalytic studies
involving ligand–metal cooperation. The H2 formation involves an intra-
molecular cooperation between the structurally remote functionality and
the metal center.109
A theoretical study on the mechanism of the acceptorless alcohol dehy-
drogenation mediated by the iridium catalyst [Cp*Ir(bpyO)] (bpyO ¼ α,α0 -
bipyridonate) suggests that the metal–ligand cooperative work involves
aromatization/dearomatization of the bpyO ligand.111 On the basis of
those results, the new ruthenium catalyst [(HMB)Ru(bpyO)(H2O)]
(HMB ¼ hexamethylbenzene) was designed (Fig. 11).111 Another ruthe-
nium catalyst formed from the Ru-hydride precursor [RuH2(PPh3)3CO]
and pincer PNP-ligands effectively promotes the dehydrogenative oxidation

Figure 10 PCP-pincer Ir complex.109

Figure 11 [(HMB)Ru(bpyO)(H2O)] ruthenium catalyst.111


ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 115

of alcohols with TOFs as high as 1.4  104 h1 after 20 min, at moderate
temperatures (<100 °C) and without an activation additive.111 The reaction
is an example of an efficient hydrogen production from a renewable alcohol
source. The proposed mechanism may also involve intramolecular con-
certed hydrogen loss from a dihydride Ru species followed by the outer-
sphere dehydrogenative step (Scheme 17).111
Ruthenium compounds also can be used for transfer hydrogenations.
Thus, a series of water-soluble Ru(II) half-sandwich complexes with
2,20 -bipyridine ligands behave as catalysts in the transfer hydrogenation of
different ketones to the corresponding alcohols, using a mixture of sodium
formate/formic acid as a hydrogen source.112 Interestingly, some obtained
alcohols can be selectively deuterated at the benzylic carbon, if the catalytic
transfer hydrogenation is performed in D2O (Scheme 18). This fact can be
explained by the fast reversible deuteration of the hydride intermediates

Scheme 17 Mechanism of alcohol–ketone interconversion coupled with the hydrogen


evolution step.111

Scheme 18 Catalytic transfer hydrogenation and deuteration of acetophenone.112


ARTICLE IN PRESS

116 Maximilian N. Kopylovich et al.

Ru(X2) (X ¼ H, D) in the presence of D2O and formic acid. The deutera-


tion is faster than the hydrogenation process; as a result, a selective deuter-
ation of phenylethanol at the benzylic carbon occurs. Using this principle,
selective deuteration of different unsaturated organic substrates can be
developed.112
The above mentioned ruthenium- or iridium-PNP type complexes cat-
alyze a one-pot dehydrogenation/aldol condensation/hydrogenation
sequence (Guerbet reaction), thus allowing to prepare β-alkylated
ketones.113 A dehydrogenative CdC bond formation occurs in this and
related cases114,115; the proposed catalytic sequence involves hydrogen
abstraction by cooperation of an acidic side arm with the metal-hydride site,
ligand exchange step with alkoxide formation, catalyst regeneration by
hydride elimination and formation of the carbonyl product; metal indepen-
dent aldol condensation; and finally, transfer of H2 to the aldol condensation
product to give the β-alkylated ketone (Scheme 19).
Substitution of noble metals in catalysts for cheap and abundant materials
is of obvious importance, and many works concern the preparation and
study of the latter systems. Thus, cobalt(II) alkyl complexes of aliphatic
PNP pincer ligands (Fig. 12) are active precatalysts for the hydrogenation
of ketones and the acceptorless dehydrogenation of alcohols under mild
conditions; in this case the alcohol dehydrogenation likely proceeds through
a cobalt(I)/(III) redox cycle.116,117

Scheme 19 Preparation of β-alkylated ketones via the dehydrogenation/aldol conden-


sation/hydrogenation sequence (Guerbet reaction).113
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 117

Figure 12 Cobalt(II) complex with a PNP pincer ligand as a catalyst precursor for accept-
orless dehydrogenation of alcohols.116,117

Scheme 20 Dehydrogenation of 2-pyridylmethanol using [CpFeCl(CO)2] as a


precatalyst.118

2-Pyridylmethanol derivatives can be effectively (with TONs up to


67,000) dehydrogenated in an acceptorless way using an iron catalyst,
e.g., derived from [CpFeCl(CO)2].118 A cocatalyst NaH is also used; its role
is possibly to provide the corresponding sodium alkoxide, which then reacts
with the above iron complex to give the active Fe-alkoxide complex
(Scheme 20). The attached pyridine ligand then displaces one of the CO
ligands, forming a Fe-N,O-metallacycle, from which an iron hydride com-
plex and 2-pyridinecarboxyaldehyde are produced by a β-hydride elimina-
tion. The oxidative addition of 2-pyridinylmethanol coupled with the H2
reductive elimination completes the catalytic cycle (Scheme 20).118 It should
be mentioned that this work is somehow related to the finding119 that in the
Fe-catalyzed oxidation of alkanes, pyrazinecarboxylic acid, and analogs have
a significant promoting effect.
In general, the iron catalysts seem to be very promising, and interesting
works on Fe-catalyzed acceptorless dehydrogenations of alcohols appear
ARTICLE IN PRESS

118 Maximilian N. Kopylovich et al.

every year. Thus, an operationally simple and economical process can be


mentioned, employing, as a catalyst, a mixture of readily available Fe(III)
acetylacetonate, 1,10-phenanthroline and K2CO3.120
In spite of generally being rather active, the above systems suffer from the
drawbacks common for homogeneous catalysts, and thus a number of het-
erogeneous catalysts have also been introduced.121 For instance, a non-
oxidative dehydrogenation of benzyl alcohol was achieved using a
hydrotalcite-supported gold catalyst.122 Nanocrystalline rhenium particles
concern another example of a reusable heterogeneous catalyst to convert
secondary and benzylic alcohols to ketones and aldehydes, through catalytic
acceptorless dehydrogenation via a novel γ-CH activation mechanism.123
Curiously, primary alcohols and aldehydes act as inhibitors of the dehydro-
genation reaction.
A heterogeneous rhodium-on-carbon system catalyzes the dehydroge-
nation of primary and secondary alcohols to the corresponding carboxylic
acids and ketones in water under basic conditions.124 Upon production
of the carboxylic acids, water played a role of an oxygen source. Pt
nanoclusters, supported on metal oxides, were also tested, and it was found
that the support, e.g., amphoteric alumina, plays an active role in the cata-
lytic process, reacting with alcohols and yielding an alkoxide and a pro-
ton.125 β-Hydrogen elimination from the alkoxide to a low coordinated
Pt0 site affords Pt-H and ketone; then protolysis of Pt-H by a neighboring
proton regenerates the Pt0 site with release of H2 gas.125
Nickel is a well-recognized promoter of hydrogen-transfer reactions and
hence it is expected that it can be applied in the reactions under discussion.
In accord, acceptor-free dehydrogenation of secondary alcohols by Ni NPs
supported on alumina has been reported.126 Low-coordinated Ni0 sites and
metal/support interfaces are believed to play significant roles in the catalytic
cycle.126
Nanocopper(0) on alumina [Cu(0)/Al2O3] NPs were prepared by a sim-
ple procedure from copper aluminum hydrotalcite and shown to be an effi-
cient catalyst for the acceptor- and oxidant-free dehydrogenation of alcohols
and amines.127
TiO2-supported Co NPs are also catalytically active in the acceptorless
dehydrogenation of various aliphatic secondary alcohols to the
corresponding ketones.128
The acceptorless dehydrogenation can be used not only for simple trans-
formation of alcohols to the corresponding ketones or aldehydes or for
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 119

hydrogen storage and production, but also as an essential part of more com-
plicated multistep procedures or tandem reactions. Moreover, apart from the
usual model alcohols, other uncommon alcohols are also being introduced
into the field and some of them are mentioned below. Thus,
dehydrogenative [Ir]-catalyzed coupling of arylhydrazines and alcohols
can be used for the selective synthesis of arylhydrazones, where no
N-alkylated by-products were generated (Scheme 21).129 This route is more
straightforward and potentially can compete with the traditional synthetic
ways to arylhydrazones based on condensation of arylhydrazines or
aryldiazonium salts with carbonyl compounds.
Dehydrogenative lactonization of diols is an efficient way to various lac-
tones (Scheme 22).130 The lactone formation is found to be catalyzed by a
recoverable stable dicationic iridium complex with 6,60 -dihydroxy-2,20 -
bipyridine ligands, and employs a variety of benzylic and aliphatic diols in
aqueous media. In comparison with the esterification of hydroxyl acids,
hydroacyloxylation of olefinic acids and Baeyer–Villiger reaction of cyclic
ketones, the dehydrogenative lactonization of diols proceeds without any
oxidant; hence, it is more environmentally benign and atom economical.
The transformation of alcohols to carboxylic acids with no oxidant or
hydrogen acceptor uses water as the oxygen atom source with concomitant
emission of dihydrogen gas.131 The reaction is catalyzed by a ruthenium
complex at a low loading (0.2 mol%) in basic aqueous solution
(Scheme 23).131 The same or related complexes are active in many other
tandem reactions which involve dehydrogenative oxidation; the proposed
mechanism of the catalysis involves a metal–ligand cooperation and both
O2 and H2 generation at a single metal center.132,133

Scheme 21 Direct synthesis of arylhydrazones through dehydrogenative [Ir]-catalyzed


coupling of arylhydrazines and alcohols.129

Scheme 22 Dehydrogenative lactonization of diols to lactones.130


ARTICLE IN PRESS

120 Maximilian N. Kopylovich et al.

Scheme 23 Conversion of alcohols to carboxylic acids with water as the oxygen source,
catalyzed by a Ru complex.131

4. OXIDATIVE DESYMMETRIZATIONS
The oxidation of alcohols is extensively used to introduce asymmetry
to starting compounds. For this, chemo-, regio- or stereospecific oxidations
are employed, e.g., upon the Oppenauer reaction, where the substrate is
oxidized by transfer of hydrogen atoms to a sacrificial ketone, such as ace-
tone. And vice versa, alcohols can be a source of hydrogen and hence a
reducing agent in asymmetric hydrogenations. For instance, asymmetric
transfer hydrogenation of ketones (Scheme 24) is an efficient and relatively
simple method to introduce asymmetry in a molecule.134
Other examples include OKR of racemic secondary alcohols
(Scheme 25A), oxidative desymmetrizations of meso-diols,136 etc. The
kinetic resolution is generally defined as a process where two enantiomers
of a racemic mixture are transformed to products at different rates. Thus,
one of the enantiomers of the racemate is selectively transformed to product,
whereas the other is left behind. This method allows to reach a maximum of
50% yield of the enantiopure remaining sec-alcohol. To overcome this lim-
itation, a modification of the method, namely dynamic kinetic resolution
(DKR), was introduced. In this case, the kinetic resolution method is com-
bined with a racemization process, where enantiomers are interconverted
while one of them is consumed (e.g., by esterification, Scheme 25B). There-
fore, a 100% theoretical yield of one enantiomer can be reached due to the
constant equilibrium shift. In most of the proposed DKR processes, several
catalytic systems, e.g., enzymes and transition-metal catalysts, work
together. Both reactions (transfer hydrogenation of ketones and the reverse
oxidation of secondary alcohols using ketone as a hydrogen acceptor) can be
promoted by a catalyst. The process can involve a temporary oxidation of a
substrate with hydrogen transfer to a transition-metal complex.
A classical example of catalytic desymmetrization is a regioselective oxi-
dation of polyoles, in particular 1,2-diols to form the corresponding
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 121

Scheme 24 An example of asymmetric transfer hydrogenation of ketones.134

Scheme 25 Oxidative kinetic resolution (A) and dynamic kinetic resolution (B) of race-
mic secondary alcohols.135

Scheme 26 Regioselective oxidation of 1,2-diols using organotin catalysts.136

α-hydroxyketones, versatile synthetic intermediates, and an important class


of biologically active products.136 Organotin compounds are renowned cat-
alysts in this transformation, which proceeds through the formation of
stannylene acetal as an intermediate (Scheme 26). Other metal catalysts,
e.g., palladium complexes with pyridine ligands137 also are used for the
chemoselective aerobic conversion of unprotected diols into the
corresponding hydroxy ketones. Pd–(NHC) complexes are catalytically
active as well; propene evolution was detected in this case, indicating that
the reaction most likely proceeds through a reductive elimination from a
palladium-allyl-hydride intermediate.138
A wide range of vicinal diols and polyols, including 1,2-diols, triols, and
tetraols, were oxidized selectively at the secondary alcohol to afford
α-hydroxy ketones, using chiral palladium complexes with pyridinyl
oxazoline derived ligands as catalysts and benzoquinone as the terminal oxi-
dant.139 The ligand geometry and environment have a significant influence
ARTICLE IN PRESS

122 Maximilian N. Kopylovich et al.

on the activity and chemoselectivity of the catalytic system. Neocuproine


can be also used as a ligand to create effective Pd catalysts for the
chemoselective oxidation of unprotected vicinal polyols under mild reaction
conditions.140 Oxidative lactonization of 1,5-diols to cyclic lactones and ste-
reospecific oxidation of (S,S)-1,2,3,4-tetrahydroxybutane [(S,S)-threitol] to
(S)-erythrulose (Scheme 27) were performed with these Pd-neocuproine
catalysts.
It is noteworthy that generally the selective oxidation of unprotected
polyols and in particular vicinal diols at the secondary alcohols is a difficult
task due to competitive CdC bond cleavage, overoxidation, poor
chemoselectivity, etc. Concerning the mechanism,140 the facile formation
of chelating protonated diolates from the vicinal diols leads to both higher
rates and chemoselectivity for oxidation of the secondary alcohol moiety
due to easy β-H elimination from the secondary Pd alkoxide. In contrast,
for mixtures of primary and secondary alkoxides, the formation of the pri-
mary Pd alkoxide is preferable and thus the selectivity for oxidation of pri-
mary alcohols to form aldehydes is observed in this case. The readily
available diamine ()-sparteine was also applied as a chiral ligand for Pd
to create an effective system for asymmetric catalysis, in particular kinetic
resolution of benzylic, allylic, and cyclopropyl secondary alcohols and des-
ymmetrization of meso-diols.141 A related oxidative desymmetrization of
meso-diols with a chiral Ir(Cp*) catalyst can be achieved with good effi-
ciency and enantiocontrol.134 In some cases, an oxidative lactonization
occurs if primary meso-diols are used as staring materials (Scheme 28).
Chiral N-sulfonyldiamine ligands are used to create effective chiral
bifunctional amidoiridium catalysts for the asymmetric aerobic oxidation
of meso- and prochiral diols to give up to >99% ee of hydroxyl ketones
and 50% ee of lactones.143 These catalysts can be also applied for an efficient
oxidative kinetic resolution of racemic secondary alcohols affording
R enantiomers with >99% ee and with 46–50% yields.135

Scheme 27 Stereospecific synthesis of (S)-erythrulose.140


ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 123

Scheme 28 Enantiospecific preparation of lactones by oxidative lactonization of pri-


mary meso-diols.142

Scheme 29 Chemoselective oxidation of a secondary alcohol moiety with Ru


catalysts.144

Chemoselective oxidation of a secondary alcohol moiety can be also per-


formed with ruthenium catalysts with phenylindenyl ligand.144 The selec-
tive oxidation of 1-phenylethanol to acetophenone from a mixture of
phenylethanol isomers, without oxidizing the other isomer, can then be
achieved (Scheme 29). In general, only the secondary alcohol moieties
are oxidized and the catalyst can be used for the chemical separation of iso-
mers or specific oxidation of highly functionalized molecules. The OKR of
unactivated racemic alcohols with dioxygen of air as the hydrogen acceptor
was effectively performed at room temperature with [(aqua)Ru(salen)] com-
plexes as catalysts.145
ARTICLE IN PRESS

124 Maximilian N. Kopylovich et al.

Scheme 30 Conversion of 2-diols to the corresponding chiral α-ketoalcohols.146

The use of cheaper, more available, and recyclable catalysts is a common


aim in the catalytic studies. Under this perspective, many recent publications
on the oxidative asymmetric resolution of alcohols involve 3d and other
abundant metals. Thus, in situ combination of copper(II) triflate and
(R,R)-Ph-BOX [BOX ¼ bis(oxazoline)] with NBS as an oxidant allowed
to achieve a desymmetrization of 2-diols to afford the corresponding chiral
α-ketoalcohols (Scheme 30).146
An air-stable, well-defined (NHC)–Ni0 complex is another effective
3d-metal catalyst for the mild anaerobic catalytic oxidation of secondary
alcohols with such functionalities as ether, tertiary amine, and alkene, using
nonanhydrous, -degassed 2,4-dichlorotoluene as both oxidant and sol-
vent.147 The use of inexpensive, stable, and ease to handle chlorinated sol-
vent as an oxidant makes this catalytic system attractive for multistep
synthesis and scaling-up; moreover, potentially harmful and/or difficult to
remove species are less likely to be formed. Additionally, primary benzylic
and alkylic alcohols are unreactive with this catalytic system and can be
recovered.
Jacobsen’s chiral Mn(III)-salen complexes constitute another example of
cheap and available catalysts for the enantioselective oxidation of racemic
secondary alcohols148; in this case, sodium hypochlorite (NaClO) was used
as oxidant. Related macrocyclic chiral Mn(III) salen complexes were applied
for the OKR of secondary alcohols with diacetoxyiodobenzene [PhI
(OAc)2] and NBS co-oxidants, in a biphasic dichloromethane-water solvent
mixture149; the catalyst can be easily recycled up to 7  without losing its
performance.
Transition metals can be eliminated from the catalytic systems. Thus, a
quinine-derived urea organocatalyst is effective in the enantioselective
oxidation of a wide range of diaryl-substituted meso-1,2-diols using bromi-
nation reagents as oxidants (Scheme 31).61 The method is simple, operates
at ambient temperature and utilizes available reagents to yield α-hydroxy
ketones in good yields (up to 94%) and enantioselectivities (up to 95% ee).
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 125

Scheme 31 Enantioselective oxidation of diaryl-substituted meso-1,2-diols.61

Moreover, both chemo- and enantioselectivities can synergistically contribute


to the specific oxidation of racemic diols bearing different substituents,
affording a hydroxy ketone as the sole product.61

5. CASCADE AND SEQUENTIAL REACTIONS


Some products of partial oxidation of alcohols, in particular aldehydes,
are widely used as starting materials in many organic transformations. Hence,
they potentially can react in situ with other components of the reaction
mixture, giving, e.g., alkenes, imines, or α-functionalized carbonyl com-
pounds. Selectivity issues will arise, but a proper choice of reaction condi-
tions, catalysts, and other additives would hopefully provide good yields and
selectivities. Sometimes the products of partial oxidation of alcohols can be
isolated from the reaction mixture after the oxidation step and transformed
further using the same catalyst.150 In this case, one deals with a sequential
transformation; an example is given by the conversion of aromatic alcohols
to the corresponding aldehydes using molecular oxygen and a copper–
TEMPO catalytic system,38e where the formed aldehydes can be isolated
and transformed further with the same copper catalyst to nitroalcohols upon
the nitroladol (Henry) reaction (Scheme 32).
If further transformations of the products of partial oxidation of alcohols
occur in one-pot, the term tandem or cascade reactions is generally used.151
Hydrogen atoms attributed to the starting alcohols can be combined with an
inorganic oxidant,152 leave the reaction environment, e.g., as H2,12 or be
transferred to another substrate, forming an added-value product. In the last
case, cheap and easy to operate alcohols are usually used as sacrificial reduc-
ing agents. At the same time, in many instances the hydrogen atoms can
ARTICLE IN PRESS

126 Maximilian N. Kopylovich et al.

Scheme 32 Sequential transformation of alcohols to aldehydes and further to


nitroalcohols.38e

O
Catalyst
R OH 1/2 R O R + H2
151
Scheme 33 Dehydrogenative coupling of alcohols to esters.

interact with an intermediate.151 Formally, a net oxidation of alcohols in this


case will not occur, but only a proton transfer. However, if one considers the
mechanism of such transformations, an oxidation (dehydrogenation) is an
essential rate-determining step. Hence, appropriate catalysts and specific
conditions of the alcohol oxidation (dehydrogenation) can serve as a starting
point in the search of new catalytic systems for the sequential or tandem
reactions of this type.
As mentioned above, hydrogen atoms, removed from the alcohol sub-
strate, can return to form the product; however, if the final hydrogenation
step could not occur, a product that is more oxidized than the starting mate-
rial is obtained. The formation of esters from alcohols and of amides from
alcohols and amines concern the most representative and studied reactions
of this type. In these cases, aldehydes, formed on the first oxidation stage
from alcohols, undergo Tishchenko- and Cannizzaro-type reactions, where
esters or carboxylates and alcohols are formed upon fusion or disproportion-
ation of aldehydes, respectively.
Thus, the dehydrogenative coupling of alcohols to esters with evolution
of H2 (Scheme 33) is one of possible effective variants of the tandem reac-
tions with net oxidation of alcohols.151 In contrast to the normal esterifica-
tion of an acid and alcohol, in which an equilibrium mixture is formed, the
evolved hydrogen, which is valuable by itself, can shift the equilibrium to
completion. Generally, the direct catalytic transformation of alcohols to
esters, without the use of the corresponding acid or acid derivative, is a very
attractive approach.
Ru(II) hydride complexes based on electron-rich PNP and PNN ligands
of the type depicted on Fig. 13 (which can undergo aromatization/
dearomatization steps) efficiently and selectively catalyze the acceptorless
dehydrogenation of primary alcohols to esters and H2 with high TONs
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 127

Figure 13 Ru(II) catalysts for acceptorless dehydrogenation of alcohols to esters.132

under relatively mild conditions.132 This provides a convenient method for


the synthesis of esters in view of its high efficiency, simplicity, and facile
product isolation. Thus, the reaction gave up to 99% of ester yield after
6 h at 115 °C, if started from 1-hexanol. Concerning the mechanism, it
was demonstrated that hemiacetal formation from the aldehyde and alcohol
followed by its dehydrogenation is more likely to occur than a Tischenko-
type reaction involving the aldehyde.132
Heterogeneous reusable catalysts, e.g., supported platinum-based ones,
can be utilized for the acceptor-free dehydrogenative coupling of alcohols
to esters under additive-free and solvent-free conditions at 180 °C, with iso-
lated yields within the 53–91% range.153 The activity depends on the sup-
port material and on the loaded transition metal. Thus, the SnO2 support
contains Lewis acid sites that activate carbonyl groups of adsorbed aldehyde
intermediates, while the Pt/SnO2 combination possesses the best promoting
activity among the studied transition metals, e.g., Ir, Re, Ru, Rh, Pd, Ag,
Co, Ni, and Cu loaded on SnO2.
Direct amide bond formation is a rather important transformation since
the amide functionality is a widely spread unit in synthetic intermediates,
pharmaceuticals, polymers, and natural products. Hence, its simple forma-
tion from cheap and convenient starting materials, such as alcohols, is of
interest. Generally, the amide bond construction is similar to the ester for-
mation, but the competing N-alkylation process significantly complicates
the picture (Scheme 34). Thus, the search for effective and selective catalysts
is of a clear need. Magnetic Fe3O4@EDTA–Cu(II) NPs can be such catalysts
if benzyl alcohols and amine hydrochloride salts are used as substrates with
TBHP as an oxidant. The corresponding amides are formed, while the cat-
alyst can be easily recovered by magnetic forces and reused several times
without loss of activity.154
To improve selectivity in the oxidation-addition cascades, the catalysts
and reducing agents should be able to maintain the catalytic cycle and
respond to the polarity differences between different radicals. Using this
approach, synthesis of side-chain-extended tetrahydrofurans from alkenols
and acceptor-substituted alkenes was achieved.155 Co-1,3-diketones were
ARTICLE IN PRESS

128 Maximilian N. Kopylovich et al.

Scheme 34 Competing processes in the dehydrogenative formation of amides from


alcohols and amines.154

used as catalysts to activate molecular oxygen for the oxidative cyclization


(Scheme 35); a sequence of polar and free-radical reactions is believed to
occur in this case.
Dehydrogenation of alcohols to aldehyde or ketone allows subsequent
bond construction steps which would not be possible for the parent alcohols.
Hence, a variety of iridium, rhodium or ruthenium phosphine, pincer and
related complexes, that are efficient catalysts for the dehydrogenation of
alcohols, can potentially be applied for the related hydrogen-transfer reac-
tions, thus leading to new added-value compounds.8 The hydrogen atoms
transfer to a sacrificial hydrogen acceptor, such as a carbonyl compound or
an olefin which is reduced to the corresponding alcohol or alkane.
Apart from the described examples, other interesting alcohol transforma-
tions can occur where different reactions are combined to give new and
sometimes unpredictable result. For instance, oxidation of one substrate
and reduction of another one can lead to the same product, thus favoring
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 129

Scheme 35 Alkenol oxidation—olefin addition cascade catalyzed by Co complexes.155

Scheme 36 Oxidation of methanol to formate salts.152

sustainability and profit, if the product is of a significant added value. The


process where methanol is effectively transformed into formate salts
(Scheme 36)152 can serve as an example of a reaction of this type. In this case,
catalytic methanol dehydrogenation is combined with bicarbonate hydroge-
nation giving high TONs (>18,000), TOFs (>1300 h1) and yields (>90%)
of formate salt. The bicarbonate behaves as a very convenient hydrogen
acceptor since the same product (formate) is formed from bicarbonate
hydrogenation and methanol dehydrogenation, while utilization of hazard-
ous gases and chemicals was avoided. The obtained formate salts are essential
chemicals with a variety of uses. The current industrial production of for-
mates involves absorption of hazardous, flammable, and difficult to transport
carbon monoxide under high pressure in solid sodium hydroxide at 160 °C.
A related direction concerns the elaboration of coupled systems, where
one substrate is oxidized into an added-value product, while another one is
reduced, also giving a product with added value. As an example, coupled
systems for selective oxidation of aromatic alcohols to aldehydes and reduc-
tion of nitrobenzene into aniline156 can be mentioned. This oxidation–
reduction coupling was realized using a cadmium-based composite material
as a photocatalyst under visible-light illumination, giving, in one-pot, 45%
and 26% yields of benzaldehyde and aniline, respectively.
ARTICLE IN PRESS

130 Maximilian N. Kopylovich et al.

6. CONVERSION OF RENEWABLE SOURCES


AND HYDROGEN PRODUCTION
6.1 Transformation of Renewable Materials into
Added-Value Compounds
Many renewable raw materials contain hydroxyl groups connected to car-
bon atom and hence can be considered as alcohols. One of the most abun-
dant renewable raw materials is lignocellulosic biomass, which is composed
of carbohydrate polymers such as cellulose or hemicellulose, and an aromatic
polymer lignin. The lignocellulose is mainly used in construction materials,
as firewood and for production of biofuels, such as biodiesel, ethanol, and
hydrogen, which are sustainable alternatives to fossil fuels with reduced
CO2 emission. However, modification and effective conversion of biomass
to fine chemicals with an added value is an active field of modern research.
For instance, partial oxidation of cellulose157 can lead to significant change in
its properties, e.g., ability to absorb water, what is important for many appli-
cations. Generally, the conversion of components of lignocellulose into
value-added chemicals is a complex process with many parallel reactions,
low yields and sometimes unpredictable results.
One of the current directions is the direct conversion of biomass to
formic or acetic acids.158 In this process, remarkable total yields up to
80% were achieved.159 Vanadium-substituted phosphomolybdic acids
(H3+nPVnMo12nO40) were used as catalysts, while molecular oxygen
was employed as an oxidant.159 The catalysts could be reusable at least
3  without significant loss of activity. Both formic and acetic acids are of
wide use in chemical, pharmaceutical, and agricultural industries; formic
acid is also recently evaluated as a perspective carrier to store or generate
H2.160,161
Other important added-value products, such as levulinic acid,162,163
sorbital,164 ethylene glycol,165 5-hydroxymethylfurfuran,155 lactic,166
glycolic,167 and gluconic168 acids can also be prepared by selective conver-
sion of cellulose and biomass-derived carbohydrates. Different conditions
were proposed to optimize the yields and selectivities; the use of supercritical
conditions or various catalysts and additives are among the most used pro-
cedures to increase effectiveness of these conversions. For instance, Keggin-
type heteropolyacid catalysts H5PV2Mo10O40 can effectively accelerate the
conversion of mono-, olygo-, and polysaccharides to formic acid with good
yields.152 Similar phosphomolybdic acid H3PMo12O40 was used as a
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 131

bifunctional catalyst to accelerate both the hydrolysis of cellulose and the


subsequent oxidation reactions to give glycolic acid with molecular oxygen
in a water medium.167
Conversion of substrates, readily available from natural sources, e.g.,
glycerol (a cheap by-product of biodiesel production), furans, or carbohy-
drates, is another direction which is under active exploration.84,169 The liq-
uid phase oxidation of glycerol with bimetallic Au/Pd and Au/Pt catalysts
supported over MgO leads to an enhanced glycerol conversion and
increased selectivity toward oxidized C3 products under mild conditions
and without the addition of a base.170 Another tested bimetallic catalyst
allows proceeding with a selective oxidation of biomass alcohols to the
corresponding aldehydes by a visible-light-driven synergistic photo-
electrochemical (PEC) catalysis system with Au/CeO2–TiO2 nanotubes
(NTs) as photocathodes and using mild conditions.171 The obtained conver-
sion of benzyl alcohols was up to 98% while the selectivity toward benzal-
dehyde was >99%.
Along with conventional catalysts, thermostable enzymes are prospective
agents to convert glycerol into materials with high added values, e.g., dihy-
droxyacetone, a synthetic precursor and sunless tanning agent. Using glyc-
erol as a second substrate, (R)-1,2-propanediol can be also produced from
hydroxyacetone in a one-enzyme bioelectrocatalytic reactor.172 In this
study, NAD-dependent Thermotoga maritime glycerol dehydrogenase
(TmGlyDH) was employed as a main catalytic agent, while the
NAD(H)-cofactor can be immobilized and regenerated electrochemically.
It was also demonstrated that TmGlyDH can be a useful catalyst for produc-
ing optically active products, e.g., for resolving a racemic mixture of 1,2-
propanediol leaving the (S)-enantiomer aside.
Other important and available derivatives of biomass are sugars, and their
conversion and application are also of a high recent interest.163,164 For
instance, novel magnetically separable carbonaceous nanohybrids were
recently prepared from porous starch.173 These porous polysaccharide-
derived materials are highly magnetic up to 450 °C and thus can be used
as recyclable catalysts, e.g., oxidations of benzyl alcohols and xylose dehy-
dration to furfural.
Gold-based materials were shown to be more active than other metal
catalysts in the first step of the 5-hydroxymethylfurfural (HMF) oxidation,
leading to 5-hydroxymethyl-2-furancarboxylic acid (HFCA) very quickly,
even though they showed less activity for the subsequent conversion of
HFCA to 2,5-furandicarboxylic acid (FDCA).174 The HMF oxidation over
ARTICLE IN PRESS

132 Maximilian N. Kopylovich et al.

Au and Pt catalysts in the presence of high amounts of NaOH was recently


investigated with the use of isotopically labeled dioxygen and water. The
source of inserted oxygen was shown to be water rather than oxygen.
Unfortunately, in all the studied gold-based samples, process efficiency
and catalyst stability were rather low. However, the modification of
Au-based catalysts with Pt or Pd metal produced stable and recyclable cat-
alysts.175 Bimetallic Au8Pd2 species supported over active carbon have the
highest activity and stability for the production of FDCA.

6.2 Alcohol Oxidation for Hydrogen Storage and Production


The reserves of fossil fuels are rapidly depleting, while the pollution caused
by their intensive exploitation constitutes another challenge to overcome.
Among the alternative and clean fuels, hydrogen is one of the most prom-
ising ones since it can be produced from renewable resources, has a high
energy density, zero carbon emission, etc. However, other physical proper-
ties of H2 make its handling a rather difficult task. Therefore, easy-to-handle
hydrogen-containing sources should be introduced, for instance, alcohols
which can be used to store and transport H2. Thus, methanol contains
12.6% hydrogen, but the hydrogen production from alcohols generally
involves rather costly oxidative steam reforming at high temperatures (over
200 °C) and pressures (25–50 bar) and with noble metal catalysts to obtain
high purity hydrogen. Furthermore, these methods usually generate green-
house gases, while the evolved CO itself may result in poisoning the
involved catalysts. Alcohol dehydrogenation in the presence of a hydrogen
acceptor (e.g., O2, H2O2, acetone) concerns another perspective, but the
application of the hydrogen acceptor contradicts the atom economy and
hence efficient acceptorless alcohol-dehydrogenation protocols are of
clear need.
Many molecular catalysts have been applied for the hydrogen production
from alcohols,12,176 ruthenium and rhodium ones being particularly effec-
tive in catalytic acceptorless dehydrogenations.107 Thus, an efficient low-
temperature aqueous-phase methanol dehydrogenation process, which is
facilitated by ruthenium complexes with pincer-type ligands, e.g.,
[RuHCl(CO)(HN(C2H4PiPr2)2)], has been described.177 Hydrogen gener-
ation by this method proceeds at 65–95 °C and ambient pressure with cat-
alyst TOFs up to 4700 h1 and TONs above 350,000; furthermore, no base
is needed. It was demonstrated for the first time that it is possible to effi-
ciently dehydrogenate the thermodynamically less-favorable primary
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 133

aliphatic alcohols below 100 °C. This eventually could make the use of
methanol as a practical hydrogen carrier a feasible task.
Concerning heterogeneous catalytic systems, Rh is an active metal for
the ethanol steam reforming to produce H2.178,179 In the reforming the sup-
ports also play an important role regarding the activity, selectivity and sta-
bility of the catalyst, and MgAl2O4 was found to be an appropriate
support for the Rh catalysts.180 This Rh/MgAl2O4 system is believed to
be a bifunctional catalyst,177 whereas the activation of ethanol takes place
both on the metal particle and on the support basic and acidic sites leading
to the formation of intermediate compounds (Scheme 37). Ethylene can also
be formed giving a lower hydrogen production [Scheme 37 (3)] due to deac-
tivation of the acidic sites with coke formation on the support.177
Primary alcohols, like ethanol, can be obtained from biomass fermenta-
tion, but their further transformations toward other carriers with higher
energy density are still under development. However, a simple but effective
non-catalytic way for the production of high purity hydrogen from primary
alcohols under basic conditions has been recently reported.181 At this reac-
tion, one mole of ethanol reacts with one mole of a base (e.g., NaOH,
Scheme 38) giving two moles of H2 and one mole of sodium acetate.
One of the main advantages of this approach is that no catalyst is required,
and no environmentally harmful gas, such as CO or CO2, is produced in the
process. The authors also report that temperature is a key factor affecting the
rate of gas generation: as long as the temperature is higher than 120 °C,
hydrogen and carboxylate are produced.
If methanol is used as a starting material, the corresponding formate salts
can be produced apart from H2. Since formates, in their turn, have been
evaluated as potent carriers to store and produce hydrogen,160,161 addition-
ally giving valuable oxalates or carbonates, the overall process has a high

Scheme 37 Ethanol steam reforming main reaction (1) and intermediate ethanol dehy-
drogenation (2) and undesirable ethanol dehydration (3).

Scheme 38 Production of hydrogen from ethanol.181


ARTICLE IN PRESS

134 Maximilian N. Kopylovich et al.

Scheme 39 Proposed177 mechanism for H2 generation with ethanol as substrate.

potential to become a large-scale industrial process. The water tolerance of


the reaction was also tested, and it was found that high water content inhibits
the hydrogen production. In spite of that, the normal bioethanol can be used
to generate high purity hydrogen at a relatively low temperature; although
the rate of the reaction might be lowered, a moderate extension of the reac-
tion time would overcome this obstacle.
The proposed177 mechanism of this transformation (Scheme 39), as
supported by GC/MS and isotope labeling studies, involves removal of
the hydroxide proton of ethanol by a base, thus giving a molecule of water
and ethanolate (Step 1). This anion rapidly reacts with water to give H2,
acetaldehyde, and hydroxide ion (Step 2). Reaction of the hydroxide ion
with the carbonyl carbon of acetaldehyde (via nucleophilic addition) forms
alkoxide which is then deprotonated and gives a dianionic Cannizzaro inter-
mediate, that subsequently reacts with another molecule of acetaldehyde to
give the final product (sodium acetate) and a molecule of ethanol (Step 3).

7. IRRADIATION-PROMOTED OXIDATIONS
Acceleration and promotion of chemical reactions by irradiation with
radiowaves of various frequency ranges is an established field with multiple
applications in laboratory practice and in industry. However, new and inter-
esting results in this topic continue to appear. The range of radiowave fre-
quencies is expanding and now includes virtually all the spectrum starting
from gamma-rays and down to the low-energy MW irradiation. Although
there is some consensus on the key steps of photochemical reactions pro-
moted by short-waves, e.g., UV and visible-light irradiations, it is still debat-
able the mechanism by which low-energy irradiation, such as MWs,
influence the reaction kinetics.182 In this section, we shall not discuss the
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 135

mechanisms of irradiation action (there are plenty of books and reviews on


that matter, e.g., see Refs. 182c,d and references therein), but focus on some
recent applications of irradiation in alcohol oxidation.

7.1 Photocatalytic Oxidations


Photochemical reactions induced by ultraviolet and visible-light irradiations
mainly involve transformations of molecules by direct absorption of light. In
these cases, the energy of the UV/vis photon is significantly higher than the
energy of Brownian motion and is generally high enough to directly cleave
molecular bonds. Application of photocatalytic and photoelectrocatalytic
methods for partial oxidations of both aliphatic and aromatic alcohols to give
aldehydes and other related products, mainly using titanium oxide as catalyst,
have been reported by several groups.183,184
Due to its direct and in many cases specific action toward certain chem-
ical bonds, many irradiation-induced processes are more sustainable, con-
sume less energy, generate less by-products, or can be conducted using
greener solvents than analogous conventional reactions. On the other hand,
sometimes the application of irradiation is crucial to perform a specific reac-
tion, which cannot be conducted with a reasonable rate or would not occur
at all, if the irradiation was not applied. For instance, piperonal, a compound
of great importance for cosmetics, agrochemical, and pharmaceutical chem-
istries, is traditionally synthesized by isomerization and subsequent oxidation
of safrole, an ingredient of some rather expensive and rear essential oils.
Other proposed routes involve harmful reagents or environmentally unsafe
heavy metals, while selective photocatalytic oxidation (Scheme 40) allows to
use piperonyl alcohol, a common chemical which is approximately 500 
less expensive than piperonal, as a starting material, under organic-free aque-
ous conditions, using UV irradiation and cheap commercial TiO2 as a
photocatalyst.184
The photocatalytic oxidation of aromatic alcohols to aldehydes in water
with rutile and anatase TiO2 NPs under UV light irradiation was stud-
ied.185,186 For example, photooxidation of four-substituted aromatic alco-
hols to the corresponding aldehydes, over rutile TiO2 NPs, showed

Scheme 40 Photocatalytic synthesis of piperonal from piperonyl alcohol.184


ARTICLE IN PRESS

136 Maximilian N. Kopylovich et al.

45–74% of selectivity.185 The effect of surface and physical properties of


TiO2 NPs on the selective photocatalytic activity under UV light irradiation
was also investigated. Photocatalytic selective oxidation of ethanol to ace-
taldedyde in a fluidized bed photoreactor was studied on new structured
photocatalysts based on direct supporting the VOx/TiO2 on the surface
of commercial ZnS-based phosphors.187
Oxidation of benzylalcohol to benzaldehyde with yields and selectivity
values higher than 40% and 80%, respectively, was reported using ferric ions
as homogeneous catalysts and oxygen as an oxidant under UV-solar simu-
lated radiation. To avoid occurrence of side reactions, in consequence of
generated undesired reactive OH radicals188 due to the possible Fe(III)
aquo-complexes photolysis, reactions were carried out at pH 0.5.189
Selective oxidation of benzyl alcohol to benzaldehyde and reduction of
nitrobenzene into aniline, under visible-light illumination, was also reported
using CdS/g-C3N4 composite as a photocatalyst.156 The conversion of the
alcohol into the aldehyde was achieved by direct holes oxidation, and
the reduction of nitrobenzene into aniline by direct electrons reduction.
The CdS/g-C3N4 photocatalyst exhibits enhanced photocatalytic activity
and excellent photostability relatively to single g-C3N4 and CdS, with an
optimum percentage of CdS of ca. 10 wt.%. Under irradiation of visible light
(420 nm) for 4 h under N2 purge conditions, the yields of benzaldehyde and
aniline are 44.6% and 26.0%, respectively.156
Recently, the first example of a MOF as a promising visible-
light photocatalyst toward the selective oxidation of alcohols to
their corresponding aldehydes was reported. The Zr-based MOFs,
Zr-benzenedicarboxylate (UiO-66) and its derivative Zr-2-NH2-benzene-
dicarboxylate (UiO-66(NH2)) were prepared via a solvothermal method
and successfully applied to photocatalytic reactions.190 A reaction mecha-
nism, involving the production of charge carriers and photogenerated elec-
trons and holes, was proposed. Oxygen reacts with an electron to form O2  ,
while the photogenerated holes can directly oxidize the organic reactive
substrates to carbonium ions. The formed superoxide radicals further react
with the carbocations, which leads to the final products.187 A recent devel-
opment in this field concerns the construction of a photoreactor with mem-
brane separation, namely pervaporation, in order to prevent overoxidation
of the aldehyde.
A significant rate acceleration was observed in the copper-catalyzed
aerobic photooxidation of sugars using a Cu–TEMPO catalytic system
(Scheme 41).191 It was suggested that the transformation of the
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 137

Scheme 41 Synthesis of tetrahydroazapanes through light-activated Cu/TEMPO-


catalyzed aerobic alcohol oxidation.191

Scheme 42 Plasmon-mediated oxidation of sec-phenethyl (R ¼ CH3) and benzyl (R ¼ H)


alcohols in the presence of supported AuNP.192

Cu(II)–TEMPO intermediate is induced photochemically. Curiously, this


reaction is highly selective for primary alcohols, and hence
tetrahydroxyazepanes can be easily synthesized via specific oxidation of a
benzyl glucoside, followed by reductive amination and nucleophilic ring
expansion (Scheme 41).
Recent technological advances and price drop in laser and light emitting
diode (LED) production allow extensive studies on how an irradiation of a
certain wavelength influences a specific chemical reaction. Thus, a compar-
ative study of laser, LED, and MW irradiations in both sec-phenethyl and
benzyl alcohol oxidations to acetophenone and benzaldehyde, respectively,
was performed.192 Excitation with monochromatic 530 nm LEDs
(Scheme 42) gave yields as good or better than the corresponding laser
and MW techniques, with a maximum conversion of 95% after 20 min.
Hence, LEDs can provide a new economical and power-efficient alternative
ARTICLE IN PRESS

138 Maximilian N. Kopylovich et al.

Scheme 43 Possible pathways in the plasmon excitation of supported metallic


nanoparticles.192

light source for plasmon-mediated reactions, where Au metal nanoparticles


(AuNPs), with a strong absorption of the surface plasmon band (SPB) within
the visible region, favor ejection of electrons and induce a variety of photo-
initiated electron transfer pathways (Scheme 43).
A highly efficient and selective catalytic oxidation of biomass alcohols to
the corresponding aldehydes was shown to be driven by a synergistic action
of visible-light and electrochemical catalysis employing Au/CeO2–TiO2
NTs as photocathodes, under mild conditions.171 At the bias potential of
0.8 V and under visible-light irradiation for 8 h, the conversion of benzyl
alcohol was 98% with the selectivity toward the benzaldehyde formation
being >99%.

7.2 MW-Promoted Oxidations


The use of MW irradiation with a lower energy than UV–vis light has
attracted much attention in synthetic organic chemistry due to the improve-
ments on efficiency, rates of reactions, and energy consumption, as well as
on selectivities, what has positioned this technology as a useful alternative
energy source in organic synthesis, with an environmentally friendly nature.
The MW-enhanced chemistry is based on the efficiency of the interac-
tion of molecules with electromagnetic waves generated by a “microwave
dielectric effect”. This process mainly depends on the ability of a specific
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 139

mixture (substrates, catalyst, and solvents) to absorb MW energy and convert


it into heat. Polar molecules have good potential to absorb MWs and con-
vert them to heat energy, thus accelerating the reactions compared with the
conventional heating. The ability of a specific material (e.g., a substrate or a
solvent) to convert electromagnetic energy into heat is known as loss tan-
gent, tan δ. A reaction medium with a high tan δ is required for efficient
absorption and consequently rapid heating. Despite alcohols and most of
organic compounds have a considerably lower dielectric constant compared
to water, they heat much rapidly in a MW field on account of their high tan
δ. Furthermore, polar components (such as ILs) can be added to increase the
absorbance of a reaction medium. However, the nature of the MW effect is
still debatable and in some cases it has been proved that it concerns a heating
effect.182 External infrared (IR) temperature controllers mainly used in
MW-assisted homogeneous reactions do not accurately monitor the sample
temperature and usually tend to understate its value.182
There are many examples of the successful application of MW-assisted
chemistry to organic synthesis; these include the use of benign reaction
media, solvent-free conditions, and application of solid supported and reus-
able catalysts. Over the past few years, it was demonstrated that many
transition-metal-catalyzed bond transformations can be significantly
enhanced by employing MW heating under sealed-vessel conditions, in
most cases without requiring an inert atmosphere.
Recently, a MW-promoted procedure for one-pot, two-step conver-
sion of aryl alcohols to aryl fluorides via aryl nonafluorobutylsulfonates
(ArONf ) was reported (Scheme 44).193 Moderate to good yields were
achieved by this MW-assisted palladium-catalyzed fluorination sequence.
The in situ conversion of aryl alcohols to aryl nonaflates using CsF as base,
followed by the MW heated (180 °C) fluorination, catalyzed by [Pd2(dba)3]
(dba ¼ dibenzylideneacetone) and t-BuBrettPhos [2-(di-tert-butylphosphino)-
20 ,40 ,60 -triisopropyl-3,6-dimethoxy-1,10 -biphenyl], allowed full conversion
after 30–60 min reaction.191

Scheme 44 MW-assisted fluorination of aryl triflates.191


ARTICLE IN PRESS

140 Maximilian N. Kopylovich et al.

Rhodium- and ruthenium-catalyzed hydrogen-transfer type oxidation


of secondary alcohols (e.g., 5-tetradecanol, cyclododecanol, cyclooctanol,
4-t-butylcyclohexanol, or 1-p-tolyl-1-hexanol) lead, in moderately to
excellent yields, to the corresponding ketones by employing MW heating
at 140 °C for 15 min, using 2 equiv. of methyl acrylate as hydrogen acceptor
and 5 mol% of [RhCl(CO)(PPh3)2] as catalyst, in a water/N,N-DMF sol-
vent mixture (Scheme 45).194 No conversion was observed in the absence
of MW irradiation. Primary alcohols, such as n-heptanol, n-tridecanol, or
1,7-heptanediol, were oxidized using a 2.5 mol% of [RuCl2(PPh3)2] and
2 equiv. of methyl vinyl ketone under solvent-free conditions and MW
heating at 120 °C for 15 min (Scheme 46).194
The accelerating effect of MW irradiation in the synthesis of ketones
from secondary alcohols with TBHP as oxidant was also reported.26,38,40,64
In fact, in the presence of the dicopper(II) [Cu2(Hedea)2(N3)2](0.25H2O)
(Hedea ¼ N-ethyldiethanolamine) complex with the {Cu2(μ–O)2} die-
thanolaminate core, the oxidation of 1-phenylethanol is dramatically accel-
erated when the reaction mixture is subject to MW irradiation, achieving a
very high yield of acetophenone (91%) after 15 min of reaction, in contrast
with the 51% acetophenone formed after 30 min when using conventional
heating.26
Several recent examples use copper(II) or Cu(II)/TEMPO catalytic
systems,38 namely alkoxy-1,3,5-triazapentadien(e/ato) copper(II) com-
plexes (yields up to 97% and TONs up to 485 after 60 min, and TOFs of
up to 1170 h1 after 10 min reaction)195 or bis- and tris-pyridyl amino
and imino thioether Cu and Fe complexes, with a maximum yield of
acetophenone of 99% after 30 min at 80 °C. The maximum TOF of
5220 h1 (corresponding to 87% yield) was achieved just after 5 min of reac-
tion time under the low MW power of 10 W.42,43

Scheme 45 MW-assisted oxidation of secondary alcohols with a Rh(I) catalyst and


methyl acrylate.194

Scheme 46 MW-assisted oxidation of primary alcohols with a Ru(II) catalyst and methyl
vinyl ketone.194
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 141

In a related study, the same group reported a series of mixed-ligand din-


uclear manganese(II) Schiff base complexes as catalysts for the MW-assisted
oxidation of alcohols (Scheme 47).64 Acetophenone yield of 81% is
obtained using a maximum of 0.4% molar ratio of [Mn(H2L)–
(py)2]2(NO3)22CH3OH (H2L ¼ hydrazone Schiff base) catalyst relative to
the substrate (1-phenylethanol) in the presence of TEMPO and in aqueous
basic solution, under mild conditions.64b
A recent publication has described the efficient ruthenium-catalyzed C-3
reductive alkylation of 4-hydroxycoumarin by dehydrogenative oxidation
of benzylic alcohols.196 The optimization of reaction parameters, such as
type of catalyst, type of solvent, activation method, reaction time, temper-
ature, and base, was performed (Scheme 48). Under optimized conditions,
using [RuCl2(PPh3)3] as catalyst in tert-amyl alcohol under MW irradiation
at 140 °C for 2 h, afforded a satisfying selectivity/conversion.
MW irradiation also permitted a shorter reaction time for the selective
solvent-free oxidation of primary, secondary, allylic, and benzylic alcohols
with pyridinum sulfonate chlorochromate and pyridinum sulfonate fluoro-
chromate as oxidizing agents, compared with the use of solvent. For exam-
ple, cholest-5-en-3-ol acetate and cholest-5-en-3-ol benzoate were
chemoselectively oxidized at position 7 in solvent-free conditions with
82% and 87% conversion, respectively.197
A CrO3-catalyzed oxidation of homopropargyl alcohols with TBHP
under MW irradiation was found to be an efficient and rapid alternative

Scheme 47 MW-assisted solvent-free oxidation of 1-phenylethanol to acetophenone


with a Mn catalyst.64

Scheme 48 MW-assisted selective C-3 alkylation of 4-hydroxycoumarin with benzyl


alcohol.196
ARTICLE IN PRESS

142 Maximilian N. Kopylovich et al.

for the preparation of 1,2-allenic ketones. The optimal conditions were


achieved when a solution of 1-phenylbut-3-yn-1-ol in CH2Cl2 was irradi-
ated with MW at 40 °C for 0.2 h in the presence of 5 mol% of CrO3 as cat-
alyst and 3 equiv. of TBHP as oxidant.198
Tetrabutylammonium decatungstate(VI) was reported to possess a cata-
lytic activity in the oxidation of selected alcohols with hydrogen peroxide as
oxidant, in 1,2-dichloroethane/water or acetonitrile/water solvent systems.
A pronounced accelerating effect on the reaction rate was observed when a
MW conditions were used.199
A highly active (NHC)-Pd catalytic system can be applied for anaerobic
oxidation of secondary alcohols at very mild temperatures. This procedure
allows assessing one-pot domino procedures for the synthesis of R-arylated
ketones from secondary aryl alcohols with very good yields.200 Recently, a
successful use of MW heating for the synthesis of [Pd(acac)Cl(NHC)] and
[PdCl2(3-Cl-pyridine)(NHC)] complexes was reported.201 This protocol
affords the desired compounds in yields comparable to those obtained using
conventional heating, but drastically reduces the reaction times. A similar
protocol was thereafter applied to the synthesis of a series of
[Ni(Cp)Cl(NHC)] complexes (Scheme 49).202 These complexes where
applied in the catalytic anaerobic oxidation of alcohols using 2,4-
chlorotoluene as solvent and oxidant.202
In recent years, the use of a room-temperature IL in MW-assisted syn-
thesis, as a (co-) solvent and/or (co-)catalyst, is becoming an increasingly
exploited area since the ionic nature of ILs allows them to absorb MW
energy very efficiently. Besides that, ILs exhibit several inherent benefits,
such as low vapor pressures, high thermal and chemical stability, and
nonflammability. The role played by the combination of an IL and MW
was explored in the activation of H2O2 in [hmim]Br (hmim ¼
3-methylimidazolium) used as catalyst and solvent, under MW irradiation,
for the metal-free chemoselective oxidation of various alcohols into the
corresponding carbonyl compounds.203 In addition, a new metal-free meth-
odology for the synthesis of anthraquinone has been reported.203

Scheme 49 MW-assisted synthesis of [Ni(Cp)Cl(NHC)] complexes, catalysts for alcohol


oxidation.202
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 143

A Co(II) complex supported on SBA-15 has been employed as a highly


active and reusable catalyst in the selective oxidation of various alcohols, in
which the improved rates of reactions (from hours to minutes), yields and
even selectivities in some cases, illustrate the usefulness of MW protocols
as alternative methodologies in organic synthesis.204 A MW-assisted
solvent-free peroxidative oxidation of benzyl alcohol to benzaldehyde cat-
alyzed by magnetic Ni-doped MgFe2O4 NPs was also successfully per-
formed.205 The catalyst is reusable and eco-friendly, is applied in a small
amount and the reaction time is short.205

7.3 Others
Application of ultrasounds is among the new alternative techniques which
can accelerate heterogeneous reactions by increasing the surface (dispersity)
of the reagents, while in homogeneous systems ultrasounds assist the even
heat distribution in a reactor.206 Thus, the application of ultrasounds or
of MW irradiation may substantially shorten the reaction time in oxidations
of alcohols with PCC, from hours to minutes.207 It is believed that the ultra-
sound produces erosion on the PCC surface and therefore accelerates its
interaction with the organic substrates. The involvement of ultrasonic irra-
diation in a drastic reduction on the amount of PCC used was also
indicated.207
An ongoing approach on the development of new environmentally
friendly protocols involves the combination of MW and ultrasonic irradia-
tion in ILs. In this context, MW and ultrasound activation methods have
been used in the oxidation of five- to eight-membered cyclanols in the pres-
ence of H2O2/H2WO4 and several hydrophobic ILs as cocatalysts.208 Quan-
titative oxidation of cyclohexanol, after only 30 min at 90 °C, was achieved
in the presence of [Aliquat][NTf2] (Aliquat ¼ N-methyl-N,N-
dioctyloctan-1-ammonium; Ntf2 ¼ bis(trifluoromethylsulfonyl)imide), pre-
pared from Aliquat 336 (Scheme 50) which is well known as a phase transfer

Scheme 50 Aliquat 336 (A) and [Aliquat][NTf2] (B).


ARTICLE IN PRESS

144 Maximilian N. Kopylovich et al.

Scheme 51 Ultrasound- and microwave-assisted one-pot oxidation of cyclohexanol to


ε-caprolactone in [bmim][BF4] ionic liquid.209

agent. The aliphatic cation of Aliquat is more efficient than the cyclic cations
of other tested ILs.208
Another example of application of MW and ultrasound methods is the
one-pot, tandem oxidation of cyclic alcohols to their respective lactones
using KHSO5 (potassium peroxy-monosulfate) as oxidant and an IL as a sol-
vent (Scheme 51).209 Ultrasound and MW irradiation reduced the reaction
time for the cyclohexanol oxidation by Oxone®, catalyzed by a TEMPO
nitroxyl radical, in the presence of tetrabutylammonium bromide
(TBAB) in [bmim][BF4] (bmim ¼ 1-butyl-3-methylimidazolium), from 8,
using normal heating, to 5 and 0.5 h, respectively, with similar yields of
ca. 80%. A new class of ILs with peroxymonosulphate anions was also syn-
thesized and employed in the model oxidation.209
As illustrated above, the activation of substrates, intermediates, and other
components of reaction mixtures by using irradiation of different kinds can
efficiently influence the reaction kinetics in the oxidation of alcohols.
Further application of such techniques should widen the spectrum of used
substrates and obtained products in this field.

8. CATALYSTS RECYCLIZATION
Many homogeneous catalytic systems, in spite of being very active and
interesting from a fundamental point of view, frequently cannot lead to a
practical solution of technological problems due to their high cost, instabil-
ity, and difficulty of isolation and recyclization. Hence, efforts have been
devoted to obtain recyclable catalytic systems, which, even if being less
active than state-of-art homogeneous representatives in a single use, can
be reused many times.
Several methods can be used to achieve recyclable catalytic systems, such
as the following ones: (i) utilization of heterogeneous solid catalytic mate-
rials; (ii) formation of dispersed nano-, sol-gel, and micellar systems;
(iii) phase division, where a homogeneous catalyst and substrate are usually
well soluble in one solvent, while the product is soluble in another solvent,
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 145

which in its turn is unmixable with the first one. Other methods that gen-
erally lead to such approaches have also been developed, e.g., immobiliza-
tion of intrinsically homogeneous catalysts on a solid support without
significant loss of their activity. In this case (supported catalysis), advantages
of homogeneous catalysts (e.g., enantioselectivity) can be combined with
easy recovery and reutilization. Another idea concerns the use of physical
forces, e.g., magnetic fields, to remove a catalyst from the reaction mixture.
Some advances in the development of these and related ideas on catalysts
recyclization and improvement of the overall activity of catalytic systems
for alcohol oxidation are discussed below.

8.1 Heterogeneous Solid Oxides, Alloys, and Related Materials


Many classical heterogeneous catalysts, e.g., oxides and related compounds
with incorporated ruthenium, gold, palladium, or platinum were found to
be effective for the aerobic oxidation of alcohols.2 Silver210 and cobalt211
were also included in this catalytic family and have raised expectations
regarding the availability of the catalysts.
Benzyl alcohol is a typical model substrate to test a chosen heterogeneous
catalyst and the reaction conditions. Apart form the formation of benzalde-
hyde, disproportionation, and dehydration can occur to give toluene or
dibenzyl ether (Scheme 52).212,213 Therefore, selective, active, and recycla-
ble heterogeneous catalysts are highly sought after. In this respect, AuPd
alloys supported on activated carbon, as well as the monometallic Pd and
Au were tested for alcohol oxidation in the presence of O2.212 The AuPd
alloy possesses a higher catalytic activity than the monometallic Pd (TOF
of 38 or 54 h1 for Pd or AuPd alloy, respectively) maintaining a high

Scheme 52 Reaction scheme for benzyl alcohol oxidation.212


ARTICLE IN PRESS

146 Maximilian N. Kopylovich et al.

selectivity (>94%) toward benzaldehyde. An unexpected result was that,


under the same conditions, Au was inactive. The better activity was
explained by the smaller interatomic distances in AuPd which leads to a bet-
ter contact between the reagents and the catalyst.
A direction in the development of heterogeneous catalysts for alcohol
oxidation concerns the synthesis of composite materials, such as Fe(III)
substituted Keggin-type clusters dispersed in amorphous silica matrix
(PWFe/SiO2).214 This composite was tested as a catalyst in the oxidation
of alcohols into aldehydes using H2O2 as a “green” oxidant. Under mild
reaction conditions, the catalyst showed a high activity and selectivity, with
good yields for all the tested substrates. The good catalyst activity was attrib-
uted to the large surface area owing to its micro and mesoporous structures
and strong acidity of the polyoxometalate component. The stability and
reusability of the catalyst was also quite good; moreover, the catalyst prep-
aration is simple and direct from cheap starting materials.
An example of a flow chemistry process is the Oppenauer oxidation of
secondary benzylic alcohols using partially hydrated zirconia and various
carbonyl compounds as oxidants (Scheme 53).215 The authors applied this
procedure to electron-rich and electron-deficient substrates, with improve-
ment in temperature (as low as 40 °C) and an easy reaction workup. The
reuse of the catalyst was performed several times, without loss in catalytic
efficiency.
Cerium(IV) oxide-based heterogeneous catalysts are of interest in
oxidation owing to the unique redox properties of cerium, i.e., if the crys-
tallite size of ceria decreases, an increase in the oxygen vacancy defect con-
centration occurs leading to attractive catalytic properties.216 On the other
hand, due to peculiar catalytic activity of gold, its combination with other
materials is also appealing, especially when the combination promotes the

Scheme 53 Example of the flow Oppenauer oxidation of secondary benzylic


alcohols.215
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 147

development of eco-friendly processes. Thus, the composite Au/CeO2


material was prepared by two distinct methods: homogeneous deposi-
tion–precipitation (ACH) and direct anionic exchange (ACD).216 The effi-
ciency of the two processes of catalyst fabrication was evaluated using the
aerobic oxidation of benzyl alcohol under solvent- and base-free reaction
conditions. The ACH catalyst exhibited a higher alcohol conversion
(64.5%) than the ACD catalyst, (53.8%), with comparable selectivities. This
difference, according to the authors, is due to presence of smaller sized gold
NPs and the higher number of oxygen vacancy sites on the ceria surface: the
ACH sample shows surface oxygen vacancies, whereas no oxygen vacancies
were found in the ACD sample. The conversion of benzyl alcohol increased
with reaction time, while the selectivity on benzaldehyde diminished with
an overproduction of benzyl benzoate. The recyclability of both catalysts
decreased after repeated use, which might be due to their structural changes.
The related Au/CeO2–Al2O3 catalyst was reported to be effective for the
heterogeneous oxidation of 1-tetradecanol, used as a model to test the
catalytic potential of the material toward other fatty alcohols.217 All the reac-
tions were performed at 80–120 °C with molecular oxygen at atmospheric
pressure and no added base. The highest conversion was 38%, while the
reaction selectivity was up to 70% for tetradecanoic acid and up to 80%
for tetradecanal.
Hydrotalcites have also attracted much attention as useful precursors for
the development of new environmentally friendly catalysts. Thus, Pt/Au
alloy NPs, supported over hydrotalcites and with soluble starch as a green
reducing and stabilizing agent, are catalysts for the selective aerobic oxida-
tion of glycerol and 1,2-propanediol.218 The reaction conditions were mild;
the oxidation was being performed in aqueous solution with no base added
and using molecular oxygen. The authors tested the individual metals as cat-
alysts and concluded that the bimetallic catalyst exhibits some synergetic
properties. The high activity and selectivity of these bimetallic catalysts sug-
gest that Pt atoms gain more electrons than Au atoms in PtxAuy–starch/
hydrotalcites as a result of the alterations of geometry and due to two types
of electron transfers: (i) from the starch ligand to both Au and Pt atoms and
(ii) from Au to Pt atoms.218
A series of cobalt-doped vanadium phosphorus oxide catalysts was pre-
pared using a classical organic method, followed by calcination, and tested
for the oxidation of benzyl alcohol with TBHP as an oxidant.219 The catalytic
activity increases with the temperature growth up to 68% at 90 °C, while
selectivity toward benzaldehyde varies in different solvents and reaches
ARTICLE IN PRESS

148 Maximilian N. Kopylovich et al.

Scheme 54 Oxidation of benzyl alcohol with cobalt-doped vanadium phosphorus


oxide catalysts.219

100% following the trend: acetonitrile > chloroform > toluene > dioxane.
The effect of the competitive adsorption between the solvents and benzyl
alcohol was discussed. The proposed mechanism (Scheme 54) involves
two active phases, where a suitable V5+/V4+ balance is required; the presence
of Co increases the average oxidation number of vanadium creating higher
amount of V5+ species, which are essential for the reversible V4+/V5+ redox
mechanism.
A porphyrin-containing cellulose derivative, namely hematin-appended
6-aminocellulose, performed well as a catalyst for the oxidation of guaiacol
and synapyl alcohol.220 The catalytic material is insoluble in most alcohols
and can be considered as a heterogeneous biomimetic catalyst. The high oxi-
dation activity and stability of the catalyst might be due to the cellulose back-
bone that inhibits the self-aggregation of the hematin moieties. To probe the
potential of the cellulose backbone as a chiral catalyst, oxidation of sinapyl
alcohol was performed, but the material showed no chiral behavior.

8.2 Supported Catalysts


Supported catalysts can be considered as heterogeneous ones which, how-
ever, exploit some features of homogeneous catalytic systems, e.g., selectiv-
ity and activity. Thus, vanadium-substituted phosphotungstic acid
immobilized on amine-functionalized MCM-41 exhibited high activity
and selectivity in the oxidation of aromatic alcohols to the corresponding
aldehydes with H2O2, even after five cycles.221 It should be mentioned that
not only catalytic systems as a whole, but also their components, in particular
the most expensive and unstable ones, can be immobilized and reused. For
instance, fullerene has been employed as a molecular support for TEMPO
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 149

and this combination was further applied for the oxidation of primary and
secondary alcohols using the Anelli protocol.76
In another study,222 MCM-41 was used as a support, to anchor
undecatungstophosphate. The thus prepared bifunctional catalyst was tested
for oxidation, as well as esterification, of benzyl alcohols. Kinetic studies rev-
ealed that the reactions follow first order kinetic patterns, and the low values
of activation energy for esterification and oxidation are indicative that the
reaction rate is governed by a chemical step. A curious feature of the
supported tungstophosphate is the drastic change in selectivity of the reac-
tion with time and increasing temperature. Thus, 100% selectivity toward
benzaldehyde was achieved for 2 h but, after 24 h, the selectivity shifted
toward benzoic acid. When the reaction temperature increases, the conver-
sion of the alcohol also grows, as well as the selectivity toward benzoic acid.
This overoxidation of benzaldehyde to benzoic acid might be due to the
high acidity of the catalyst or an effect of the support.
The effect of a support can provide a key to achieve a high catalytic effi-
ciency: for instance, TONs up to 63,000, with selectivities of 99% and con-
versions between 71% and 99% were reached when the oxidation of
activated, nonactivated, and heterocyclic alcohols were studied with
1 atm of molecular oxygen and a zirconia supported ruthenium catalyst.223
Such high TONs were attributed to the properties of ZrO2 surface, in par-
ticular hydroxyl groups and coordinatively unsaturated Lewis acidic-basic
Zr4+ and O2 pairs. Developments in support materials, which can facilitate
electron transfer, increase long-term stability of the catalysts, etc., are highly
desirable.
With Pd dispersed over mesoporous SiO2, the oxidation of crotyl and
cinnamyl alcohols (Scheme 55) can be achieved with high efficiency.224,225

Scheme 55 Cinnamyl alcohol oxidation.224


ARTICLE IN PRESS

150 Maximilian N. Kopylovich et al.

The generation of atomically dispersed Pd2+ surface species at low palladium


loadings promotes the high activity in the oxidation of allylic alcohols. The
hierarchically ordered nanoporous Pd/SBA-15 was shown to be a key factor
to obtain aldehydes upon oxidation of sterically hindered allylic alcohols,
such as phytol and farnesol. The results show how important is the capability
of support materials to stabilize the metal oxide and to provide a specific pore
size thus promoting the mass-transfer.
A robust matrix material can also be employed as a support to accommo-
date active metals, metal oxides, polymers, etc. Thus, three-dimensional
graphene-based frameworks were used to support copper phthalocyanines
and show improved thermal and chemical stability combined with a com-
petitive overall cost and availability.226 This catalyst was successfully tested
for the selective aerobic oxidation of alcohols to the corresponding carbonyl
compounds. The high catalytic activity of the material was related to π–π
interactions between benzene moieties of reactants and graphene that favor
the interaction of the reagents and catalytic sites. On the other hand, the
presence of basic sites on the support also helps to improve the selectivity.
The use of more active and stable reducible oxides, like TiO2, has some
advantages over the nonreducible ones, such as SiO2 and Al2O3. Thus,
mesoporous titanium dioxide (anatase) was applied as a support for the depo-
sition of gold NPs and the thus prepared material was used for the vapor
phase oxidation of benzyl alcohol.227 The activity of the catalyst decreased
in terms of TOFs against metal loading, what is probably due to agglomer-
ation of gold NPs and hence lower number of available active metal sites.

8.3 Nano, Dispersed and Micellar Catalysts


One of the ways to enhance the activity of heterogeneous catalytic systems
consists in providing a high surface/volume ratio, i.e., its dispersion down to
nano-scale. Thus, nanoporous stainless steel (NPSS) electrode materials with
copper and a film of palladium were fabricated and it was found that their
porosity and the presence of Cu improve the long-term stability of the
Pd film on the surface.228 The presence of Cu has a significant effect on
the catalytic activity, the reaction kinetics, and poisoning tolerance of the
NPSS/Cu/Pd electrode. The electrode was tested for the electrooxidation
of glycerol, with comparable results of those of palladium-carbon catalysts,
possibly due to the large electrochemically active surface area.
Palladium clusters can be encapsulated in a microporous silica shell; the
thus prepared heterogeneous catalyst was tested for the solvent-free aerobic
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 151

oxidation of various hydrocarbons and alcohols and showed a high activity,


with TOF values of up to 54,740 h1.229 The activity and selectivity depend
on the sizes of pores and substrates. The recyclability was also tested and
reached 20 without any loss of activity for benzyl alcohol oxidation;
the corresponding overall TONs were estimated to be ca. 280,000.
Palladium NPs supported on carbon NTs functionalized with various
organosilane modifiers have been tested for the selective aerobic oxidation
of benzyl alcohol and quasi-TOFs based on the active surface area as high as
288,755 h1 were obtained.230 The high selectivity toward benzaldehyde
was ascribed to the low surface hydrogen concentration leading to dimin-
ished formation of toluene, and to the low surface basicity that hampers
the disproportionation of benzaldehyde to form benzoic acid. In addition,
the basic catalyst support facilitates small and highly dispersed Pd NPs with
narrow size distribution, what favors the high activity of this catalyst even
after five consecutive runs. Another related catalytic system, composed of
conjugated microporous polymers, with encapsulated palladium NPs with
1.6–3.5 nm in size, showed a high catalytic activity for the benzyl alcohol
oxidation to benzaldehyde with conversions and yields up to 74%.231
Gold-containing poly(urea-formaldehyde) microparticles were prepared
by in situ polymerization using a series of stabilization agents and tested for
the selective oxidation of glycerol.232 The glycerol conversion and the
glyceric acid selectivity have opposite behaviors when the size of gold par-
ticle decreases. If the surface of stabilizer is hydrophobic, then the selectivity
to C3 products in the resulting catalysts is enhanced. If stabilizers with
hydrophilic surfaces are applied, the formation of C–C bond cleavage prod-
ucts is preferable.
Besides oxidants, solvent, and other components of the reaction mixture,
the environmental compatibility of the catalyst support also deserves to be
addressed. Novel biocompatible thiol-functionalized fructose-derived
nanoporous carbon support produced by hydrothermal carbonization can
significantly diminish the environmental impact.233 This porous carbon
material with supported gold NPs was highly active in selective aerobic oxi-
dation of several alcohols to the corresponding aldehydes and ketones. The
catalyst was easily recovered and reused 6  without leaching of metals or
loss of activity. The proposed mechanism for the catalytic oxidation of alco-
hols involves the base promoted deprotonation of alcohol to form alkoxide
on the Au surfaces. Then gold catalyzes the β-hydrogen elimination to
produce the corresponding aldehyde, along with the formation of O2
and H2O.233
ARTICLE IN PRESS

152 Maximilian N. Kopylovich et al.

Another remarkable direction concerning the promotion of the catalyst


active area and its recyclability consists in the use of micelle catalytic systems.
They can combine the advantages of homogeneous and heterogeneous cat-
alyzes, because, on one hand, the catalyst and substrate are soluble in one
solvent, while the product possesses a high solubility in another one. The
transfer of the product to another phase shifts the equilibrium, while the cat-
alyst is responsible for the kinetics. Usually micellar structures are used,
where catalysts are confined within the small droplets of one solvent, sepa-
rated from another one and commonly stabilized by surfactants. Thus,
enzyme-inspired star block-copolymers with limited branching were tested
in catalytic systems for the oxidation of alcohols in water, in particular for a
Cu/TEMPO-catalyzed alcohol oxidation reaction34 90% conversion of
benzyl alcohol to benzaldehyde was obtained after 44 h of reaction. The fact
that the polymers are able to preconcentrate molecular oxygen is of partic-
ular significance for further developments.

8.4 ILs and Related Systems with Phase Division


The use of ILs234 and supercritical fluids235 in catalytic oxidation has been
regarded as a new possibility for catalyst recycling and enhancing the product
yield and selectivity. For instance, recycling of expensive TEMPO is not a
trivial task due to the homogeneous character of most of the TEMPO cat-
alytic systems. Moreover, in some cases there are also drawbacks due to
overoxidation that lower conversion and selectivity. The replacement of
organic solvents by ILs can provide an effective strategy to avoid such prob-
lems.236 Thus, a vanadium-based catalyst, TEMPO and sulfonic acid cocat-
alysts were grafted on the [C4py][BF4] IL and used for oxidation of alcohols
with H2O2 as oxidant, exhibiting good activity and recyclability.237
In another example, TEMPO was incorporated into supported [C6mim]
[BF4] IL (Fig. 14A) and exhibited high activity for alcohol oxidation using
bis(acetoxy)iodobenzene (BAIB) as the terminal oxidant.236 The catalyst can

Figure 14 Strategies for immobilizing TEMPO on ILs. [C4mim][BF4] supported TEMPO


(A) and IL@SBA-15-TEMPO (B).79
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 153

be recycled together with the IL without loss of the efficiency for several
cycles. TEMPO can be also grafted on SBA-15 solid support be combined
with 1-methyl-3-butylimidazolium ([C4mim][Br]) IL thus forming the
IL@SBA-15-TEMPO catalytic system (Fig. 14B).79 This system possesses
improved selectivity and good recyclability for the oxidation of alcohols
to aldehydes and ketones with TBN as an oxidant in AcOH.
Deep eutective solvents (DESs) are a novel class of ILs that are generally
obtained by the interfusion of quaternary ammonium salts and hydrogen
bond donors (e.g., amides, amines, alcohols, and carboxylic acids). Their
ionic nature and relatively high polarity provide good solubility for many
ionic species, such as metal salts. They also have other advantages over com-
mon ILs, such as the simple and easy preparation as pure phases from cheap
and easily available components or high chemical stability toward atmo-
spheric moisture and temperature. These novel DES were used to incorpo-
rate Fe(NO3)3 9H2O in TEMPO; the DES–TEMPO/Fe(NO3)3 system
showed good performances in the selective oxidation of various alcohols
to the corresponding aldehydes and ketones, using molecular oxygen as
an oxidant and under mild and solvent-free conditions.238 As expected,
the DES was easily recovered and recycled up to 5  without significant loss
of catalytic activity.
Glycols constitute other media which have been widely used in organic
transformations as environmentally benign solvents and soluble supports for
liquid phase synthesis. ILs and polyethylene glycols can also be combined in
the solvent–free aerobic oxidation of alcohols to give an excellent catalytic
effect and easy catalyst recovery.239 The bifunctionalized combined
IL-glycol PEG1000 catalytic system ([Imim-PEG1000-TEMPO][CuCl2])
shows catalytic properties similar to those of its nonsupported counterpart
in terms of yields as high as 95% with 100% conversion and selectivity
toward ketone. Moreover, ([Imim-PEG1000-TEMPO][CuCl2]) could be
recycled and reused without significant loss of catalytic activity after
five runs.
The combination of several of the above described approaches can be also
used. Thus, precious metal catalysts can be supported on nanomaterials and
combined with ILs.240 The supported gold NPs on graphene oxide (GO)
with an ionic liquid framework (Au@GO-IL) has been shown to be a highly
active, and leaching tests, such as hot filtration test and AAS analysis, indicate
that the catalytic reaction is mainly heterogeneous in nature. The reusability
of this catalyst was tested for 5  without a significant decrease in its catalytic
activity.240 Also using gold NPs, but performing oxidation under MW
ARTICLE IN PRESS

154 Maximilian N. Kopylovich et al.

irradiation, 2-hydroxybenzyl alcohol was converted successfully to


2-hydroxybenzaldehyde in the presence of 3-chloroperoxybenzoic acid
and hydrogen peroxide in methanol.241
Pd NPs are commonly used in oxidations of alcohols and can be incor-
porated into an IL.242 The morphology of the particles can be suited to dif-
ferent applications, e.g., flower-like particles, due to their concave
tetrahedral subunits, exhibited a high electrocatalytic activity toward ethanol
and methanol oxidation compared with that of the commercial Pd black
catalyst.242 A Pd complex containing triphenylphosphine and a Schiff base
catalyst was used243 for the study of the solvent effect in carbonylation of
primary and secondary alcohols to aldehydes and ketones, in the presence
of NaOCl as an oxidant. By kinetic study of different proportions between
the imidazolium-based IL ([C2mim][PF6]), it was shown that the accelera-
tion of the reaction depends on the mixing proportion and that the best ratio
was 1:1.
Another interesting aspect of using ILs instead of molecular solvents is
the possibility of bypassing steps. Thus, in a propane oxidation using rho-
dium (palladium)–copper–chloride catalytic systems immobilized in ILs,
propane is oxidized to acetone, bypassing the isopropanol formation step.244
Methane was also studied and is oxidized under more severe conditions than
propane, giving methyl trifluoroacetate as the main product.
Several hydrophobic methylimidazolium-based ILs were studied in
the oxidation of cyclohexanol to cyclohexanone with H2O2 and WO3 as
a catalyst.245 In the biphasic cyclohexanol-ILs system, 1-octyl-3-
methylimidazolium chloride ([C8mim]Cl) IL was found to effectively pro-
mote cyclohexanol oxidation to 100% conversion of cyclohexanol and
100% selectivity to cyclohexanone, what is accounted for by the biphasic
character of the system. The oxidation of cyclohexanol occurs in aqueous
phase containing H2O2 and the catalyst, while the produced cyclohexanone
is transferred to the organic phase, minimizing its further oxidation. Higher
concentrations of [C8mim]Cl favor the oxidation possibly by stabilizing
reaction intermediates in the catalytic process.245
A combination of a tungsten species (in particular, tungstic acid H2WO4)
and an IL allowed to obtain very good results in the oxidation of five- to
eight-membered cyclanols, using aqueous H2O2 MW or ultrasound activa-
tion with the ammonium-based IL Aliquat 336.208 The oxidation reaction
was studied in several ILs and no effect of the aromaticity of the IL cation was
observed. However, the anion of the IL seems to be important and the yields
increased with its size. This can be a key factor for the choice of ILs.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 155

The use of lanthanides with ILs for oxidation reactions is still scarce, but a
hydrogen peroxide–urea adduct and catalytic (CF3SO3)3La in an IL was
recently used for the oxidation of a secondary alcohol to ketone.246
A number of 1,2-diols, α-hydroxyketones and other aromatic, and aliphatic
secondary alcohols have been successfully oxidized to the corresponding
ketones with yields from 74% up to 92% and reaction times between 0.5
and 3 h.
Imidazolium-based ILs were used as nonconventional media in alcohol
dehydrogenase (ADH)-catalyzed reactions in enzymatic catalysis.247 When
containing up to 50% of the IL, the overall conversion could be improved in
some cases, while the stereoselectivity of the enzyme remained unaltered.247
Besides enzymatic catalysis, the development and use of green and
efficient methods to transform lignocellulosic biomass (along with cellulose
and hemicellulose) into fuels and high value-added chemicals is another
appealing area. Thus, one-pot sequential oxidation and aldol conden-
sation reactions of veratryl alcohol in the basic ionic liquid (BIL)
1-butyl-3-methylimidazolium 5-nitrobenzimidazolide, which acted as the
solvent and basic additive, was studied.248 The effects of different factors, such
as the type of catalyst, reaction time, reaction temperature, and the amount of
BIL, on the oxidation reaction were investigated. It was shown that the cat-
alytic performance of individual Ru@ZIF-8 (zeolitic imidazolate framework-
8) or CuO was very poor for the oxidation of veratryl alcohol to veratryl alde-
hyde. Interestingly, Ru@ZIF-8 + CuO was very efficient for the oxidation
reaction and a high yield of veratryl aldehyde could be obtained, indicating
the synergistic effect of the two catalysts in the BIL. The veratryl aldehyde
generated by the oxidation of veratryl alcohol could react directly with ace-
tone to form, in high yield, 3,4-dimethoxybenzylideneacetone by aldol con-
densation reaction catalyzed by the BIL.

8.5 Other Directions


Supercritical CO2 is the most used supercritical fluid due to its favorable
characteristics. Its low toxicity, relatively low critical temperature and stabil-
ity allow most compounds to be extracted with little damage. In addition,
the solubility of many extracted compounds in CO2 varies with pressure,
allowing selective extractions. These possibilities are particularly useful in
the reactions involving gaseous reagents such as oxidation with O2. Con-
cerning the selective aerobic oxidation of alcohols, CO2 can dry wet mate-
rial thus allowing to achieve a high selectivity to aldehyde by suppressing the
ARTICLE IN PRESS

156 Maximilian N. Kopylovich et al.

formation of carboxylic acid via the favored hydration of aldehyde.249 In


another work250 it was reported that the conversion of benzyl alcohol to
benzaldehyde in CO2 increased 50% for a pressure increase of 7% because
the substrate and products are distributed differently in the organic and
supercritical phases, thus affecting both the reaction rate and selectivity.
A promising perspective for catalyst recyclization concerns the applica-
tion of magnetically recoverable catalysts that can be readily collected by
magnetic attraction without the use of traditional isolation methods, such
as filtration, extraction or centrifugation. For instance, magnetic CoFe2O4
NPs can efficiently catalyze the oxidation of alcohols to the corresponding
carbonyl products and then be easily recovered with assistance of a magnetic
field and reused several times.251 If the catalytically active particles are non-
magnetic, they can be immobilized on the surface of a magnetic carrier, e.g.,
Fe2O3 or Fe3O4, be applied for the oxidation of various alcohols, and then
be recovered by application of a magnetic field. Thus, superparamagnetic
Fe3O4@EDTA–Cu(II) NPs were readily prepared and identified as an
effective catalyst for the tandem transformation of benzyl alcohols and amine
hydrochloride salts into the corresponding amides with TBHP as an
oxidant.154 After completion of the reaction, the catalyst can be removed
from the reaction vessel by assistance of an external magnet and reused at
least 6  without significant loss of its activity.
Bifunctional bimetallic alloys in which both catalytic and magnetic func-
tions are simultaneously provided can be applied. For example, Co-Pd bime-
tallic alloy NP catalysts were prepared and employed for the aerobic oxidation
of a variety of alcohols in water.252 The catalysts then were magnetically
recovered and reused for further oxidation. Leaching of Co and Pd was in
orders of only 103 mol% and 106 mol%, respectively. The proposed expla-
nation for the low leaching involves an electrolytic “protection” of the more
expensive Pd component: Eox of Pd(0)/Pd(II) (0.95 V) is higher than Eox of
Co(0)/Co(II) (0.28 V). Hence, Co(0) is oxidized by Pd(II) to give Co(II),
while Pd(0) aggregates with initial NPs and remains within the catalyst.
To conclude, heterogeneous catalysis with a wide range of possible
supports is rather appealing for industrial applications. The easy recovery
and recycling ability are two of the most desirable features. Although
some concerns still remain regarding their limited catalytic activity and
deactivation,253 many of the recently introduced protocols for the alcohol
oxidation involve supported transitional metal catalysts.174 The pursuit of
oxidation systems without such limitations and which are readily accessible,
stable, inexpensive, environmentally acceptable, and can promote selective
oxidation under mild reaction conditions is still under way.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 157

ACKNOWLEDGMENTS
This work has been partially supported by the Foundation for Science and
Technology (FCT), Portugal (FCT Doctoral Program CATSUS, UID/QUI/00100/
2013 and “Investigador 2013” [IF/01270/2013/CP1163/CT0007] programs). N. M. R.
M., A. P. C. R., and M. N. K. express gratitude to FCT for doctorate, post-doc
fellowships and working contract, respectively.

REFERENCES
1. Tojo G, Fernandez M. Oxidation of Alcohols to Aldehydes and Ketones. New York, USA:
Springer; 2006.
2. Mallat T, Baiker A. Oxidation of alcohols with molecular oxygen on solid catalysts.
Chem Rev. 2004;104:3037–3058.
3. Sarathy SM, Oßwald P, Hansen N, Kohse-H€ oinghaus K. Alcohol combustion chem-
istry, progress in energy and combustion. Science. 2014;44:40–102.
4. Bianchini C, Shen PK. Palladium-based electrocatalysts for alcohol oxidation in half
cells and in direct alcohol fuel cells. Chem Rev. 2009;109:4183–4206.
5. Sheldon RA, Arends IWCE, Dijksman A. New developments in catalytic alcohol
oxidations for fine chemicals synthesis. Catal Today. 2000;57:157–166.
6. Allen SE, Walvoord RR, Padilla-Salinas R, Kozlowski MC. Aerobic copper-catalyzed
organic reactions. Chem Rev. 2013;113:6234–6645.
7. Sheldon RA, Arends IWCE, Ten Brink G-J, Dijksman A. Green catalytic oxidations of
alcohols. Acc Chem Res. 2002;35:774–781.
8. Dobereiner GE, Crabtree RH. Dehydrogenation as a substrate-activating strategy in
homogeneous transition-metal catalysis. Chem Rev. 2010;110:681–703.
9. Sigman MS, Jensen DR. Ligand-modulated palladium-catalyzed aerobic alcohol
oxidations. Acc Chem Res. 2006;39:221–229.
10. Ryland BL, Stahl SS. Practical aerobic oxidations of alcohols and amines with homo-
geneous copper/TEMPO and related catalyst systems. Angew Chem Int Ed Engl.
2014;53:8824–8838.
11. Liang Z-X, Zhao TS. Catalysts for Alcohol-Fuelled Direct Oxidation Fuel Cells.
Cambridge: RSC; 2012.
12. Trincado M, Banerjee D, Grutzmacher H. Molecular catalysts for hydrogen produc-
tion from alcohols. Energy Environ Sci. 2014;7:2464–2503.
13. (a) Backvall J-E, ed. Modern Oxidation Methods. Weinheim: Wiley; 2010.
(b) Arends IWCE, Sheldon RA. Modern oxidation of alcohols using environmentally
benign oxidants. In: Backvall J-E, ed. Modern Oxidation Methods. Weinheim: Wiley;
2010. (c) Ishii Y, Sakaguchi S, Obora Y. Aerobic oxidations and related reactions cat-
alyzed by n-hydroxyphthalimide. In: Backvall J-E, ed. Modern Oxidation Methods.
Weinheim: Wiley; 2010. (d) Browne WR, de Boer JW, Pijper D, Brinksma J,
Hage R, Feringa BL. Manganese-catalyzed oxidation with hydrogen peroxide.
In: Backvall J-E, ed. Modern Oxidation Methods. Weinheim: Wiley; 2010.
(e) Murahashi S-I, Komiya N. Ruthenium-catalyzed oxidation for organic synthesis.
In: Backvall J-E, ed. Modern Oxidation Methods. Weinheim: Wiley; 2010.
14. Mizuno N. Modern Heterogeneous Oxidation Catalysis. Weinheim: Wiley-VCH Verlag
GmbH & Co. KGaA; 2009.
15. Sheldon RA, Arends IWCE. Organocatalytic oxidations mediated by nitroxyl radicals.
Adv Synth Catal. 2004;346:1051–1071.
16. Silva JAL, Silva JJRF, Pombeiro AJL. Oxovanadium complexes in catalytic oxidations.
Coord Chem Rev. 2011;255:2232–2248.
17. Kumpulainen ETT, Koskinen AMP. Catalytic activity dependency on catalyst compo-
nents in aerobic copper–TEMPO oxidation. Chem Eur J. 2009;15:10901–10911.
ARTICLE IN PRESS

158 Maximilian N. Kopylovich et al.

18. Hoover JM, Ryland BL, Stahl SS. Copper/TEMPO-catalyzed aerobic alcohol oxidation:
mechanistic assessment of different catalyst systems. ACS Catal. 2013;3:2599–2605.
19. Parmeggiani C, Cardona F. Transition metal based catalysts in the aerobic oxidation of
alcohols. Green Chem. 2012;14:547–564.
20. Wigington BN, Drummond ML, Cundari TR, Thorn DL, Hanson SK, Scott SL.
A biomimetic pathway for vanadium-catalyzed aerobic oxidation of alcohols:
evidence for a base-assisted dehydrogenation mechanism. Chem Eur J. 2012;18:
14981–14988.
21. Zakzeski J, Bruijnincx PCA, Jongerius AL, Weckhuysen BM. The catalytic valoriza-
tion of lignin for the production of renewable chemicals. Chem Rev.
2010;10:3552–3599.
22. Roberts SM, Whittall J. Regio- and Stereo-Controlled Oxidations and Reductions.
Weinheim: Wiley-VCH; 2007.
23. Griffith WP. Ruthenium Oxidation Complexes. Dordrecht: Springer; 2011.
24. (a) Schultz MJ, Sigman MS. Recent advances in homogeneous transition metal-
catalyzed aerobic alcohol oxidations. Tetrahedron. 2006;62:8227–8241. (b) Zhan B-Z,
Thompson A. Recent developments in the aerobic oxidation of alcohols. Tetrahedron.
2004;60:2917–2935.
25. Besson M, Gallezot P. Selective oxidation of alcohols and aldehydes on metal catalysts.
Catal Today. 2000;57:127–141.
26. Figiel PJ, Kirillov AM, Guedes da Silva MFC, Lasri J, Pombeiro AJL. Self-assembled
dicopper(II) diethanolaminate cores for mild aerobic and peroxidative oxidation of
alcohols. Dalton Trans. 2010;39:9879–9888.
27. Figiel PJ, Kirillov AM, Karabach YY, Kopylovich MN, Pombeiro AJL. Mild aerobic
oxidation of benzyl alcohols to benzaldehydes in water catalyzed by aqua-soluble
multicopper(II) triethanolaminate compounds. J Mol Catal A Chem.
2009;305:178–182.
28. Zhang G, Proni G, Zhao S, et al. Chiral tetranuclear and dinuclear copper(II) com-
plexes for TEMPO-mediated aerobic oxidation of alcohols: are four metal centres bet-
ter than two? Dalton Trans. 2014;43:12313–12320.
29. Raisanen MT, Al-Hunaiti A, Atosuo E, Kemell M, Leskela M, Repo T. Mn(II) acetate:
an efficient and versatile oxidation catalyst for alcohols. Catal Sci Technol.
2014;4:2564–2573.
30. (a) John LC, Gunay A, Wood AJ, Emmert MH. Catalysts for convenient aerobic alco-
hol oxidations in air: systematic ligand studies in Pd/pyridine systems. Tetrahedron.
2013;69:5758–5764. (b) Bailie DS, Clendenning GMA, McNamee L, Muldoon MJ.
Anionic N,O-ligated Pd(II) complexes: highly active catalysts for alcohol oxidation.
Chem Commun. 2010;46:7238–7240.
31. Gowrisankar S, Neumann H, G€ ordes D, Thurow K, Jiao H, Beller M. A convenient
and selective palladium-catalyzed aerobic oxidation of alcohols. Chem Eur J.
2013;19:15979–15984.
32. Ingram AJ, Solis-Ibarra D, Zare RN, Waymouth RM. Trinuclear Pd3O2 intermediate
in aerobic oxidation catalysis. Angew Chem Int Ed Engl. 2014;53:5648–5652.
33. Giachi G, Frediani M, Oberhauser W, Passaglia E. Aerobic alcohol oxidation catalyzed
by polyester-based Pd(II) macrocomplexes. J Polym Sci A Polym Chem.
2012;50:2725–2731.
34. Mugemana C, Chen B-T, Bukhryakov KV, Rodionov V. Star block-copolymers:
enzyme-inspired catalysts for oxidation of alcohols in water. Chem Commun.
2014;50:7862–7865.
35. Hu Z, Kerton FM. Room temperature aerobic oxidation of alcohols using CuBr2 with
TEMPO and a tetradentate polymer based pyridyl-imine ligand. Appl Catal A Gen.
2012;413–414:332–339.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 159

36. Zhang G, Han X, Luan Y, Wang Y, Wen X, Ding C. l-Proline: an efficient N,


O-bidentate ligand for copper-catalyzed aerobic oxidation of primary and secondary
benzylic alcohols at room temperature. Chem Commun. 2013;49:7908–7910.
37. Kirillov AM, Kirillova MV, Shul’pina LS, et al. Mild oxidative functionalization of
alkanes and alcohols catalyzed by new mono- and dicopper(II) aminopolyalcoholates.
J Mol Catal A Chem. 2011;350:26–34.
38. (a) Mahmudov KT, Kopylovich MN, Silva MFCG, Figiel PJ, Karabach YY,
Pombeiro AJL. New copper(II) dimer with 3-(2-hydroxy-4-nitrophenylhydrazo)-
pentane-2,4-dione and its catalytic activity in cyclohexane and benzyl alcohol oxida-
tions. J Mol Catal A Chem. 2010;318:44–50. (b) Kopylovich MN, Mahmudov KT,
Silva MFCG, et al. Ortho-hydroxyphenylhydrazo-β-diketones: tautomery, coordina-
tion ability, and catalytic activity of their copper(II) complexes toward oxidation
of cyclohexane and benzylic alcohols. Inorg Chem. 2011;50:918–931.
(c) Kopylovich MN, Karabach YY, Mahmudov KT, et al. Heterometallic
copper(II)–potassium 3D coordination polymers driven by multifunctionalized azo
derivatives of β-diketones. Cryst Growth Des. 2011;11:4247–4252.
(d) Kopylovich MN, Gajewska MJ, Mahmudov KT, et al. Copper(II) complexes with
a new carboxylic-functionalized arylhydrazone of β-diketone as effective catalysts for
acid-free oxidations. New J Chem. 2012;36:1646–1654. (e) Kopylovich MN,
Mizar A, Silva MFCG, MacLeod TCO, Mahmudov KT, Pombeiro AJL. Template
syntheses of copper(II) complexes from arylhydrazones of malononitrile and their cat-
alytic activity towards alcohol oxidations and the nitroaldol reaction: hydrogen bond-
assisted ligand liberation and E/Z isomerisation. Chem Eur J. 2013;19:588–600.
(f ) Mahmudov KT, Kopylovich MN, Sabbatini A, et al. Cooperative metal-ligand
assisted E/Z isomerization and cyano activation at Cu(II) and Co(II) complexes of
arylhydrazones of active methylene nitriles. Inorg Chem. 2014;53:9946–9958.
39. Borthakur R, Asthana M, Saha M, Kumar A, Pal AK. An efficient oxidation of alcohols
using a new trinuclear copper complex as a reusable catalyst under solvent free condi-
tions. RSC Adv. 2014;4:21638–21643.
40. Nasani R, Saha M, Mobin SM, et al. Copper-organic frameworks assembled from
in situ generated 5-(4-pyridyl)tetrazole building blocks: synthesis, structural features,
topological analysis and catalytic oxidation of alcohols. Dalton Trans. 2014;43:
9944–9954.
41. Fei H, Shin J, Meng YS, et al. Reusable oxidation catalysis using metal-
monocatecholato species in a robust metal–organic framework. J Am Chem Soc.
2014;136:4965–4973.
42. Fernandes RR, Lasri J, da Silva MFCG, da Silva JAL, da Silva JJR F, Pombeiro AJL.
Bis- and tris-pyridyl amino and imino thioether Cu and Fe complexes. Thermal
and microwave-assisted peroxidative oxidations of 1-phenylethanol and cyclohexane
in the presence of various N-based additives. J Mol Catal A Chem. 2011;351:
100–111.
43. Fernandes RR, Lasri J, da Silva MFCG, da Silva JAL, Fraústo da Silva JJR,
Pombeiro AJL. Mild alkane C–H and O–H oxidations catalysed by mixed-N,
S copper, iron and vanadium systems. Appl Catal A Gen. 2011;402:110–120.
44. Lenze M, Bauer EB. Chemoselective, iron(ii)-catalyzed oxidation of a variety of sec-
ondary alcohols over primary alcohols utilizing H2O2 as the oxidant. Chem Commun.
2013;49:5889–5891.
45. (a) Al-Hunaiti A, Niemi T, Sibaouih A, Pihko P, Leskela M, Repo T. Solvent
free oxidation of primary alcohols and diols using thymine iron(III) catalyst. Chem
Commun. 2010;46:9250–9252. (b) Join B, M€ oller K, Ziebart C, et al. Selective iron-
catalyzed oxidation of benzylic and allylic alcohols. Adv Synth Catal. 2011;353:
3023–3030.
ARTICLE IN PRESS

160 Maximilian N. Kopylovich et al.

46. (a) Hanson SK, Wu R, Silks LAP. Mild and selective vanadium-catalyzed oxidation of
benzylic, allylic, and propargylic alcohols using air. Org Lett. 2011;13:1908–1911.
(b) Ohde C, Limberg C. From surface-inspired oxovanadium silsesquioxane models
to active catalysts for the oxidation of alcohols with O2-the cinnamic acid/
metavanadate system. Chem Eur J. 2010;16:6892–6899. (c) Sedai B, Dı́az-Urrutia C,
Baker RT, Wu R, Silks LAP, Hanson SK. Comparison of copper and vanadium homo-
geneous catalysts for aerobic oxidation of lignin models. ACS Catal. 2011;1:794–804.
(d) Son S, Toste FD. Non-oxidative vanadium-catalyzed CO bond cleavage: appli-
cation to degradation of lignin model compounds. Angew Chem Int Ed Engl.
2010;122:3879–3882.
47. Zhang G, Scott BL, Wu R, Silks LAP, Hanson SK. Aerobic oxidation reactions catalyzed
by vanadium complexes of bis(phenolate) ligands. Inorg Chem. 2012;51:7354–7361.
48. D€onges M, Amberg M, Stapf G, Kelm H, Bergsträßer U, Hartung J. cis-2,6-
Bis-(methanolate)-piperidine oxovanadium(V) complexes as catalysts for oxidative
alkenol cyclization by tert-butyl hydroperoxide. Inorg Chim Acta. 2014;420:120–134.
49. Han L, Xing P, Jiang B. Selective aerobic oxidation of alcohols to aldehydes, carboxylic
acids, and imines catalyzed by a Ag-NHC complex. Org Lett. 2014;16:3428–3431.
50. Liu L, Yu M, Wayland BB, Fu X. Aerobic oxidation of alcohols catalyzed by
rhodium(III) porphyrin complexes in water: reactivity and mechanistic studies. Chem
Commun. 2010;46:6353–6355.
51. Taketoshi A, Beh XN, Kuwabara J, Koizumi T, Kanbara T. Aerobic oxidative dehy-
drogenation of benzyl alcohols to benzaldehydes by using a cyclometalated ruthenium
catalyst. Tetrahedron Lett. 2012;53:3573–3576.
52. Wang S-S, Zhang J, Zhou C-L, Vo-Thanh G, Liu Y. An ionic compound containing
Ru(III)-complex cation and phosphotungstate anion as the efficient and recyclable cat-
alyst for clean aerobic oxidation of alcohols. Catal Commun. 2012;28:152–154.
53. Borthakur R, Asthana M, Kumar A, Koch A, Lal RA. Solvent free selective oxidation
of alcohols catalyzed by a trinuclear complex with a dicopper(II)-monozinc(II) centre
using hydrogen peroxide as an oxidant. RSC Adv. 2013;3:22957–22962.
54. Sanmartin-Matalobos J, Garcia-Deibe AM, Briones-Miguens L, Gonzalez-Bello C,
Portela-Garcia C, Fondo M. Dual role of 2-tosylaminomethylaniline as a ligand and
a nucleophile in the copper-mediated oxidation of methanol. Dalton Trans.
2014;43:722–728.
55. Cao Q, Dornan LM, Rogan L, Hughes NL, Muldoon MJ. Aerobic oxidation catalysis
with stable radicals. Chem Commun. 2014;50:4524–4543.
56. Tong X, Sun Y, Yan Y, Luo X, Liu J, Wu Z. Transition metal-free catalytic oxidation
of aromatic alcohols with molecular oxygen in the presence of a catalytic amount of
N-bromosuccinimide. J Mol Catal A Chem. 2014;391:1–6.
57. Prebil R, Stavber G, Stavber S. Aerobic oxidation of alcohols by using a completely
metal-free catalytic system. Eur J Org Chem. 2014;395–402.
58. Ma X, Li Z, Liu F, Cao S, Rao H. Tetra-n-butylammonium bromide: a simple but
efficient organocatalyst for alcohol oxidation under mild conditions. Adv Synth Catal.
2014;356:1741–1746.
59. Guo S, Yu J-T, Dai Q, Yang H, Cheng J. The Bu4NI-catalyzed alfa-acyloxylation of
ketones with benzylic alcohols. Chem Commun. 2014;50:6240–6242.
60. Al-Hunaiti A, Räisänen M, Pihko P, Leskelä M, Repo T. Organocatalytic oxidation of
secondary alcohols using 1,2-di(1-naphthyl)-1,2-ethanediamine (NEDA). Eur J Org
Chem. 2014;6141–6144.
61. Rong Z-Q, Pan H-J, Yan H-L, Zhao Y. Enantioselective oxidation of 1,2-diols with
quinine-derived urea organocatalyst. Org Lett. 2013;16:208–211.
62. Ciriminna R, Pagliaro M. Industrial oxidations with organocatalyst TEMPO and its
derivatives. Org Process Res Dev. 2009;14:245–251.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 161

63. Anelli PL, Biffi C, Montanari F, Quici S. Fast and selective oxidation of primary alco-
hols to aldehydes or to carboxylic acids and of secondary alcohols to ketones mediated
by oxoammonium salts under two-phase conditions. J Org Chem. 1987;52:2559–2562.
64. (a) Sabbatini A, Martins LMDRS, Mahmudov KT. Microwave-assisted and solvent-
free peroxidative oxidation of 1-phenylethanol to acetophenone with a CuII–TEMPO
catalytic system. Catal Commun. 2014;48:69–72. (b) Sutradhar M, Martins LMDRS,
Guedes da Silva MFC, Alegria ECBA, Liu C-M, Pombeiro AJL. Dinuclear
Mn(II, II) complexes: magnetic properties and microwave assisted oxidation of alco-
hols. Dalton Trans. 2014;43:3966–3977. (c) Alexandru M, Cazacu M, Arvinte A.
μ-Chlorido-bridged dimanganese(II) complexes of the Schiff base derived from [2
+2] condensation of 2,6-diformyl-4-methylphenol and 1,3-bis(3-aminopropyl)
tetramethyldisiloxane: structure, magnetism, electrochemical behaviour, and catalytic
oxidation of secondary alcohols. Eur J Inorg Chem. 2014;120–131.
65. Hayashi M, Sasano Y, Nagasawa S, Shibuya M, Iwabuchi Y. 9-Azanoradamantane
N-oxyl (Nor-AZADO): a highly active organocatalyst for alcohol oxidation. Chem
Pharm Bull. 2011;59:1570–1573.
66. (a) Sheldon RA, Arends IWCE. Catalytic oxidations mediated by metal ions and
nitroxyl radicals. J Mol Catal A Chem. 2006;251:200–214. (b) Tebben L, Studer A.
Nitroxides: applications in synthesis and in polymer chemistry. Angew Chem Int Ed
Engl. 2011;50:5034–5068. (c) Wertz S, Studer A. Nitroxide-catalyzed transition-
metal-free aerobic oxidation processes. Green Chem. 2013;15:3116–3134. (d) Ishii Y,
Sakaguchi S, Iwahama T. Innovation of hydrocarbon oxidation with molecular oxygen
and related reactions. Adv Synth Catal. 2001;343:393–427. (e) Recupero F, Punta C.
Free radical functionalization of organic compounds catalyzed by
N-hydroxyphthalimide. Chem Rev. 2007;107:3800–3842. (f ) Coseri S. Phthalimide-
N-oxyl (PINO) radical, a powerful catalytic agent: its generation and versatility towards
various organic substrates. Catal Rev. 2009;51:218–292.
67. (a) Gartshore CJ, Lupton DW. Readily accessible oxazolidine nitroxyl radicals: bifunc-
tional cocatalysts for simplified copper based aerobic oxidation. Adv Synth Catal.
2010;352:3321–3328. (b) Cheng L, Wang J, Wang M, Wu Z. Mechanistic insight into
alcohol oxidation mediated by an efficient green CuII-bipy catalyst with and without
TEMPO by density functional methods. Dalton Trans. 2010;39:5377–5387. (c) Lu Z,
Ladrak T, Roubeau O. Selective, catalytic aerobic oxidation of alcohols using CuBr2
and bifunctional triazine-based ligands containing both a bipyridine and a TEMPO
group. Dalton Trans. 2009;3559–3570. (d) Gamez P, Arends IWCE, Sheldon RA,
Reedijk J. Room temperature aerobic copper-catalysed selective oxidation of primary
alcohols to aldehydes. Adv Synth Catal. 2004;346:805–811. (e) Figiel PJ, Leskelä M,
Repo T. TEMPO-copper(II) diimine-catalysed oxidation of benzylic alcohols in
aqueous media. Adv Synth Catal. 2007;349:1173–1179. (f ) Figiel PJ, Sibaouih A,
Ahmad JU. Aerobic oxidation of benzylic alcohols in water by 2,2,6,6-
tetramethylpiperidine-1-oxyl(TEMPO)/copper(II) 2-N-arylpyrrolecarbaldimino
complexes. Adv Synth Catal. 2009;351:2625–2632. (g) Salinas Uber J, Vogels Y, van
den Helder D. Pyrazole-based ligands for the [copper–TEMPO]-mediated oxidation
of benzyl alcohol to benzaldehyde and structures of the Cu coordination compounds.
Eur J Inorg Chem. 2007;4197–4206. (h) Jiang N, Ragauskas AJ. Copper(II)-catalyzed
aerobic oxidation of primary alcohols to aldehydes in ionic liquid [bmpy]PF6. Org Lett.
2005;7:3689–3692. (i) Striegler S. Aerobic oxidation of primary alcohols under mild
aqueous conditions promoted by a dinuclear copper(II) complex. Tetrahedron.
2006;62:9109–9114. ( j) Mannam S, Alamsetti SK, Sekar G. Aerobic, chemoselective
oxidation of alcohols to carbonyl compounds catalyzed by a DABCO-copper complex
under mild conditions. Adv Synth Catal. 2007;349:2253–2258. (k) Yang G, Ma J,
Wang W. Heterogeneous Cu–Mn oxides mediate efficiently TEMPO-catalyzed
ARTICLE IN PRESS

162 Maximilian N. Kopylovich et al.

aerobic oxidation of alcohols. Catal Lett. 2006;112:83–87. (l) Yang G, Zhu W,


Zhang P. Recyclable carbon supported copper-manganese oxide for selective aerobic
oxidation of alcohols in combination with 2,2,6,6-tetramethylpiperidyl-1-oxyl under
neutral condition. Adv Synth Catal. 2008;350:542–546. (m) Mahmudov KT,
Kopylovich MN, Sabbatini A. Cooperative metal–ligand assisted E/Z isomerization
and cyano activation at CuII and CoII complexes of arylhydrazones of active methylene
nitriles. Inorg Chem. 2014;53:9946–9958.
68. Steves JE, Stahl SS. Copper(I)/ABNO-catalyzed aerobic alcohol oxidation: alleviating
steric and electronic constraints of Cu/TEMPO catalyst systems. J Am Chem Soc.
2013;135:15742–15745.
69. Sasano Y, Nagasawa S, Yamazaki M, Shibuya M, Park J, Iwabuchi Y. Highly
chemoselective aerobic oxidation of amino alcohols into amino carbonyl compounds.
Angew Chem Int Ed Engl. 2014;53:3236–3240.
70. (a) Wang N, Liu R, Chen J, Liang X. NaNO2-activated, iron-TEMPO catalyst system
for aerobic alcohol oxidation under mild conditions. Chem Commun.
2005;42:5322–5324. (b) Wang X, Liang X. Aerobic oxidation of alcohols to carbonyl
compounds catalyzed by Fe(NO3)3/4-OH-TEMPO under mild conditions. Chin
J Catal. 2008;29:935–939. (c) Yin W, Chu C, Lu Q, Tao J, Liang X, Liu R. Iron chlo-
ride/4-acetamido-TEMPO/sodium nitrite-catalyzed aerobic oxidation of primary
alcohols to the aldehydes. Adv Synth Catal. 2010;352:113–118. (d) Miao C-X,
Wang J-Q, Yu B. Synthesis of bimagnetic ionic liquid and application for selective aer-
obic oxidation of aromatic alcohols under mild conditions. Chem Commun.
2011;47:2697–2699. (e) Ma S, Liu J, Li S. Development of a general and practical iron
nitrate/TEMPO-catalyzed aerobic oxidation of alcohols to aldehydes/ketones: cataly-
sis with table salt. Adv Synth Catal. 2011;353:1005–1017.
71. (a) Cecchetto A, Fontana F, Minisci F, Recupero F. Efficient Mn–Cu and Mn–Co–
TEMPO-catalysed oxidation of alcohols into aldehydes and ketones by oxygen under
mild conditions. Tetrahedron Lett. 2001;42:6651–6653. (b) Minisci F, Recupero F,
Rodinò M, Sala M, Schneider A. A convenient nitroxyl radical catalyst for the selective
oxidation of primary and secondary alcohols to aldehydes and ketones by O2 and
H2O2 under mild conditions. Org Process Res Dev. 2003;7:794–798. (c) Minisci F,
Recupero F, Cecchetto A. Mechanisms of the aerobic oxidation of alcohols to alde-
hydes and ketones, catalysed under mild conditions by persistent and non-persistent
nitroxyl radicals and transition metal salts—polar, enthalpic, and captodative effects.
Eur J Org Chem. 2004;109–119. (d) Gilhespy M, Lok M, Baucherel X. Polymer-
supported nitroxyl radical catalyst for selective aerobic oxidation of primary alcohols
to aldehydes. Chem Commun. 2005;1085–1086. (e) Gilhespy M, Lok M,
Baucherel X. Polymer-supported nitroxyl radical catalysts for the hypochlorite and aer-
obic oxidation of alcohols. Catal Today. 2006;117:114–119.
72. Ben-Daniel R, Alsters P, Neumann R. Selective aerobic oxidation of alcohols with a
combination of a polyoxometalate and nitroxyl radical as catalysts. J Org Chem.
2001;66:8650–8653.
73. Du Z, Ma J, Ma H, Wang M, Huang Y, Xu J. Vanadyl sulfate: a simple catalyst for
oxidation of alcohols with molecular oxygen in combination with 2,2,6,6-tetramethyl-
piperidyl-1-oxyl. Catal Commun. 2010;11:732–735.
74. (a) Shibuya M, Osada Y, Sasano Y, Tomizawa M, Iwabuchi Y. Highly efficient,
organocatalytic aerobic alcohol oxidation. J Am Chem Soc. 2011;133:6497–6500.
(b) Bogart JA, Lee HB, Boreen MA, Jun M, Schelter EJ. Fine-tuning the oxidative
ability of persistent radicals: electrochemical and computational studies of substituted
2-pyridylhydroxylamines. J Org Chem. 2013;78:6344–6349. (c) Hamada S,
Furuta T, Wada Y, Kawabata T. Chemoselective oxidation by electronically tuned
nitroxyl radical catalysts. Angew Chem Int Ed Engl. 2013;52:8093–8097.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 163

75. Zhu J, Wang P, Lu M. Synthesis of novel magnetic silica supported hybrid ionic liquid
combining TEMPO and polyoxometalate and its application for selective oxidation of
alcohols. RSC Adv. 2012;2:8265–8268.
76. Beejapur HA, Campisciano V, Franchi P, Lucarini M, Giacalone F, Gruttadauria M.
Fullerene as a platform for recyclable TEMPO organocatalysts for the oxidation of alco-
hols. ChemCatChem. 2014;6:2419–2424.
77. Wang L, Li J, Lv Y, Zhao G, Gao S. Selective aerobic oxidation of alcohols catalyzed by
iron chloride hexahydrate/TEMPO in the presence of silica gel. Appl Organomet Chem.
2012;26:37–43.
78. Wang L, Li J, Zhao X, Lv Y, Zhang H, Gao S. An efficient and scalable room temper-
ature aerobic alcohol oxidation catalyzed by iron chloride hexahydrate/mesoporous
silica supported TEMPO. Tetrahedron. 2013;69:6041–6045.
79. Karimi B, Badreh E. SBA-15-functionalized TEMPO confined ionic liquid: an effi-
cient catalyst system for transition-metal-free aerobic oxidation of alcohols with
improved selectivity. Org Biomol Chem. 2011;9:4194–4198.
80. Obermayer D, Balu AM, Romero AA, Goessler W, Luque R, Kappe CO.
Nanocatalysis in continuous flow: supported iron oxide nanoparticles for the hetero-
geneous aerobic oxidation of benzyl alcohol. Green Chem. 2013;15:1530–1537.
81. Ishida T, Watanabe H, Takei T, Hamasaki A, Tokunaga M, Haruta M. Metal oxide-
catalyzed ammoxidation of alcohols to nitriles and promotion effect of gold
nanoparticles for one-pot amide synthesis. Appl Catal A Gen. 2012;425–426:85–90.
82. Liu J, Ma S. Aerobic oxidation of indole carbinols using
Fe(NO3)3 9H2O/TEMPO/NaCl as catalysts. Org Biomol Chem. 2013;11:4186–4193.
83. Sedai B, Dı́az-Urrutia C, Baker RT, Wu R, Silks LAP, Hanson SK. Aerobic oxidation
of β-1 lignin model compounds with copper and oxovanadium catalysts. ACS Catal.
2013;3:3111–3122.
84. Painter RM, Pearson DM, Waymouth RM. Selective catalytic oxidation of glycerol to
dihydroxyacetone. Angew Chem Int Ed Engl. 2010;49:9456–9459.
85. Spatareanu A, Bercea M, Budtova T, Harabagiu V, Sacarescu L, Coseri S. Synthesis,
characterization and solution behaviour of oxidized pullulan. Carbohydr Polym.
2014;111:63–71.
86. Amberg M, D€ onges M, Stapf G, Hartung J. Formation of 3-acyloxy-γ-
butyrolactones from 4-pentenols in vanadium-catalyzed oxidations. Tetrahedron.
2014;70:5321–5331.
87. Liu J, Ma S. Iron-catalyzed aerobic oxidation of allylic alcohols: the issue of C]C bond
isomerization. Org Lett. 2013;15:5150–5153.
88. Liu J, Ma S. Room temperature Fe(NO3)39H2O/TEMPO/NaCl-catalyzed aerobic
oxidation of homopropargylic alcohols. Tetrahedron. 2013;69:10161–10167.
89. Ouyang X-H, Song R-J, Li J-H. Iron-catalyzed oxidative 1,2-carboacylation of acti-
vated alkenes with alcohols: a tandem route to 3-(2-oxoethyl)indolin-2-ones. Eur J Org
Chem. 2014;2014:3395–3401.
90. Ben-David H, Iron MA, Neumann R. Platinum complexes of cationic ligands for the
aerobic oxidation of “inert” perfluoro-substituted alcohols. Chem Commun.
2013;49:1720–1722.
91. Bowman RK, Brown AD, Cobb JH. Synthesis of HCV replicase inhibitors: base-
catalyzed synthesis of protected α-hydrazino esters and selective aerobic oxidation with
catalytic Pt/Bi/C for synthesis of imidazole-4,5-dicarbaldehyde. J Org Chem.
2013;78:11680–11690.
92. (a) Duschek A, Kirsch SF. 2-Iodoxybenzoic acid—a simple oxidant with a dazzling
array of potential applications. Angew Chem Int Ed Engl. 2011;50:1524–1552.
(b) Satam V, Harad A, Rajule R, Pati H. 2-Iodoxybenzoic acid (IBX): an efficient
hypervalent iodine reagent. Tetrahedron. 2010;66:7659–7706. (c) Uyanik M,
ARTICLE IN PRESS

164 Maximilian N. Kopylovich et al.

Ishihara K. Hypervalent iodine-mediated oxidation of alcohols. Chem Commun.


2009;2086–2099. (d) Zhdankin VV. Organoiodine(V) reagents in organic synthesis.
J Org Chem. 2011;76:1185–1197.
93. Thottumkara AP, Bowsher MS, Vinod TK. In situ generation of o-iodoxybenzoic acid
(IBX) and the catalytic use of it in oxidation reactions in the presence of oxone as a
co-oxidant. Org Lett. 2005;7:2933–2936.
94. Xu S, Itto K, Satoh M, Arimoto H. Unexpected dehomologation of primary alcohols
to one-carbon shorter carboxylic acids using o-iodoxybenzoic acid (IBX). Chem
Commun. 2014;50:2758–2761.
95. Li X-H, Meng X-G, Liu Y, Peng X. Direct oxidation of secondary alcohol to ester by
performic acid. Green Chem. 2013;15:3332–3336.
96. Miura T, Nakashima K, Tada N, Itoh A. An effective and catalytic oxidation using
recyclable fluorous IBX. Chem Commun. 2011;47:1875–1877.
97. Majee A, Kundu SK, Santra S, Hajra A. Combination of NH2OHHCl and NaIO4: a
new and mild oxidizing agent for selective oxidation of alcohols to carbonyl com-
pounds. Tetrahedron Lett. 2012;53:4433–4435.
98. Bartlett SL, Beaudry CM. High-yielding oxidation of β-hydroxyketones to
β-diketones using o-iodoxybenzoic acid. J Org Chem. 2011;76:9852–9855.
99. Berini C, Brayton DF, Mocka C, Navarro O. Homogeneous, anaerobic (N-
Heterocyclic Carbene)Pd or Ni catalyzed oxidation of secondary alcohols at mild
temperatures. Org Lett. 2009;11:4244–4247.
100. Hamasaki A, Kuwada H, Tokunaga M. tert-Butylnitrite as a convenient and easy-
removable oxidant for the conversion of benzylic alcohols to ketones and aldehydes.
Tetrahedron Lett. 2012;53:811–814.
101. Holan M, Jahn U. Anaerobic nitroxide-catalyzed oxidation of alcohols using the
NO+/NO Redox Pair. Org Lett. 2013;16:58–61.
102. Fernandes RA, Bethi V. An expedient osmium(VI)/K3Fe(CN)6-mediated selective
oxidation of benzylic, allylic and propargylic alcohols. RSC Adv. 2014;4:
40561–40568.
103. Hayashi M, Shibuya M, Iwabuchi Y. Oxidation of alcohols to carbonyl compounds
with diisopropyl azodicarboxylate catalyzed by nitroxyl radicals. J Org Chem.
2012;77:3005–3009.
104. Zhu Y, Zhao B, Shi Y. Highly efficient Cu(I)-catalyzed oxidation of alcohols to
ketones and aldehydes with diaziridinone. Org Lett. 2013;15:992–995.
105. William JM, Kuriyama M, Onomura O. Electrochemical oxidation of 1,2-diols to
α-hydroxyketones in water. Electrochemistry. 2013;81:374–376.
106. William JM, Kuriyama M, Onomura O. Boronic acid-catalyzed selective oxi-
dation of 1,2-diols to α-hydroxy ketones in water. Adv Synth Catal. 2014;356:
934–940.
107. Nielsen M, Kammer A, Cozzula D, Junge H, Gladiali S, Beller M. Efficient hydrogen
production from alcohols under mild reaction conditions. Angew Chem Int Ed Engl.
2011;50:9593–9597.
108. (a) Baratta W, Bossi G, Putignano E, Rigo P. Pincer and diamine Ru and Os
diphosphane complexes as efficient catalysts for the dehydrogenation of alcohols to
ketones. Chem Eur J. 2011;17:3474–3481. (b) Prades A, Peris E, Albrecht M. Oxida-
tions and oxidative couplings catalyzed by triazolylidene ruthenium complexes.
Organometallics. 2011;30:1162–1167. (c) Zhang J, Balaraman E, Leitus G,
Milstein D. Electron-rich PNP- and PNN-type ruthenium(II) hydrido borohydride
pincer complexes. Synthesis, structure, and catalytic dehydrogenation of alcohols
and hydrogenation of esters. Organometallics. 2011;30:5716–5724. (d) Shahane S,
Fischmeister C, Bruneau C. Acceptorless ruthenium catalyzed dehydrogenation of
alcohols to ketones and esters. Catal Sci Technol. 2012;2:1425–1428. (e) Schley ND,
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 165

Dobereiner GE, Crabtree RH. Oxidative synthesis of amides and pyrroles via
dehydrogenative alcohol oxidation by ruthenium diphosphine diamine complexes.
Organometallics. 2011;30:4174–4179.
109. (a) Royer AM, Rauchfuss TB, Gray DL. Organoiridium pyridonates and their role in
the dehydrogenation of alcohols. Organometallics. 2010;29:6763–6768. (b) Musa S,
Shaposhnikov I, Cohen S, Gelman D. Ligand–metal cooperation in PCP pincer com-
plexes: rational design and catalytic activity in acceptorless dehydrogenation of alco-
hols. Angew Chem Int Ed Engl. 2011;50:3533–3537. (c) Kawahara K, Fujita K,
Yoshida T, Imori Y, Yamaguchi R. Dehydrogenative oxidation of primary and sec-
ondary alcohols catalyzed by a Cp*Ir complex having a functional C,N-chelate
ligand. Org Lett. 2011;13:2278–2281. (d) Kawahara R, Fujita K, Yamaguchi R.
Dehydrogenative oxidation of alcohols in aqueous media using water-soluble and
reusable Cp*Ir catalysts bearing a functional bipyridine ligand. J Am Chem Soc.
2012;134:3643–3646. (e) Kawahara R, Fujita K, Yamaguchi R. Cooperative catalysis
by iridium complexes with a bipyridonate ligand: versatile dehydrogenative oxidation
of alcohols and reversible dehydrogenation–hydrogenation between 2-propanol and
acetone. Angew Chem Int Ed Engl. 2012;51:12790–12794. (f ) Polukeev AV,
Petrovskii PV, Peregudov AS, Ezernitskaya MG, Koridze AA. Dehydrogenation
of alcohols by bis(phosphinite) benzene based and bis(phosphine) ruthenocene based
iridium pincer complexes. Organometallics. 2013;32(4):1000–1015.
110. Zhang D-D, Chen X-K, Liu H-L, Huang X-R. A computational mechanistic study
of pH-dependent alcohol dehydrogenation catalyzed by a novel [C, N] or [C, C]
cyclometalated Cp*Ir complex in aqueous solution. New J Chem. 2014; 38:3862–3873.
111. Zeng G, Sakaki S, Fujita K, Sano H, Yamaguchi R. Efficient catalyst for acceptorless
alcohol dehydrogenation: interplay of theoretical and experimental studies. ACS Catal.
2014;4(3):1010–1020.
112. Aliende C, Pérez-Manrique M, Jalón FA, Manzano BR, Rodrı´guez AM, Espino G.
Arene ruthenium complexes as versatile catalysts in water in both transfer hydrogena-
tion of ketones and oxidation of alcohols. Selective deuterium labeling of rac-1-
phenylethanol. Organometallics. 2012;31:6106–6123.
113. Musa S, Ackermann L, Gelman D. Dehydrogenative cross-coupling of primary and
secondary alcohols. Adv Synth Catal. 2013;355:3077–3080.
114. (a) Yang J, Liu X, Meng D-L. Efficient iron-catalyzed direct β-alkylation of secondary
alcohols with primary alcohols. Adv Synth Catal. 2012;354:328–334. (b) Xu C,
Goh LY, Pullarkat SA. Efficient iridium-thioether-dithiolate catalyst for β-alkylation
of alcohols and selective imine formation via N-alkylation reactions. Organometallics.
2011;30:6499–6502. (c) Miura T, Kose O, Li F, Kai S, Saito S. CuI/H2/NaOH-
catalyzed cross-coupling of two different alcohols for carbon–carbon bond formation:
“borrowing hydrogen”? Chem Eur J. 2011;17:11146–11151. (d) Kose O, Saito S.
Cross-coupling reaction of alcohols for carbon-carbon bond formation using
pincer-type NHC/palladium catalysts. Org Biomol Chem. 2010;8:896–900.
(e) Gnanamgari D, Sauer ELO, Schley ND, Butler C, Incarvito CD, Crabtree RH.
Iridium and ruthenium complexes with chelating N-heterocyclic carbenes: efficient
catalysts for transfer hydrogenation, β-alkylation of alcohols, and N-alkylation of
amines. Organometallics. 2008;28:321–325.
115. (a) Shimizu K, Sato R, Satsuma A. Direct C–C cross-coupling of secondary and pri-
mary alcohols catalyzed by a γ-alumina-supported silver subnanocluster. Angew Chem
Int Ed Engl. 2009;121:4042–4046. (b) Shimizu K, Sato R, Satsuma A. Direct C–C
cross-coupling of secondary and primary alcohols catalyzed by a γ-alumina-supported
silver subnanocluster. Angew Chem Int Ed Engl. 2009;48:3982–3986.
116. Zhang G, Hanson SK. Cobalt-catalyzed acceptorless alcohol dehydrogenation: synthe-
sis of imines from alcohols and amines. Org Lett. 2013;15:650–653.
ARTICLE IN PRESS

166 Maximilian N. Kopylovich et al.

117. Zhang G, Vasudevan KV, Scott BL, Hanson SK. Understanding the mechanisms of
cobalt-catalyzed hydrogenation and dehydrogenation reactions. J Am Chem Soc.
2013;135:8668–8681.
118. Kamitani M, Ito M, Itazaki M, Nakazawa H. Effective dehydrogenation of
2-pyridylmethanol derivatives catalyzed by an iron complex. Chem Commun.
2014;50:7941–7944.
119. Kirillov AM, Shul’pin GB. Pyrazinecarboxylic acid and analogs: highly efficient
co-catalysts in the metal-complex-catalyzed oxidation of organic compounds. Coord
Chem Rev. 2013;257:732–754.
120. Song H, Kang B, Hong SH. Fe-catalyzed acceptorless dehydrogenation of secondary
benzylic alcohols. ACS Catal. 2014;4:2889–2895.
121. (a) Mitsudome T, Mikami Y, Funai H, Mizugaki T, Jitsukawa K, Kaneda K. Oxidant-
free alcohol dehydrogenation using a reusable hydrotalcite-supported silver nanoparti-
cle catalyst. Angew Chem Int Ed Engl. 2008;47:138–141. (b) Fang W, Zhang Q, Chen J,
Deng W, Wang Y. Gold nanoparticles on hydrotalcites as efficient catalysts for oxidant-
free dehydrogenation of alcohols. Chem Commun. 2010;46:1547–1549. (c) Zhang Q,
Deng W, Wang Y. Effect of size of catalytically active phases in the dehydrogenation of
alcohols and the challenging selective oxidation of hydrocarbons. Chem Commun.
2011;47:9275–9292. (d) Fang W, Chen J, Zhang Q, Deng W, Wang Y. Hydrotalcite-
supported gold catalyst for the oxidant-free dehydrogenation of benzyl alcohol: studies
on support and gold size effects. Chem Eur J. 2011;17:1247–1256.
122. Chen J, Fang W, Zhang Q, Deng W, Wang Y. A comparative study of size effects in the
Au-catalyzed oxidative and non-oxidative dehydrogenation of benzyl alcohol. Chem
Asian J. 2014;9:2187–2196.
123. Yi J, Miller JT, Zemlyanov DY. A reusable unsupported rhenium nanocrystalline cat-
alyst for acceptorless dehydrogenation of alcohols through γ-C–H activation. Angew
Chem Int Ed Engl. 2014;53:833–836.
124. Sawama Y, Morita K, Yamada T. Rhodium-on-carbon catalyzed hydrogen scavenger-
and oxidant-free dehydrogenation of alcohols in aqueous media. Green Chem.
2014;16:3439–3443.
125. Kon K, Hakim Siddiki SMA, Shimizu K. Size- and support-dependent Pt nanocluster
catalysis for oxidant-free dehydrogenation of alcohols. J Catal. 2013;304:63–71.
126. Shimizu K, Kon K, Shimura K, Hakim SSMA. Acceptor-free dehydrogenation of sec-
ondary alcohols by heterogeneous cooperative catalysis between Ni nanoparticles and
acid–base sites of alumina supports. J Catal. 2013;300:242–250.
127. Damodara D, Arundhathi R, Likhar PR. Copper nanoparticles from copper aluminum
hydrotalcite: an efficient catalyst for acceptor- and oxidant-free dehydrogenation of
amines and alcohols. Adv Synth Catal. 2014;356:189–198.
128. Shimizu K, Kon K, Seto M, Shimura K, Yamazaki H, Kondo JN. Heterogeneous
cobalt catalysts for the acceptorless dehydrogenation of alcohols. Green Chem.
2013;15:418–424.
129. Li F, Sun C, Wang N. Catalytic acceptorless dehydrogenative coupling of
arylhydrazines and alcohols for the synthesis of arylhydrazones. J Org Chem.
2014;79:8031–8039.
130. Fujita K, Ito W, Yamaguchi R. Dehydrogenative lactonization of diols in aqueous
media catalyzed by a water-soluble iridium complex bearing a functional bipyridine
ligand. ChemCatChem. 2014;6:109–112.
131. Balaraman E, Khaskin E, Leitus G, Milstein D. Catalytic transformation of alcohols to
carboxylic acid salts and H2 using water as the oxygen atom source. Nat Chem.
2013;5:122–125.
132. (a) Gnanaprakasam B, Milstein D. Synthesis of amides from esters and amines with
liberation of H2 under neutral conditions. J Am Chem Soc. 2011;133:1682–1685.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 167

(b) Gnanaprakasam B, Ben-David Y, Milstein D. Ruthenium pincer-catalyzed acyla-


tion of alcohols using esters with liberation of hydrogen under neutral conditions. Adv
Synth Catal. 2010;352:3169–3173. (c) Balaraman E, Gnanaprakasam B, Shimon LJW,
Milstein D. Direct hydrogenation of amides to alcohols and amines under mild condi-
tions. J Am Chem Soc. 2010;132:16756–16758. (d) Gunanathan C, Ben-David Y,
Milstein D. Direct synthesis of amides from alcohols and amines with liberation of
H2. Science. 2007;317:790–792. (e) Zhang J, Leitus G, Ben-David Y, Milstein D. Facile
conversion of alcohols into esters and dihydrogen catalyzed by new ruthenium
complexes. J Am Chem Soc. 2005;127:10840–10841. (f ) Gunanathan C, Milstein D.
Metal-ligand cooperation by aromatization–dearomatization: a new paradigm in bond
activation and “Green” catalysis. Acc Chem Res. 2011;44:588–602.
133. Li H, Hall MB. Mechanism of the formation of carboxylate from alcohols and water
catalyzed by a bipyridine-based ruthenium complex: a computational study. J Am Chem
Soc. 2013;136:383–395.
134. Bartoszewicz A, Ahlsten N, Martı́n-Matute B. Enantioselective synthesis of alcohols
and amines by iridium-catalyzed hydrogenation, transfer hydrogenation, and related
processes. Chem Eur J. 2013;19:7274–7302.
135. Arita S, Koike T, Kayaki Y, Ikariya T. Aerobic oxidative kinetic resolution of racemic
secondary alcohols with chiral bifunctional amido complexes. Angew Chem Int Ed Engl.
2008;47:2447–2449.
136. William JM, Kuriyama M, Onomura O. An efficient method for selective oxidation of
1,2-diols in water catalyzed by Me2SnCl2. RSC Adv. 2013;3:19247–19250.
137. Bettucci L, Bianchini C, Filippi J, Lavacchi A, Oberhauser W. Chemoselective
aerobic diol oxidation by palladium(II)-pyridine catalysis. Eur J Inorg Chem.
2011;1797–1805.
138. Bettucci L, Bianchini C, Oberhauser W, Hsiao T-H, Lee HM. Chemoselective aerobic
oxidation of unprotected diols catalyzed by Pd-(NHC) (NHC¼N-heterocyclic
carbene) complexes. J Mol Catal A Chem. 2010;322:63–72.
139. De Crisci AG, Chung K, Oliver AG, Solis-Ibarra D, Waymouth RM. Chemoselective
oxidation of polyols with chiral palladium catalysts. Organometallics.
2013;32:2257–2266.
140. Chung K, Banik SM, De Crisci AG. Chemoselective Pd-catalyzed oxidation of
polyols: synthetic scope and mechanistic studies. J Am Chem Soc. 2013;135:7593–7602.
141. Ebner DC, Bagdanoff JT, Ferreira EM. The palladium-catalyzed aerobic kinetic reso-
lution of secondary alcohols: reaction development, scope, and applications. Chem Eur
J. 2009;15:12978–12992.
142. Suzuki T, Morita K, Matsuo Y, Hiroi K. Catalytic asymmetric oxidative lactonizations
of meso-diols using a chiral iridium complex. Tetrahedron Lett. 2003;44: 2003–2006.
143. Moritani J, Hasegawa Y, Kayaki Y, Ikariya T. Aerobic oxidative desymmetrization of
meso-diols with bifunctional amidoiridium catalysts bearing chiral N-sulfonyldiamine
ligands. Tetrahedron Lett. 2014;55:1188–1191.
144. Manzini S, Urbina-Blanco CA, Nolan SP. Chemoselective oxidation of secondary
alcohols using a ruthenium phenylindenyl complex. Organometallics. 2013;32:660–664.
145. Mizoguchi H, Uchida T, Katsuki T. Ruthenium-catalyzed oxidative kinetic resolution
of unactivated and activated secondary alcohols with air as the hydrogen acceptor at
room temperature. Angew Chem Int Ed Engl. 2014;53:3178–3182.
146. Onomura O, Arimoto H, Matsumura Y, Demizu Y. Asymmetric oxidation of 1,2-
diols using N-bromosuccinimide in the presence of chiral copper catalyst. Tetrahedron
Lett. 2007;48:8668–8672.
147. Berini C, Winkelmann OH, Otten J, Vicic DA, Navarro O. Rapid and selective cat-
alytic oxidation of secondary alcohols at room temperature by using (N-heterocyclic
carbene)-Ni0 systems. Chem Eur J. 2010;16:6857–6860.
ARTICLE IN PRESS

168 Maximilian N. Kopylovich et al.

148. Zhang Y, Zhou Q, Ma W, Zhao J. Enantioselective oxidation of racemic secondary


alcohols catalyzed by chiral Mn(III)-salen complex with sodium hypochlorite as oxi-
dant. Catal Commun. 2014;45:114–117.
149. Bera PK, Maity NC, Abdi SHR, Khan N-uH, Kureshy RI, Bajaj HC. Macrocyclic
Mn(III) salen complexes as recyclable catalyst for oxidative kinetic resolution of sec-
ondary alcohols. Appl Catal A Gen. 2013;467:542–551.
150. Willemsen JS, van Hest JCM, Rutjes FPJT. Potassium formate as a small molecule
switch: controlling oxidation-reduction behaviour in a two-step sequence. Chem
Commun. 2013;49:3143–3145.
151. Hamid MHSA, Slatford PA, Williams JMJ. Borrowing hydrogen in the activation of
alcohols. Adv Synth Catal. 2007;349:1555–1575.
152. Liu Q, Wu L, G€ ulak S, Rockstroh N, Jackstell R, Beller M. Towards a sustainable syn-
thesis of formate salts: combined catalytic methanol dehydrogenation and bicarbonate
hydrogenation. Angew Chem Int Ed Engl. 2014;53:7085–7088.
153. Moromi SK, Siddiki HSMA, Ali MA, Kon K, Shimizu K. Acceptorless
dehydrogenative coupling of primary alcohols to esters by heterogeneous Pt catalysts.
Catal Sci Technol. 2014;4:3631–3635.
154. Azizi K, Karimi M, Nikbakht F, Heydari A. Direct oxidative amidation of benzyl alco-
hols using EDTA@Cu(II) functionalized superparamagnetic nanoparticles. Appl Catal
A Gen. 2014;482:336–343.
155. Fries P, Halter D, Kleinschek A, Hartung J. Functionalized tetrahydrofurans from
alkenols and olefins/alkynes via aerobic oxidation-radical addition cascades. J Am Chem
Soc. 2011;133:3906–3912.
156. Dai X, Xie M, Meng S, Fu X, Chen S. Coupled systems for selective oxidation of
aromatic alcohols to aldehydes and reduction of nitrobenzene into aniline using
CdS/g-C3N4 photocatalyst under visible light irradiation. Appl Catal B Environ.
2014;158–159:382–390.
157. Saito T, Isogai A. TEMPO-mediated oxidation of native cellulose. The effect of oxi-
dation conditions on chemical and crystal structures of the water-insoluble fractions.
Biomacromolecules. 2004;5:1983–1989.
158. Wolfel R, Taccardi N, Bosmann A, Wasserscheid P. Selective catalytic conversion of
biobased carbohydrates to formic acid using molecular oxygen. Green Chem.
2011;13:2759–2763.
159. Zhang J, Sun M, Liu X, Han Y. Catalytic oxidative conversion of cellulosic biomass to
formic acid and acetic acid with exceptionally high yields. Catal Today. 2014; 233:77–82.
160. Fellay C, Dyson PJ, Laurenczy G. A viable hydrogen-storage system based on selective
formic acid decomposition with a ruthenium catalyst. Angew Chem Int Ed Engl.
2008;47:3966–3968.
161. Boddien A, Loges B, Gartner F, et al. Iron-catalyzed hydrogen production from formic
acid. J Am Chem Soc. 2010;132:8924–8934.
162. Weingarten R, Conner WC, Huber GW. Production of levulinic acid from cellulose
by hydrothermal decomposition combined with aqueous phase dehydration with a
solid acid catalyst. Energy Environ Sci. 2012;5:7559–7574.
163. Van de Vyver S, Thomas J, Geboers J, et al. Catalytic production of levulinic acid from
cellulose and other biomass-derived carbohydrates with sulfonated hyperbranched
poly(arylene oxindole)s. Energy Environ Sci. 2011;4:3601–3610.
164. Fukuoka A, Dhepe PL. Catalytic conversion of cellulose into sugar alcohols. Angew
Chem Int Ed Engl. 2006;45:5161–5163.
165. Ji N, Zhang T, Zheng MY, et al. Direct catalytic conversion of cellulose into ethylene
glycol using nickel-promoted tungsten carbide catalysts. Angew Chem Int Ed Engl.
2008;47:8510–8513.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 169

166. Holm MS, Saravanamurugan S, Taarning E. Conversion of sugars to lactic acid deriv-
atives using heterogeneous zeotype catalysts. Science. 2010;328:602–605.
167. Zhang JZ, Liu X, Sun M, Ma XH, Han Y. Direct conversion of cellulose to glycolic
acid with a phosphomolybdic acid catalyst in a water medium. ACS Catal.
2012;2:1698–1702.
168. Zhang JZ, Liu X, Hedhili MN, Zhu YH, Han Y. Highly selective and complete con-
version of cellobiose to gluconic acid over Au/Cs2HPW12O40 nanocomposite catalyst.
ChemCatChem. 2011;3:1294–1298.
169. Carrettin S, McMorn P, Johnston P, Griffin K, Hutchings GJ. Selective oxidation of
glycerol to glyceric acid using a gold catalyst in aqueous sodium hydroxide. Chem
Commun. 2002;696–697.
170. Brett L, He Q, Hammond C, et al. Selective oxidation of glycerol by highly active
bimetallic catalysts at ambient temperature under base-free conditions. Angew Chem
Int Ed Engl. 2011;50:10136–10139.
171. Zhang Y, Zhao G, Zhang Y, Huang X. Highly efficient visible-light-driven
photoelectro-catalytic selective aerobic oxidation of biomass alcohols to aldehydes.
Green Chem. 2014;16:3860–3869.
172. Beauchamp J, Gross PG, Vieille C. Characterization of Thermotoga Maritima glycerol
dehydrogenase for the enzymatic production of dihydroxyacetone. Appl Microbiol
Biotechnol. 2014;98:7039–7050.
173. Ojeda M, Balu AM, Romero AA, et al. MAGBONS: novel magnetically separable car-
bonaceous nanohybrids from porous polysaccharides. ChemCatChem.
2014;6:2847–2853.
174. Davis SE, Houk LR, Tamargo EC, Datye AK, Davis RJ. Oxidation of
5-hydroxymethylfurfural over supported Pt, Pd and Au catalysts. Catal Today.
2011;160:55–60.
175. Villa A, Schiavoni M, Campisi S, Veith GM, Prati L. Pd-modified Au on carbon as an
effective and durable catalyst for the direct oxidation of HMF to 2,5-furandicarboxylic
acid. ChemSusChem. 2013;6:609–612.
176. Ni M, Leung DYC, Leung MKH. A review on reforming bio-ethanol for hydrogen
production. Int J Hydrog Energy. 2007;32:3238–3247.
177. Nielsen M, Alberico E, Baumann W, et al. Low-temperature aqueous-phase methanol
dehydrogenation to hydrogen and carbon dioxide. Nature. 2013;495:85–89.
178. Valant A, Garron A, Bion N, Duprez D, Epron F. Effect of higher alcohols on the per-
formances of a 1%Rh/MgAl2O4/Al2O3 catalyst for hydrogen production by crude bio-
ethanol steam reforming. Int J Hydrog Energy. 2011;36:311–318.
179. Duprez D. Selective steam reforming of aromatic compounds on metal-catalysts. Appl
Catal A Gen. 1992;82:111–157.
180. Hung CC, Chen SL, Liao YK, Chen CH, Wang JH. Oxidative steam reforming of eth-
anol for hydrogen production on M/Al2O3. Int J Hydrog Energy. 2012;37:4955–4966.
181. Qu YC, Wei XL, Zuo Y, et al. A novel catalyst-free process for producing hydrogen
and carboxylate from biomass-derived alcohols. Int J Hydrog Energy.
2014;39:13136–13140.
182. (a) Herrero MA, Kremsner JM, Kappe CO. Nonthermal microwave effects revisited: on
the importance of internal temperature monitoring and agitation in microwave chemis-
try. J Org Chem. 2007;73:36–47. (b) Obermayer D, Kappe CO. On the importance of
simultaneous infrared/fiber-optic temperature monitoring in the microwave-assisted
synthesis of ionic liquids. Org Biomol Chem. 2010;8:114–121. (c) Kappe CO. Controlled
microwave heating in modern organic synthesis. Angew Chem Int Ed Engl.
2004;43:6250–6284. (d) Hoz A, Diaz-Ortiz A, Moreno A. Microwaves in organic
synthesis. Thermal and non-thermal microwave effects. Chem Soc Rev. 2005;34:164–178.
ARTICLE IN PRESS

170 Maximilian N. Kopylovich et al.

183. (a) Farràs P, Maji S, Benet-Buchholz J, Llobet A. Synthesis, characterization, and reac-
tivity of dyad ruthenium-based molecules for light-driven oxidation catalysis. Chem Eur
J. 2013;19:7162–7172. (b) Ohzu S, Ishizuka T, Hirai Y, Fukuzumi S, Kojima T.
Photocatalytic oxidation of organic compounds in water by using ruthenium(ii)–
pyridylamine complexes as catalysts with high efficiency and selectivity. Chem Eur J.
2013;19:1563–1567. (c) Guillo P, Hamelin O, Batat P, Jonusauskas G,
McClenaghan ND, Menage S. Photocatalyzed sulfide oxygenation with water as the
unique oxygen atom source. Inorg Chem. 2012;51:2222–2230.
184. Bellardita M, Loddo V, Palmisano G, Pibiri I, Palmisano L, Augugliaro V. Photo-
catalytic green synthesis of piperonal in aqueous TiO2 suspension. Appl Catal
B Environ. 2014;144:607–613.
185. (a) Palmisano G, Garcı́a-López E, Marcı́ G, et al. Advances in selective conversions by
heterogeneous photocatalysis. Chem Commun. 2010;46:7074–7089. (b) Augugliaro V,
Palmisano L. Green oxidation of alcohols to carbonyl compounds by heterogeneous
photocatalysis. ChemSusChem. 2010;3:1135–1138. (c) Yurdakal S, Palmisano G,
Loddo V, Alagoz O, Augugliaro V, Palmisano L. Selective photocatalytic oxidation
of 4-substituted aromatic alcohols in water with rutile TiO2 prepared at room temper-
ature. Green Chem. 2009;11:510–516.

186. Ozcan L, Yurdakal S, Augugliaro V, et al. Photoelectrocatalytic selective oxidation of
4-methoxybenzyl alcohol in water by TiO2 supported on titanium anodes. Appl Catal
B Environ. 2013;132–133:535–542.
187. Sannino D, Vaiano V, Ciambelli P. Innovative structured VOx/TiO2 photocatalysts
supported on phosphors for the selective photocatalytic oxidation of ethanol to acet-
aldehyde. Catal Today. 2013;205:159–167.
188. Dvoranová D, Barbieriková Z, Brezová V. Radical intermediates in photoinduced
reactions on TiO2. Molecules. 2014;19:17279–17304.
189. Spasiano D, Marotta R, Di Somma I, Andreozzi R, Caprio V. Fe(III)-photocatalytic
partial oxidation of benzyl alcohol to benzaldehyde under UV-solar simulated radia-
tion. Photochem Photobiol Sci. 2013;12:1991–2000.
190. Shen L, Liang S, Wu W, Liang R, Wu L. Multifunctional NH2-mediated zirconium
metal-organic framework as an efficient visible-light-driven photocatalyst for selective
oxidation of alcohols and reduction of aqueous Cr(VI). Dalton Trans. 2013;
42:13649–13657.
191. Gassama A, Hoffmann N. Selective light supported oxidation of hexoses using
air as oxidant—synthesis of tetrahydroxyazepanes. Adv Synth Catal. 2008;
350:35–39.
192. Hallett-Tapley GL, Silvero MJ, Gonzı́lez-Béjar M, Grenier M, Netto-Ferreira JC,
Scaiano JC. Plasmon-mediated catalytic oxidation of sec-phenethyl and benzyl alcohols.
J Phys Chem C. 2011;115:10784–10790.
193. Wannberg J, Wallinder C, Ünl€ usoy M, Sk€ old C, Larhed MJ. One-pot, two-step,
microwave-assisted palladium-catalyzed conversion of aryl alcohols to aryl fluorides
via aryl nonaflates. Org Chem. 2013;78:4184–4189.
194. Takahashi M, Oshima K, Matsubara S. Hydrogen transfer type oxidation of alcohols by
rhodium and ruthenium catalyst under microwave irradiation. Tetrahedron Lett.
2003;44:9201–9203.
195. (a) Figiel PJ, Kopylovich MN, Lasri J, da Silva MFC G, da Silva JJR F, Pombeiro AJL.
Solvent-free microwave-assisted peroxidative oxidation of secondary alcohols to the
corresponding ketones catalyzed by copper(II) 2,4-alkoxy-1,3,5-triazapentadienato com-
plexes. Chem Commun. 2010;46:2766–2768. (b) Kopylovich MN, YaYu Karabach,
da Silva MFC Guedes, Figiel PJ, Lasri J, Pombeiro AJL. Alkoxy-1,3,5-triazapentadien
(e/ato) copper(II) complexes: template formation and applications for the preparation
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 171

of pyrimidines and as catalysts for oxidation of alcohols to carbonyl products. Chem Eur J.
2012;18:899–914.
196. Montagut-Romans A, Boulven M, Lemaire M, Popowycz F. Efficient C-3 reductive
alkylation of 4-hydroxycoumarin by dehydrogenative oxidation of benzylic alcohols
through ruthenium catalysis. New J Chem. 2014;38:1794–1801.
197. Bekhradnia AR, Arshadi S. Efficient selective oxidation of organic substrates using
pyridinum sulfonate halochromate under solvent-free conditions and microwave irra-
diation: experimental and theoretical study. Synth React Inorg Met-Org Nano-Met Chem.
2012;42:705–710.
198. Zhang XY, Qu YY, Wang YY, Fan XS. Synthesis of 1,2-allenic ketones through oxi-
dation of homopropargyl alcohols with CrO3(cat.)/TBHP under MWI. Chin Chem
Lett. 2011;22:268–271.
199. Galica M, Kasprzyk W, Bednarz S, Bogda D. Microwave-assisted oxidation of alcohols
by hydrogen peroxide catalysed by tetrabutylammonium decatungstate. Chem Pap.
2013;67:1240–1244.
200. Landers B, Berini C, Wang C, Navarro O. (N-heterocyclic carbene)-Pd-catalyzed
anaerobic oxidation of secondary alcohols and domino oxidationarylation reactions.
J Org Chem. 2011;76:1390–1397.
201. Winkelmann OH, Navarro O. Microwave-assisted synthesis of N-heterocyclic
carbene-palladium(II) complexes. Adv Synth Catal. 2010;352:212–214.
202. Landers B, Navarro O. Microwave-assisted synthesis of (N-heterocyclic carbene)
Ni(Cp)Cl complexes. Inorg Chim Acta. 2012;380:350–353.
203. Kumar R, Sharma N, Sharma N, Sharma A, Sinha AK. Metal-free activation of H2O2
by synergic effect of ionic liquid and microwave: chemoselective oxidation of benzylic
alcohols to carbonyls and unexpected formation of anthraquinone in aqueous condi-
tions. Mol Divers. 2011;15:687–695.
204. Rajabi F, Balu AM, Toreinia F, Luque R. A versatile supported cobalt(II) complex for
heterogeneously catalysed processes: conventional vs. microwave irradiation protocols.
Catal Sci Technol. 2011;1:1051–1059.
205. Yousefi M, Kamel MH, Kar A, Attar Kar MK. Ni-doped MgFe2O4 nanoferrite as a
novel catalyst for selective oxidation of benzyl alcohol to benzaldehyde. Main Group
Chem. 2013;12:177–184.
206. Suslick KS, Price GJ. Applications of ultrasound to materials chemistry. Annu Rev Mater
Sci. 1999;29:295–326.
207. Chakraborty V, Bordoloi M. Microwave-assisted oxidation of alcohols by pyridinium
chlorochromate. J Chem Res. 1999;2:118–119.
208. Chatel G, Monnier C, Kardos N, Voiro C, Andrioletti B, Draye M. Green, selective
and swift oxidation of cyclic alcohols to corresponding ketones. Appl Catal A Gen.
2014;478:157–164.
209. Matuszek K, Zawadzk P, Czardybon W, Chrobok A. New strategies for the synthesis
of lactones using peroxymonosulphate salts, ionic liquids and microwave or ultrasound
irradiation. New J Chem. 2014;38:237–241.
210. Beier MJ, Hansen TW, Grunwaldt JD. Selective liquid-phase oxidation of alcohols cat-
alyzed by a silver-based catalyst promoted by the presence of ceria. J Catal.
2009;266:320–330.
211. Zhu J, Faria JL, Figueiredo JL, Thomas A. Reaction mechanism of aerobic oxidation of
alcohols conducted on activated-carbon-supported cobalt oxide catalysts. Chem Eur J.
2011;17:7112–7117.
212. Sankar M, Nowicka E, Tiruvalam R, et al. Controlling the duality of the mechanism in
liquid-phase oxidation of benzyl alcohol catalysed by supported Au–Pd nanoparticles.
Chem Eur J. 2011;17:6524–6532.
ARTICLE IN PRESS

172 Maximilian N. Kopylovich et al.

213. Villa A, Wang D, Su DS, Prati L. New challenges in gold catalysis: bimetallic systems.
Catal Sci Technol. 2015;5:55–68.
214. Farsani MR, Yadollahi B. Synthesis, characterization and catalytic performance of a Fe
polyoxometalate/silica composite in the oxidation of alcohols with hydrogen peroxide.
J Mol Catal A Chem. 2014;392:8–15.
215. Chorghade R, Battilocchio C, Hawkins JM, Ley SV. Sustainable flow Oppenauer oxi-
dation of secondary benzylic alcohols with a heterogeneous zirconia catalyst. Org Lett.
2013;15:5698–5701.
216. Sudarsanam P, Mallesham B, Durgasri DN, Reddy BM. Physicochemical and catalytic
properties of nanosized Au/CeO2 catalysts for eco-friendly oxidation of benzyl alcohol.
J Ind Eng Chem. 2014;20:3115–3121.
217. Corberán VC, Gómez-Avilésa A, Martı́nez-Gonzáleza S, Ivanovab S, Domı́nguez MI,
González-Péreza ME. Heterogeneous selective oxidation of fatty alcohols: oxidation of
1-tetradecanol as a model substrate. Catal Today. 2014;238:49–53.
218. Tongsakul D, Nishimura S, Ebitani K. Platinum/gold alloy nanoparticles-supported
hydrotalcite catalyst for selective aerobic oxidation of polyols in base-free aqueous solu-
tion at room temperature. ACS Catal. 2013;3:2199–2207.
219. Mahdavi V, Hasheminasab HR. Vanadium phosphorus oxide catalyst promoted by
cobalt doping for mild oxidation of benzyl alcohol to benzaldehyde in the liquid phase.
Appl Catal A Gen. 2014;482:189–197.
220. Ozawa M, Fukutome A, Sannami Y, Kamitakahara H, Takano T. Preparation and
evaluation of the oxidation ability of hematin-appended 6-amino-6-deoxycellulose.
J Wood Chem Technol. 2014;34:262–272.
221. Dong X, Wang D, Li K, Zhen Y, Hu H, Xue G. Vanadium-substituted
heteropolyacids immobilized on amine-functionalized mesoporous MCM-41: a recy-
clable catalyst for selective oxidation of alcohols with H2O2. Mater Res Bull.
2014;57:210–220.
222. Patel A, Singh S. Undecatungstophosphate anchored to MCM-41: an ecofriendly and
efficient bifunctional solid catalyst for non-solvent liquid-phase oxidation as well as
esterification of benzyl alcohol. Microporous Mesoporous Mater. 2014;195:240–249.
223. Kim YH, Hwang SK, Kim JW, Lee YS. Zirconia-supported ruthenium catalyst for
efficient aerobic oxidation of alcohols to aldehydes. Ind Eng Chem Res. 2014;
53:12548–12552.
224. Parlett CMA, Keshwalla P, Wainwright SG, et al. Hierarchically ordered nanoporous
Pd/SBA-15 catalyst for the aerobic selective oxidation of sterically challenging allylic
alcohols. ACS Catal. 2013;3:2122–2129.
225. Parlett CMA, Bruce DW, Hondow NS, Newton MA, Lee AF, Wilson K. Mesoporous
silicas as versatile supports to tune the palladium-catalyzed selective aerobic oxidation of
allylic alcohols. ChemCatChem. 2013;5:939–950.
226. Mahyari M, Laeini MS, Shaabani A. Aqueous aerobic oxidation of alkyl arenes and
alcohols catalyzed by copper(II) phthalocyanine supported on three-dimensional
nitrogen-doped graphene at room temperature. Chem Commun. 2014;50:7855–7857.
227. Kumar A, Kumar VP, Kumar BP, Komandur VV, Chary VR. Vapor phase oxidation of
benzyl alcohol over gold nanoparticles supported on mesoporous TiO2. Catal Lett.
2014;144:1450–1459.
228. Rezaei B, Havakeshian E, Ensafi AA. Fabrication of a porous Pd film on nanoporous
stainless steel using galvanic replacement as a novel electrocatalyst/electrode design for
glycerol oxidation. Electrochim Acta. 2014;136:89–96.
229. Qiao ZA, Zhang P, Chai SH, et al. Lab-in-a-shell: encapsulating metal clusters for size
sieving catalysis. J Am Chem Soc. 2014;136:11260–11263.
230. Yan Y, Chen Y, Jia X, Yang Y. Palladium nanoparticles supported on organosilane-
functionalized carbon nanotube for solvent-free aerobic oxidation of benzyl alcohol.
Appl Catal B Environ. 2014;156–157:385–397.
ARTICLE IN PRESS

Catalytic Oxidation of Alcohols 173

231. Ishida T, Onumab Y, Kinjo K, et al. Preparation of microporous polymer-encapsulated


Pd nanoparticles and their catalytic performance for hydrogenation and oxidation.
Tetrahedron. 2014;70:6150–6155.
232. Gil S, Jiménez-Borja C, Martin-Campo J, Romero A, Valverde JL, Sánchez-Silva L.
Stabilizer effects on the synthesis of gold-containing microparticles. Application to the
liquid phase oxidation of glycerol. J Colloid Interface Sci. 2014;431:105–111.
233. Mahyari M, Shaabani A, Behbahani M, Bagheri A. Thiol-functionalized fructose-
derived nanoporous carbon as a support for gold nanoparticles and its application for
aerobic oxidation of alcohols in water. Appl Organomet Chem. 2014;28:576–583.
234. Parvulescu VI, Hardacre C. Catalysis in ionic liquids. Chem Rev. 2007;107:2615–2665.
235. Han X, Poliakoff M. Continuous reactions in supercritical carbon dioxide: problems,
solutions and possible ways forward. Chem Soc Rev. 2012;41:1428–1436.
236. Fall A, Sene M, Gaye M, Gomez G, Fall Y. Ionic liquid-supported TEMPO as catalyst
in the oxidation of alcohols to aldehydes and ketones. Tetrahedron Lett. 2010;
51:4501–4504.
237. Wang SS, Popovic Z, Wu HH, Liu Y. A homogeneous mixture composed of vanadate,
acid, and TEMPO functionalized ionic liquids for alcohol oxidation by H2O2.
ChemCatChem. 2011;3:1208–1213.
238. Zhang Y, Lu F, Cao X, Zhao J. Deep eutectic solvent supported TEMPO for oxidation
of alcohols. RSC Adv. 2014;4:40161–40169.
239. Wang Z, Cao XYJ, Lu M. Bi-functionalized PEG1000 ionic liquid ([Imim-PEG1000-
TEMPO][CuCl2]): an efficient and reusable catalytic system for solvent-free aerobic
oxidation of alcohols. New J Chem. 2014;38:4149–4154.
240. Movahed SK, Lehi NF, Dabiri M. Gold nanoparticle supported on ionic liquid mod-
ified graphene oxide as an efficient and recyclable catalyst for one-pot oxidative
A3-coupling reaction of benzyl alcohols. RSC Adv. 2014;4:42155–42158.
241. Lee SE, Ko JW, Ko WB. Synthesis of gold nanoparticles in presence of ionic liquid
[Bmin][CF3SO3] with humic acid under microwave irradiation and its application as
a nanocatalyst. Asian J Chem. 2013;25:9949–9995.
242. Li Z, Gong H, Mu T, Luan Y. Ionic liquid-assisted synthesis of unusual Pd particles
with enhanced electrocatalytic performance for ethanol and methanol oxidation.
CrystEngComm. 2014;16:4038–4044.
243. Dileep R, Bhat BR, Kumara THS. Palladium complex in a room temperature ionic
liquid: a convenient recyclable reagent for catalytic oxidation. Green Chem Lett Rev.
2014;7:32–36.
244. Chepaikin EG, Bezruchko AP, Menchikova GN, Moiseeva NI, Gekhman AE. Direct
catalytic oxidation of lower alkanes in ionic liquid media. Pet Chem. 2014; 545:374–381.
245. Chen L, Zhou T, Chen L, et al. Selective oxidation of cyclohexanol to cyclohexanone
in the ionic liquid 1-octyl-3-methylimidazolium chloride. Chem Commun. 2011;
47:9354–9356.
246. Saluja P, Magoo D, Khurana JM. Lanthanum triflate-catalyzed rapid oxidation of sec-
ondary alcohols using hydrogen peroxide urea adduct (UHP) in ionic liquid. Synth
Commun. 2014;44:800–806.
247. Paul CE, Lavandera I, Gotor-Fernandez V, Gotor V. Imidazolium-based ionic liquids
as non-conventional media for alcohol dehydrogenase-catalysed reactions. Top Catal.
2014;57:332–338.
248. Fan H, Yang Y, Song J, et al. One-pot sequential oxidation and aldol condensation
reactions of veratryl alcohol catalyzed by the Ru@ZIF-8+CuO/basic ionic liquid sys-
tem. Green Chem. 2014;16:600–604.
249. Schneider MS, Baiker A. Aerogels in catalysis. Catal Rev Sci Eng. 1995;37:515–556.
250. Caravati M, Grunwaldt JD, Baiker A. Solvent-modified supercritical CO2: a beneficial
medium for heterogeneously catalyzed oxidation reactions. Appl Catal A Gen. 2006;
50:298–303.
ARTICLE IN PRESS

174 Maximilian N. Kopylovich et al.

251. Sadri F, Ramazani A, Massoudi A, et al. Magnetic CoFe2O4 nanoparticles as an efficient


catalyst for the oxidation of alcohols to carbonyl compounds in the presence of oxone as
an oxidant. Bull Kor Chem Soc. 2014;35:2029–2032.
252. Ito Y, Ohta H, Yamada YMA, Enoki T, Uozumi Y. Bimetallic Co-Pd alloy
nanoparticles as magnetically recoverable catalysts for the aerobic oxidation of alcohols
in water. Tetrahedron. 2014;70:6146–6149.
253. Vinod CP, Wilson K, Lee AF. Recent advances in the heterogeneously catalysed aer-
obic selective oxidation of alcohols. J Chem Technol Biotechnol. 2011;86:161–171.

Anda mungkin juga menyukai